Karen Muskavitch

Commentary On

This case raises two very important questions with regard to research conducted in the collaborative setting of an academic laboratory: "Whose lab is it?" and the corollary "Whose research is it?" These questions are most pertinent when they concern research conducted by a post-doctoral fellow or a graduate student, as is the situation here. In the biology laboratories with which I am most familiar, the research of a graduate student like Archibald is typically the basis of his dissertation and in that sense is "his," but the work is part of a larger project on which the entire laboratory is working, and will continue working after he completes his degree and moves on. In this sense, the research is not his but rather belongs to the lab and the director of the lab. Most people are now aware that the research notebooks belong to the lab, and in many cases the convention is that the research questions stay in the original lab as well.

Many people are typically involved in a research project in an academic laboratory including the faculty member who is the principal investigator (PI) on most of the grants supporting the laboratory, a few post-docs trying to get their CVs in shape for the job market, some graduate students working toward their degrees, perhaps some undergraduates, and a few technicians. The technicians may range from those with advanced training in the field, even doctorates, to those who came to the lab with no special training and may only be able to carry out relatively routine tasks. Linking all these people is a complex web of relationships that can sometimes become strained or frayed.

This scenario focuses on one of these relationships, that between a graduate student and the faculty member who directs the laboratory. You will note that I have avoided using the term "mentor" to describe the faculty member. Contrary to what is usually assumed in the sciences, a graduate student's research adviser might not be the student's mentor. As noted in Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering, a recent book from the National Academy Press,

A fundamental difference between a mentor and an adviser is that mentoring is more than advising; mentoring is a personal as well as a professional relationship. An adviser might or might not be a mentor, depending on the quality of the relationship. . . Everyone benefits from having multiple mentors of diverse talents, ages, and personalities." (National Academy of Sciences et al., 1997, p. 15)

Because we do not know what the quality of the relationship between Archibald and Baker has been, I will simply use the term "research adviser" to describe Baker's relationship to Archibald.

Serving as a research adviser to a graduate student includes a number of responsibilities. (I will discuss the student's responsibilities in the commentary on Discussion Question 6.) These include guiding the student's research project by communicating effectively with the student, reviewing and providing regular feedback on the student's progress, and helping the student to acquire and develop the skills needed by independent researchers in their scientific field. In this case, we see that Archibald is meeting with Baker on a regular basis and that she reviews his work over the past week, looks at the primary data (not just the summaries that Archibald presents to her), and gives him concrete ideas on what to try next. This pattern of behavior is very good, and it seems to fulfill the first of the responsibilities of advisers. However, the way in which the conflict between Archibald and Baker is presented in this case leads one to wonder how well Baker has communicated the overall goal of the laboratory's research to her lab, and to Archibald in particular. He seems to be focused on the short-term goal of purifying cambin as quickly as possible and by whatever means so that he can do his experiments, write his dissertation and finish his degree. Baker, on the other hand, seems to be focused on the long term, on working with proteins purified in a unique way without the use of detergents. It is not clear whether Archibald just doesn't care about the long-term goals of the lab, or whether Baker has failed to communicate them to her collaborators. If the latter, then she has also failed to help Archibald to develop one of the skills he will need in a future career in science: the ability to see the big picture as well as determine the details of the next protocol that should be tested. In fact, her practice of making detailed notes in Archibald's notebook for what he should try next makes one wonder if she is doing too much directing of his work. What might be appropriate direction for a technician would not be appropriate for a senior graduate student who should be practicing experimental design skills. (See Discussion Question 3.)

As noted, we do not know what kind of relationship Archibald and Baker have had up to the exchanges recorded in this case study. What is likely to happen as this case is used with a group is that each person will project his/her own experiences and biases onto these two characters. That is good for the discussion if it engages the participants and helps them to reflect on their own relationships and what could be improved. However, it could be a problem if the participants start making assumptions about the personalities or motives of these two characters and then base their ethical analyses on these assumptions. We don't know if Baker is a long-suffering junior faculty member working with a graduate student who can't seem to see beyond his own dissertation, or if Archibald is a bright, motivated graduate student struggling under an adviser who doesn't tell lab members what the overall plan is and who wants to control every aspect of every experiment run in her lab. In facilitating discussion of this case, I suggest taking a neutral view of both characters. Assume that they are acting in good faith, and beware of assumptions that discussion participants may be making. However, the discussion also should explore the possible differences if we assume that Baker is a micro-manager or Archibald a short-sighted student. The possible consequences of a proposed course of action might change, but usually the affected parties' rights and interests, and the ethical principles and obligations, do not.

Some people might question whether the conflicts presented in this case aren't more issues of etiquette than of ethics. Because they deal with how people ought to treat each other, they are ethical issues. Many scientific societies and writers in the field of research ethics have argued that the treatment of graduate students is an issue in research ethics. In their report Responsible Science: Ensuring the Integrity of the Research Process, an NAS committee includes "[i]nadequately supervising research subordinates or exploiting them" among questionable research practices, "actions which violate traditional values of the research enterprise and that may be detrimental to the research process." In describing best practices, they note that "[s]cientists in universities accept the obligation to pass along knowledge and skills to the next generation of scientists," and that "[t]he mentor has the responsibility to supervise the trainee's progress closely and to interact personally with the trainee on a regular basis in such as way as to make the training experience a meaningful one." (National Academy of Sciences, 1992, Vol. 1, pp. 28, 141-42) Weil and Arzbaecher assert that with regard to relationships within research groups going sour "[w]e can collect these ways of going astray under broader ethical questions about how to wield power responsibly and how to behave responsibly as one dependent on the power of others. As we proceed to point out the kinds of standards and practices that are needed, we thereby delineate role responsibilities in research groups. To fail to fulfill these role responsibilities would be to behave irresponsibly, that is, unethically." (Weil and Arzbaecher, 1997, p. 78)

Discussion Questions

Questions 1 and 2

Baker's reaction to Archibald's announcement that he had gone ahead and tried the CTAB indicates that there may have been a better way either to go about the experiment, or to tell Baker about it. That does not mean that doing the experiment was "wrong." Archibald was not squandering significant laboratory resources or endangering other members of the lab, and he did try Baker's suggestion first. He was trying something that others had used with success but which Baker had told him not to do. It is not clear why she told him not to try CTAB. Was it because she wanted to control every detail of work in the lab, or because protein purified in the presence of a detergent like CTAB was worthless for their research? It is unreasonable to expect that an adviser should okay the details of everything a graduate student does. However, Archibald could have done things a little differently and possibly avoided Baker's angry response. For instance, he could have asked Baker earlier for a clarification as to why she opposed his testing CTAB. The ensuing discussion might have led to some sort of understanding. Or he could have presented the results differently. Instead of announcing the wonderful purification as he did, he could have started by describing how he carefully tried all Baker's suggestions and then decided to try CTAB while he was at it. He could have told Baker, "I know that protein purified with detergent is not useful for our studies, but I was starting to wonder if active cambin could be purified at all. At least I now know that it is possible, and we just have to figure out how to keep it active in the absence of detergent." He needs to respect his adviser-student relationship with Baker, but he must also remember that he is part of a research team and not just a pair of hands.

Archibald's chosen course of action, although not "wrong," probably was not the best choice. The tone of Baker's response, however, was clearly inappropriate and arguably "wrong." She responded as if she were scolding a child, not talking with a junior colleague in the presence of other members of the lab. (Recall that the setting is a lab meeting; we can assume others are present even if we do not hear from them.) In addition, it would take a very special set of circumstances to justify the command to a graduate student that he "never conduct experiments without my explicit approval!" It might be justified if he were a first year student just starting in research, or if he had a history of endangering others or wasting time and materials on poorly designed, inconclusive experiments. However, the essence of science is exploration and discovery: To deny a student the opportunity to try some of his own ideas is to deny him the opportunity to develop into a mature scientific investigator.

Question 3

This question asks whether Baker has the authority to control all experiments in her laboratory. For a number of reasons, the ultimate answer is "yes." I would add several qualifiers, however: that she should include others in her decision making, and that she should be sure to provide opportunities for graduate students and post-docs to participate in the decision-making process as a part of their training. However, she is the one held responsible for the funds granted to the lab, for the safety of all in the lab, for the validity of work published by the lab, and for the lab's progress in its research. Therefore, she does and must have final authority for what is done in her name in her laboratory.

Although she has the authority, that does not give her the right to act in a dictatorial or arbitrary manner. In addition, the different types of researchers in her laboratory need to have different amounts of freedom in their design of experimental approaches. A post-doc is like an apprentice scientist, just one step away from independent research and often the recipient of a stipendiary grant and funds for research materials. However, the post-doc usually has received the grant to do a certain project in a certain lab and is still considered a trainee. Thus, some guidance and supervision is appropriate. At the other end of the spectrum is the relatively unskilled technician who follows protocols prepared by others and may not even participate in the interpretation of the data collected. Between the post-doc apprentice and the hired hands of the technician is the graduate student. As part of their training, graduate students must be part of the experimental design process so that they can learn and develop their skills. The level of their participation should increase over time as they complete their graduate work. Thus, the level of faculty guidance given to a first-year student would not be appropriate for a fourth-year student. However, a completely hands-off style is never appropriate for reasons of student training and faculty responsibility.

Deciding whether Archibald's committee needs to be informed about this incident requires that we know if it was an isolated occurrence on a particularly bad Monday morning, or if it represents a pattern of micro-management and dictatorial behavior by Baker toward Archibald. Archibald could experience negative consequences if he takes this conflict outside the lab, even if it is to his dissertation committee. Thus, he must weigh his options carefully, and, if possible, unemotionally. If this incident does represent a pattern, then Archibald should go to the dissertation committee to seek redress of a situation in which he, and possibly other students in the lab, is not being trained as a predoctoral student should be.

Questions 4 and 5

No level of pressure of any type on Baker would justify a disrespectful and dictatorial response to a graduate student. However, because of the fact that she is responsible for the use of grant funds and for the reasons mentioned in the comments above, Baker does have the authority and responsibility to oversee the experiments carried out in her laboratory. She needs to change the way in which she exerts this authority.

We often hear people say that the pressures of contemporary science justify inappropriate actions, even fraud. "Pressure" is not a valid ethical factor. True, we do need to be cognizant of the pressures confronting us and try to reduce them if possible, but we can't use them to excuse inappropriate actions. The pressure on a junior faculty member to secure continued funding is not only related to getting tenure. It also involves concerns about having enough money to keep members of the lab employed, maintain student support, and be able to pay the bills for expensive reagents so that all can do their experiments. Baker may see the use of a unique, detergent-free purification for the proteins studied in the lab as the hook that will secure the continued funding, but she needs to explain her reasoning to others in her lab so that they will understand and learn from her.

Question 6

In discussions of cases like this one, we frequently spend a lot of time talking about the rights of graduate students, probably because these rights are often disregarded. However, it is also important to explore the other side -- the responsibilities of students. After all, education is not a passive endeavor. In this case, we learn that Archibald has been reading papers describing purification protocols similar to his own, and that is exactly what he should be doing. But I am puzzled as to why he does not understand the significance of the detergent-free protocol used in the Baker lab. From the information given in the case, it is not clear if the fault for this lapse lies primarily with Archibald or Baker. Has Baker failed to be clear or forthcoming with her reasons? Has Archibald failed to ask, or has he failed to pay attention to Baker's answers? We don't know, but both must bear some of the blame for the situation.

It has been noted that "The term 'mentoring' refers to an interactive process; The role of the mentored person is not a passive one. That person has a responsibility to seek information and guidance and to be ready to make use of it." (Weil and Arzbaecher, 1997, p. 77) A student should be open to, and even seek out, additional information and the perspectives of others, particularly those who are more experienced. Then the student should develop a reasoned position of his/her own to contribute to the discussion. In the end, it is hoped that student and adviser will arrive at a consensus as to how to proceed; failing that, however, the authority of the adviser who is head of the lab must be respected. This situation differs from that in History, for instance, where students typically work independently of all others in libraries or archives, and the dissertation adviser may not be a coauthor on any work that is published. But all graduate students should acknowledge the greater experience of their adviser and the fact that they asked this faculty member to guide their work, and so act on their adviser's suggestions or at the very least give them serious consideration.

Question 7

Consideration of the two principal questions raised here will probably be the most valuable part of the discussion of this case. How could this situation have been avoided? And what should Archibald and Baker do in the future? As noted above, it is not clear who bears the greatest share of blame for the current situation, nor do we know what Archibald and Baker's previous relationship has been like. Therefore, there are no definitive answers to these questions. Rather, they serve to help all of us to consider how to improve communication and thus relationships within our own research groups. Brainstorming and sharing ideas and experiences will be very helpful if coupled with an evaluation of what is likely to be most successful in a given situation.

I offer two suggestions. It would be helpful if there were an opportunity for members of a research group to discuss their expectations of each other before a crisis occurs. Perhaps this case or the vignette entitled "The Lab of Last Resorts" (Weil and Arzbaecher 1997, p. 79) could be used to trigger the discussion. Baker's lab and others also could benefit from more discussion of the "big picture" by the lab director so that all would know how their work fits together into a whole. This orientation could be provided through regular presentations by the director at lab meetings, or by cooperative preparation of grant applications.

References

  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering. Washington, D. C.: National Academy Press, 1997. Available online at http://www.nap.edu/readingroom/books/mentor.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Responsible Science: Ensuring the Integrity of the Research Process, Vol. 1. Washington, D. C.: National Academy Press, 1992.
  • V. Weil and R. Arzbaecher. "Relationships in Laboratories and Research Communities" in D. Elliott and J. E. Stern. Research Ethics: A Reader. Hanover, N.H.: University Press of New England, 1997.

This case raises a number of issues concerning the use of animals in biomedical research and the reactions of people who interact with these animals. In this commentary, I will consider the general issues related to the use of animals in research first and then turn to the more specific issues concerning Frank.

Whether or how animals should be used in research has been an issue of great debate for a number of years, a debate that shifts as our knowledge and understanding of animals grows. Many nineteenth century biomedical scientists viewed animals much as Descartes did, as similar to machines and incapable of feeling emotion or pain. "[T]hey interpreted the cries of an animal during vivisection as the mere creaking of the animal 'clockwork'" (Rudacille, 2000). Today, most researchers are concerned about the welfare of their animals and willingly comply with rules and regulations. They consider alternatives, the three R's of replacement, reduction and refinement (Russell and Burch, 1959) when preparing research proposals, but they also wish that IACUC review didn't take so long. The evolution in the way in which researchers view and treat their animal subjects has come as a result of our increasing knowledge about animals and their lives, as well as through interactions with what has come to be known as the animal rights movement (Orlans, 1993 and Rudacille, 2000). Our ways of thinking about the moral status of nonhuman animals have also changed over time and the lively debate currently includes, among others, those who would ascribe rights to nonhuman animals because they have inherent value since they are "subjects of a life," and others who argue that nonhuman animals do not have any rights per se and that more, rather than less, biomedical research using animals should be done because of its benefit to the human community (Orlans, 1993 and Orlans et al., 1998).

This case, because it deals with chimpanzees, has an additional layer of complexity and controversy. There are currently no nonprimate animal models for AIDS research, and there is a movement to include all great apes, human and nonhuman, in a community of equals. This movement is the Great Ape Project and is best described by its Declaration on Great Apes (Great Ape Project):

We demand the extension of the community of equals to include all great apes: human beings, chimpanzees, gorillas and orang-utans. The "community of equals" is the moral community within which we accept certain basic moral principles or rights as governing our relations with each other and enforceable by law. Among these principles or rights are the following: 1. The Right to Life . . . 2. The Protection of Individual Liberty . . . 3. The Prohibition of Torture.

At the same time that our more detailed knowledge of the life history and behavior of nonhuman great apes is leading some of the scientists who study them, such as Jane Goodall, to call for an end to their captivity and use in experiments, we also have the AIDS pandemic affecting millions of human beings. Chimpanzees are the only nonhuman animals that can be infected with HIV, and they will also eventually develop AIDS. While monkeys will become ill if infected with a simian version of HIV, it is not clear how analogous this disease is to HIV-caused AIDS. Thus, chimpanzees are considered the best candidates for a nonhuman animal model in which to learn more about disease progression and test potential AIDS vaccines.

Issues that are specific to Frank's situation are of two types: those that concern ethics, and those that concern his goals and feelings. Ethical concerns include the general issues discussed previously in this commentary and specific issues such as whether Vern's housing and medical care meet accepted standards. It is understandable that a chimpanzee infected with HIV has been isolated from other chimpanzees, but Frank might ask if steps have been taken to enrich Vern's environment and provide other social interactions, perhaps with suitably protected humans. Frank might also question if enough is being done to minimize Vern's pain and suffering. If these are concerns, Frank should raise these specific issues as well as the general ones concerning the use of chimpanzees in biomedical research. However, to be effective he needs to do so in a nonaccusatory, questioning manner, and he may need to seek out further information from other sources to educate himself on these issues. Respectful dialogue about the use of nonhuman animals in research should be part of the culture in a facility such as the one described in this case. If there are major problems with the way in which the chimpanzees are treated at the facility and Frank cannot get his supervisors to take his concerns seriously, he may need to alert people higher in the organization. If he believes that the work done at the facility is immoral, although it is in compliance with current animal use regulations, he may need to quit his job, and possibly work to change others' opinions.

This case also raises concerns related to Frank's goals and feelings. He began work at this animal facility because he was interested in doing graduate work in immunobiology. By the end of the case, he is no longer interested in work in this field, presumably because he is uncomfortable with the costs of this research to the animals. This need not be an ethical issue, but can be one having to do with personal emotions and preferences. Many of us in academic research chose graduate school over medical school because we realized that we were not comfortable dealing with people who are sick and/or in pain. Similarly, I know several biologists who work in plant rather than animal systems because they are not comfortable dissecting or drawing blood from animals. These people are not morally opposed to animal research; they are not vegetarians, but this type of work is not for them. Choosing a field of research involves finding a niche where one is excited by the research questions and is also comfortable with the techniques employed. That is not usually an ethical issue, but one of personal interests and aptitudes.

With all the complexities involved in this case, its discussion would benefit from preparatory research by the discussion participants into such topics as views on the moral status of animals, current regulations concerning care of primates used in research, model systems used in AIDS research, and the actual case of the first chimpanzee to develop AIDS (Novembre et al., 1997).

References

  • The Great Ape Project "A Declaration on Great Apes" Online. http://www.greatapeproject.org/gapintroduction.html, January 2002.
  • Novembre, F.J., M. Saucier, D.C. Anderson, S.A. Klumpp, S.P. Oneill, C.R. Brown, C.E. Hart, P.C. Guenthner, R.B. Swenson, and H.M. McClure. "Development of AIDS in a Chimpanzee Infected with Human Immunodeficiency Virus Type1." Journal of Virology 71 (5, 1997): 4086-4091.
  • Orlans, F. Barbara. In the Name of Science: Issues in Responsible Animal Experimentation. New York: Oxford University Press, 1993.
  • Orlans, F. Barbara, Tom L. Beauchamp, Rebecca Dresser, David B. Morton and John P. Gluck. "Moral Issues about Animals" in The Human Use of Animals: Case Studies in Ethical Choice. New York: Oxford University Press, 1998.
  • Rudacille, Deborah. The Scalpel and the Butterfly: The War between Animal Research and Animal Protection. New York: Farrar, Straus, and Giroux, New York, p. 21.
  • Russell, William M.S, and Rex L. Burch. The Principles of Humane Experimental Technique. London: Methuen and Company, 1959.
Commentary On

This case describes a tricky situation that is resolved well, but could have gone very wrong. The positive resolution was possible because all parties acted in a mature manner, communicated openly and respectfully with each other, and accepted responsibility as their obligations dictated.

Field studies -- especially long-term studies -- can be difficult and plagued with unexpected complications. Discussion of this case could help all those concerned with such studies to anticipate problems that might arise, both in experimental design and in interpersonal relationships, and how they might be resolved.

The actions of two characters in this scenario, Jane and Professor Maple, deserve further comment. Jane recognized the problem with determining the locations of the plots, welcomed others' suggestions (Anastasia's, for example), sought further information on her own (called Dr. Ilex), but then discussed the situation with Maple. She demonstrated an ability and willingness to act and make decisions independently, as a graduate student should, as well as to seek out information and advice when needed, also important for a grad student. When she could not arrive at a clear solution, she did not waste time worrying about the problems, nor did she minimize or walk away from them. Rather she presented them clearly and respectfully to Maple, along with her concerns about her graduate career.

Some graduate students in her situation might have just continued with the research hoping it would all work out in the end, or they might have made up some excuse for wanting to work on a different project with another professor. Others might have loudly accused Maple of lax supervision of previous students, incompetence, or even misconduct. Jane chose a better course of action: respectfully asking Maple what he thinks is the best thing to do while asserting her legitimate interests in completing her degree in a reasonable amount of time with a scientifically valid thesis.

Similarly, Maple responded to the situation very well. He did not dismiss Jane's concerns with some comment like, "Look, I've been doing field studies for longer than you've been alive. Who are you to question where I say the plots are, or my previous students' work?" Rather, he deliberated with Jane, and together they devised a plan to bring in a surveyor, a skilled third party, to mark smaller uniform plots so that Jane and her team could complete the final set of field measurements; arranged to gather measurements so that Maple would be able to evaluate possible amounts of error; and agreed on a research question for Jane's thesis work that was limited to data that she could evaluate and take responsibility for.

Maple, in turn, assumed responsibility for reviewing the data and conclusions that were derived from this study site over 40 years. Presumably he would inform the scientific community if he discovered significant errors that had the potential to invalidate previously published work. Determining what constituted significant error and how best to communicate this information to the scientific community would require judgment calls on Maple's part. His willingness to deliberate with a first year graduate student on the best way to complete his 40-year study suggests that he would make these decisions in consultation with others and in a manner that showed an awareness for his responsibilities to those who will refer to his publications, and a greater concern for the facts than for his pride - admirable characteristics in a scientist.

This case should trigger a good discussion on the importance of effective communication and respectful relationships within a research group, how to achieve such a desirable lab climate, and what to do if things go wrong. It has been my experience that we scientists don't talk very much about interpersonal relationships within our research groups and have little to no training in how to facilitate effective communication. After all, we did not go into scientific research because we were "people persons." Unfortunately, although good research can be done in an unpleasant lab climate, poor communication and relationships usually translate into time and energy wasted on bitter thoughts and feelings, plus opportunities missed because of a lack of cooperation and sharing of ideas. We wouldn't tolerate such a situation if the problem were lack of skill in an experimental technique; we would seek out advice, information or training. Dealing with an interpersonal conflict is the last thing most scientists want to do, but as we can see from this case, we can't always just pretend the problem doesn't exist.

How does one facilitate effective, respectful communication within a research group? Someone with training in management or psychology might be better suited to answer that question. I know of one institutional program designed to help academics learn effective, respectful communications skills (Klomparens and Beck, 2002), which could also be incorporated as one of the goals of an educational program on the responsible conduct of research (RCR).

In addition, I have some observations and suggestions from my years in the lab. One requirement for good communication is regular meetings: of the entire group, of individuals with the faculty PI, and of small collaborative working groups, if these exist. These meetings should not only feature formal presentations and reports, but should include a lively discussion and questioning of the material presented as well as an exchange of ideas. Meetings of the entire group might sometimes include discussion of cases, like this one. They are also the perfect forum for the discussion of "lab business" concerning policies on keeping the darkroom clean, the proper place and state in which to return the Geiger counter, and the like. After meetings of small working groups, it may be desirable to write up and circulate among the collaborators a record of what was decided at the meetings as a way to avoid misunderstandings and perceived territorial infringements. Regularly scheduled individual meetings enable the PI to keep in touch with what everyone is doing and give lab members a chance to voice concerns without having to make a big thing of it by asking for a special meeting with to PI. Regular meetings help people to become accustomed to talking with each other and understand that the exchange of information is one of the expectations of the research group. Regular meetings also ease informal, day to day communication as people get to know each other better. Informal communication can be further facilitated by nonscientific gatherings of the group (going out for lunch, for instance) and by the PI's frequent participation in the daily discourse of the lab. Just coming in regularly for a morning cup of coffee and some conversation will help the PI keep in touch with the real climate of the lab. Yes, we're all too busy, but regular informal interactions are essential. The most important thing that PIs can do to promote a good climate s in their research groups is to model open, clear, respectful communication with all members of their groups at all times. Monson's decision to privately ask Paul to begin to manage the daily activities of the lab, and then not to discuss this decision or even make it known to the other members of the lab clearly did not help the climate in his lab.

What to do when things go wrong within a research group? The best course is to intervene early. That doesn't mean that you don't give people a little space for occasionally having a bad day, or forgetting to do something they ought. But one shouldn't wait too long to deal with repeated lapses or a simmering tension. A climate where questions were accepted, even expected, would have helped a lot in the case presented here. However, no one wants to bother Monson, nor can they talk with each other in anything other than an accusatory tone. Richard, an undergrad and thus quite low in the lab pecking order, feels that his only recourse is to spend as little time in the lab as possible. If someone had alerted Monson, or if he had spent enough time with the people in his group to realize that something was not right, the critical incident might have been averted. It might also have helped to have a mechanism to defuse the tension when Lisa began leaving equipment dirty. (See Commentary on "The Rat Race," p. 53.)

Discussing what Richard should do after he observes Paul's suspicious behavior is a good opportunity to practice imagination and moral reasoning. First the group can brainstorm all sorts of things Richard might do, and then they can be called upon to determine which option they judge to be the best and explain their reasoning. Brainstorming is a good exercise because people in difficult situations frequently think that they have only two or three options. It takes imagination to find creative middle ground. In Richard's case, he might: call the city police, call Laboratory Safety and leave a message, confront Paul one-on-one in the lab that very evening, do nothing, put a radioactive hazard label on Lisa's chair and walk out, call Monson, call another faculty member, call the New York Times, and so on. There are lots of possibilities if you separate the imaginative from the evaluative process.

After brainstorming, the discussants need to evaluate the options generated, a process usually referred to as moral reasoning. Several guides on how to teach and practice this skill are available (see, for instance, Bebeau, et al., 1995 or Elliott and Stern, 1997). One needs to consider and then balance the moral and legal obligations of the protagonist, the other people who might be affected by his actions and their interests, and the possible consequences of different courses of action. Frequently, it is easiest to start by eliminating the possible courses of action that are unacceptable, being sure to explain one's reasoning, and then move on to consideration of the relative value of the remaining, acceptable options. Just as there is more than one way to build a bridge across a river, there are usually several acceptable options. The discussants may disagree on which is the "best" course of action for Richard to follow because they may weight different obligations or interests differently, and that is fine if their reasoning is sound. There is usually no single right answer to such problems.

In this scenario, Richard chooses the acceptable option of alerting Lisa. One option that is not acceptable is that he do nothing. After all, the contamination threatens not just Lisa, but others who might come into the lab such as the janitorial staff or another grad student who drops by to talk. One could argue that alerting Monson might have been a better course of action. Since it is his lab that is affected, he might be better able to shield Richard from any fallout, and the license for use of radioactive materials is in Monson's name. This is a good topic for discussion.

References

  • Bebeau, M.J., K.D. Pimple, K.M.T. Muskavitch, S.L.Borden, and D.H. Smith. Moral Reasoning in Scientific Research, Cases for Teaching and Assessment 1995. http://www.indiana.edu/~poynter/mr main.html.
  • Elliott, Deni, and Judy E. Stern. Research Ethics, A Reader. Hanover, N.H.: University Press of New England, 1997.
  • Klomparens, Karen, and John Beck. "Conflict Resolution." The Graduate School, Michigan State University. http://grad.msu.edu/conflict.htm.
Commentary On

These days, we most frequently think of research ethics as fabrication, falsification and plagiarism (FF&P). This case, however, deals with ethics in the more general sense of a concern for what one ought to do. Here we are asked to consider what Cindy ought to do in her interactions with others in the laboratory, and also what Tom, the PI of the lab, ought to do in managing and leading his research group.

Before moving to consider what these people ought to do to in this laboratory, I want to note two reasons why consideration of this case is not so far removed from FF&P as one might initially suppose. First, the quality of intralab communications and relationships is a major contributor to the climate in the lab, as we see in this case. Climate, in turn, has been found to correlate with the frequency of perceived misconduct. Researchers have observed that in departments where the climate emphasizes competition and individualism over cooperation there is a higher probability that students will report observing misconduct committed by others in their department (Anderson, et al. 1994). Second, some have suggested that "significant misbehavior that . . . intentionally impedes the progress of research" should be included in the definition of research misconduct (Commission on Research Integrity, 1995), and Beth's actions with the joystick, if intentional, might be seen by some to cross over into sabotage and misconduct.

The more general ethical question of how Cindy and Tom, and by association others in the lab, ought to interact hinges on respect. We are accustomed to thinking about respect for persons as critical for the determination of ethical conduct in research involving human subjects, but we frequently forget that that is just a special case of the general requirement that we as ethical people treat others as autonomous individuals capable of making informed choices. Immanuel Kant wrote that one ought to "[a]ct so that you treat humanity . . . always as an end and never as a means only" (Kant as quoted in Rachels, 1993). That is, we are not to manipulate people to achieve our own ends, and the primary way in which people manipulate others is through incomplete or biased information. Respect in interpersonal relationships requires clear and honest communication by all parties.

It appears that Cindy and Beth try to have good communication, although belatedly, with regard to the work on the research proposal. They manage over time to come to the conclusion that "the misunderstanding had been a result of poor communication." However, they do not seem to try to avoid future problems by making sure that the lines of communication are more open. Instead, when another situation arises two months later, Cindy confronts Beth with an accusation of sabotage, rather than seeking to understand what might have happened. If it had been a problem with an experimental technique, Cindy and Beth would probably have sought out advice and training, probably from Tom first. That is what they should do in this situation too. Unfortunately, they don't view their communication problems in the same light, and it appears that nothing is done.

Tom, the professor who is the head of the lab, may believe that he is respecting the independence of the members of his research group by allowing them to solve their problems without his interference, but what he is really doing is neglecting his responsibility to train these students in the skills needed to work in the larger research groups we see today. One assumes that he would not take such a hands-off approach if they were deficient in experimental or analytical skills, but interpersonal skills are also critical for collaborative research. Granted, most professors did not go into academic research because they discovered that they were people persons, and very few have training in management, but there are some things that can be done. The most important is to promote clear and honest communication, even if it is uncomfortable at times. Sign up sheets are a good start, but facilitating timely communication among lab members is also essential. He should have made sure that both Beth and Cindy knew what he had requested of each of them before they "met" in the lab that night to work on the grant application. Then, when Cindy brought up the problem with Beth and the joystick, it was his responsibility to step in if the two could not work things out. The fact that the two are not talking to each other and that the lab climate has changed are clear indications that it is time for Tom to step in. If nothing else, he owes it to the others in the lab, who did not cause the friction but are affected by it. Tom may need to go so far as to force Cindy and Beth and/or the whole lab to sit down, talk about the situation, and figure out a way in which the lab members can work together, perhaps with outside help.

How might the situation described at the end of this case have been prevented? Having in place a lab culture that values and facilitates open communication would help a great deal. Tom would need to make it a habit to meet with lab members individually or in working groups, and he would need to be present in the lab now and then to see how things were going. In order to prevent misunderstandings on "territory," I have known some research groups to draft written minutes of working group meetings so that all are clear on who is doing what. Discussion of cases like this one at lab group meetings can also help. It gives everyone a chance to consider a hypothetical situation in a less emotionally charged environment while also learning the group's expectations for appropriate behavior before a crisis. Then there is the "doughnut penalty," a mechanism for defusing the animosity that can be generated by an error by one lab member that affects the work of others. If someone mistakenly leaves a piece of equipment in a place where others can not find it and so impedes their research, for instance, the offender is required to make the situation right and then bring in doughnuts, or a similar treat, for the lab the next day. In this way the offender makes a type of public confession and penance, and all can then move forward with less danger of bitter feelings.

References

  • Anderson, M.S., K. Seashore Louis, and J. Earle. "Disciplinary and Departmental Effects on Observations of Faculty and Graduate Student Misconduct," Journal of Higher Education 65 (3, 1994): 331-350.
  • Commission on Research Integrity. Integrity and Misconduct in Research. Washington, D. C.: U.S. Department of Health and Human Services, Public Health Service, 1995.
  • Rachels, J. The Elements of Moral Philosophy. 2d edition. New York: McGraw-Hill, 1993, p. 128.

This case is likely to prompt those who use it to reflect on all three ethical principles put forth in the Belmont Report1 - respect for persons, beneficence and justice - and to consider how well they are addressed in the research described. However, the focus of the case is on the first principle, treating people as autonomous agents, and its application in the process of informed consent.

The right of potential research subjects to choose for themselves in a free and informed manner is central to all of the guidelines that have been written for research involving human subjects such as the Nuremberg Code,2 the Declaration of Helsinki,3 the Belmont Report1 and the CIOMS International Ethical Guidelines for Biomedical Research Involving Human Subjects.4 These ethical guidelines have then been used as the basis for regulations and more applied guidelines that can be used to evaluate proposed research protocols.5 -7

Ellen's situation is particularly interesting because all the basic ethical requirements for international collaborative research of the type described seem to have been met. Yet Ellen, sitting in on subject recruitment, feels that all is not as it should or could be. Yes, approvals have been received from both the PI's American university and the local, collaborating university. These approvals presumably came from the universities' human subjects committees, duly constituted to be aware of the applicable regulations and of the local cultural norms. These committees have approved the consent process and forms that are being used and, as is recommended for such research7, the community elders have been consulted and have given their consent. The potential risks associated with the experimental treatment are quite minimal, and the researchers are providing benefits to the research subjects in the form of obstetric and gynecologic exams and treatments, and to the entire community by building a health care facility that will remain after the end of the study. In addition, this is not a case of Western researchers coming in to test a treatment that those in the subject pool could never afford. Rather, the experimental treatment that is being tested is one that could benefit this community and others just like it. Despite all of these facts, however, Ellen is uncomfortable.

The questions in this case go beyond simple compliance with regulations, as important as that is, to considering congruence with the spirit of central ethical principles such as respect for persons.

Informed consent includes three elements: "information, comprehension and voluntariness".1 It seems that Tefera is reading to Sebena all the relevant information about the research study, but Ellen has concerns about Sebena's ability to understand the information she is being given because of her cultural background. In addition, Ellen questions the voluntariness of Sebena's offer to put her mark on the consent form. Voluntariness requires that the consent be "free of coercion and undue influence".1 Is there an element of coercion? Is Sebena simply obeying the perceived authority of the nurse and the researchers, or the community elders?8 (pp. 68-69) Or do the benefits of the research constitute an undue influence? Does Sebena feel an obligation to consent to be a research subject because of what the researchers have done for her community, or because of the medical care they promise for her and her unborn child if she joins the study? Since she cannot talk directly to Sebena, it is impossible for Ellen to answer these questions immediately, but SebenaÀs body language tells Ellen that there is cause for concern.

In following up on her concerns, it is important for Ellen to be aware that applying Western ethical standards to other cultures is not as simple as it might initially seem. The idea of autonomy is based on a Western understanding of individualism. In other cultures, the community is central to oneÀs life, and the individual is much less important. Important decisions may be made by the community as a whole or its elders, rather than by the individuals directly involved. This locus of decision making can result in an ideological conflict for an ethically concerned researcher such as Ellen. On one hand, it seems a good thing to respect other cultures and avoid coming in as an outsider and challenging the traditional authority structure of these communities. However, there is also good in working to empower less well represented individuals in these communities and sharing with them a different world view in which they as individuals have inherent worth. This tension can be particularly acute if the less powerful individuals are women and the community elders are all male. Some of these concerns and tensions are discussed in an article by Dena Davis.9

What should Ellen do? First she needs to creatively consider the wide variety of options that may be available to her. In a discussion of this case, this is a point where brainstorming possible options could be very productive. Then the discussion can move to evaluation of the various courses of action suggested. Most would agree that Ellen should do something, and similarly that she should be wary of being too righteous or heavy handed in her approach. One good approach is to spend a lot of time asking questions.10 In this way she could learn more and perhaps cause others to acknowledge their own uncertainties and/or consider the questions Ellen raises.

If there really is a problem and Ellen can gain support from those with more authority in the study, then some changes could be made. Perhaps the vocabulary used in the consent process could be revised so that it is more consistent with common usage in the town. Any proposed changes would need to be approved by the appropriate IRBs,5,7 and it is important to avoid misunderstandings such as those that resulted from the use of the term "bad blood" in the Tuskegee study. 8 (pp. 71-73) Perhaps there could be a program of community education in current medical concepts such as disease, treatment and research. This strategy would be consistent with the understanding that informed consent is an ongoing process of communication between subject and researcher, not a one-time signature on a form. Perhaps the consent process for the individual women could be moved out of the clinic and into a more familiar setting. These are just a few suggestions.

A final note: Just as our awareness and understanding of ethical issues associated with international and cross-cultural research are developing, so are the associated guidelines and regulations.11 Therefore, what might be considered to be the best course of action now may not be evaluated in the same way several years later.

Footnotes

Commentary On

Experimental Design of the Study

The Informed Consent Process

The Dissertation Committee


The primary questions in this case cluster around beneficence, the ethical principle that one should avoid doing harm and seek to do good. How one answers these questions will then affect how one designs the informed consent process for this study. A secondary set of questions concerns the dynamics between a graduate student and the faculty on her dissertation committee.

Experimental Design of the Study

Conducting a risk/benefit analysis is more than just a required task that a researcher must perform in filling out an application for IRB approval. Coming from the ethical principle of beneficence in the Belmont Report,(1)  it is the obligation of researchers to scrutinize the experimental design of their proposed studies, and to put themselves in the proposed subjects' shoes to consider what potential harms and benefits may result from their work. If one looks at the mandated criteria for IRB approval of research,(2)  one notes an emphasis on evaluating potential risks and benefits of a proposed study in the context of its research design. While the federal regulations do not clearly require IRBs to review a proposed study's experimental design,(3)  some institutions do ask this review of their IRBs (see, for example,(4) ) and, as in this case, it frequently becomes important in the consideration of potential risks and benefits.

A number of questions need to be answered before a decision can be made about making the results of the hCG analyses available to the subjects. First, what is the goal of this study? What is the hypothesis being tested? What does Wilma hope to learn from this study and then contribute to the scientific literature? The case as written offers no clear answers to these questions, and so these points may need to be addressed in the case discussion.

Then, knowing the study's goal(s), one can begin considering possible alternatives and refinements to the experimental design. Who are the subjects of this study? What, if anything, will be presented to them as the potential benefits of participation in this study? How will potential subjects be recruited or excluded? Will the men in each couple answer the questionnaire and consequently be subjects, or will only the women fill out the questionnaire as well as providing urine samples? Will the nurses gather any additional information when they collect the month's samples? Will Wilma contact the subjects for any follow-up data, such as reports of pregnancies clinically confirmed within three months of the completion of the study? All of these aspects of the experimental design affect the potential risks and benefits to the subjects and thus affect how Wilma will design the informed consent process and form. Let me give a couple of examples.

If only women are recruited, complete the questionnaire and provide samples, then only they are subjects of the research, and their male partners have no claim upon data generated by the study. However, possible harms and benefits to the male partners who are not subjects should also be considered in the research design. Stakeholders beyond the immediate subjects of the research, such as the subjects' families and communities, should also have their interests considered and protected.

If Wilma recruits from among patients of clinics specializing in the treatment of infertility, then her subjects, and possibly their doctors as well, will be very interested in obtaining their hCG data even six to twelve months later. However, if she recruits from among the general population and restricts her study to women who have been trying to conceive for less than three months, she is better able to argue that there is very little benefit in making the data available. Nevertheless, one must wonder what benefits or inducements Wilma can offer her subjects to make them willing to collect approximately 90 daily urine samples. Women who have been unsuccessfully trying to conceive for more than more than a year may be more willing to collaborate if there were the promise of data that might help them and their doctors determine the source of the problem. How (pure numbers or with interpretation) and to whom (the woman or her physician) the data are given could be important in minimizing possible distress if an otherwise unrecognized pregnancy is indicated.

Throughout the process of reviewing and fine-tuning the experimental design, Wilma, like all researchers, must take care to evaluate the potential harms and benefits of the research and work to minimize possible risks in a manner that is consistent with good science, even if the risks appear minimal.

The Informed Consent Process

The design of the informed consent process and form for this study will depend on how Wilma answers the questions posed in the previous section concerning the identity of her potential subjects. Both Ready and Supply are overstating their positions, but there is some truth in each of their statements, as well as quite a bit of room for a creative middle ground. While the regulations do not require that the researcher provide test results to research subjects,(2) one can argue that there is little other benefit that Wilma can offer her subjects for all their inconvenience. Unlike Supply's assertion, daily hCG levels could be useful to women having difficulty conceiving even if the data are received six to twelve months after the samples are collected, but these data could also cause distress to the woman if they reveal an early spontaneous abortion. Determining the study's potential harms and benefits and the steps that should be taken to minimize the potential harms will depend upon the details of the experimental design and the pool of potential subjects. Once these are determined, however, the informed consent process and form must be designed to clearly inform the potential subjects of "any reasonably foreseeable risks or discomforts . . . any benefits to the subject or to others which may reasonably be expected from the research," as well as stating who will have access to confidential data generated by the study and in what circumstances.(2) The consent process should also be consistent with the ethical principle of respect for persons.(1)  Perhaps the best option for Wilma is to give her potential subjects a choice about whether they want to receive their hCG data.

The Dissertation Committee

The interactions of Wilma with Knowledge, her adviser, and Ready and Supply, members of her dissertation committee, are typical and can be used to initiate discussion of several aspects of graduate students' relationships with faculty members.

First, we see Wilma consulting with her adviser on the proposal that will be submitted to the IRB. Knowledge carefully reviews the proposal and recommends some additions and modifications based on his more complete knowledge of the ethical and regulatory concerns associated with research involving human subjects. That is as it should be. It is Knowledge's responsibility as Wilma's research adviser to help her learn how to design an experiment and acquaint her with the norms of the profession.(5) In addition, at many universities faculty advisers must cosign IRB applications, indicating that they will supervise the research and are aware that they are responsible for seeing that it conforms to ethical and regulatory standards (see, for example,(6) ).

Second is the committee meeting where Ready and Supply provide Wilma with very different advice. That is not necessarily a bad thing. One usually tries to get faculty with varied expertise and a variety of perspectives on a dissertation committee(5) . However, if the conversation does not continue toward a compromise, it could be a problem. First Wilma should present her reasoning and offer some other possible courses of action. Then Knowledge should speak up. As Wilma's research adviser, he needs to help Wilma work with her committee to devise an ethically acceptable course of action upon which all can agree. If that is not possible, it may be time to rethink the project or change the committee membership.

  • (1) a b The National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research. The Belmont Report: Ethical Principles and Guidelines for the Protection of Human Subjects of Research, 1979. http://ohrp.osophs.dhhs.gov/humansubjects/guidance/belmont.htm.
  • (2) a b c 45 CFR 46, http://ohrp.osophs.dhhs.gov/humansubjects/guidance/45cfr46.htm.
  • (3)Office for Human Research Protections, The Institutional Review Board Guidebook. http://ohrp.osopht.dhhs.gov/irb/irb_guidebook.htm, Chap. IV.
  • (4)Office of Human Subjects Research, National Institutes of Health, IRB Protocol Review Standards. http://ohsr.od.nih.gov/info/checklist_IRB_protocol.html.
  • (5) a b Committee on Science, Engineering and Public Policy of the National Academy of Sciences, the National Academy of Engineering, and the Institute of Medicine. Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering. Washington, D. C.: National Academy Press, 1997, p. 7 and Chapter 2. http://www.nap.edu/readingroom/books/mentor.
  • (6)Office for Research and the University Graduate School, Indiana University, Researcher Responsibility. http://www.indiana.edu/∼resrisk/resresp.html, and IRB Application Packet Forms http://www.indiana.edu/∼resrisk/forms.pdf.
Commentary On

Identification and Minimization of Risk
Payment of Cash Incentives
Funding of Graduate Students

This case raises three primary issues: the funding of graduate students, the payment of cash incentives to those who volunteer to be the human subjects of research, and the identification and minimization of risk to human subjects. In a discussion of this case, one might choose to cover all or only one or two of these issues. I will consider each of these issues in the reverse order.

Identification and Minimization of Risk

Researchers are responsible for identifying and minimizing risk because of beneficence, the second ethical principle identified in the Belmont Report(1) . "In this document, beneficence is understood . . . as an obligation. Two general rules have been formulated as complementary expressions of beneficent actions in this sense: (1) do not harm and (2) maximize possible benefits and minimize possible harms." Putting this principle into practice, the Common Rule(2)  that regulates government-funded research involving human subjects states that one of the criteria for IRB approval of the research is that "[r]isks to subjects are minimized: (i) by using procedures which are consistent with sound research design . . . and (ii) whenever appropriate, by using procedures already being performed on the subjects for diagnostic or treatment purposes." It also states that a potential subjects should receive "a description of any reasonably foreseeable risks or discomforts to the subject" as well as "for research involving more than minimal risk, an explanation as to whether any compensation and an explanation as to whether any medical treatments are available if injury occurs . . . ." The FDA rules(3) that govern clinical trials of drugs and medical devices have the same requirements.

In this case, then, both the integrity of the research and concern about the subjects require that the research team and the IRB think creatively about possible risks. Frequently, only the potential risks that could come directly from the research protocol are considered. However, other factors in the lives of the subjects could combine with the research to produce more indirect adverse effects. Thus, even a question about the potential subjects' use of prescription drugs might not be sufficient to protect the subjects. In a case like this one, both the taking of nonprescription remedies and the illegal use of controlled substances as well as participation in other clinical trials might need to be determined before the potential risks could be truly evaluated and a person admitted to the study. The regulations make clear that the researcher is responsible for securing this information. It is not the obligation of the potential subject to guess and then volunteer any information that might be relevant.

Payment of Cash Incentives

It has long been common practice to pay people for their participation in research. Some say payment is to compensate them for their time and trouble. Others assert that payment increases subject compliance with the research protocol. Many researchers see compensation as a way to increase their recruitment of potential subjects. As Dickert and Grady note(4) , "this practice is one of the most controversial methods of recruitment. Despite discussions over many years, ethical issues about payment remain unresolved."

Concerns about compensation arise out of two of the ethical principles of the Belmont Report(1) , the principles of respect for persons, and justice. Respect for persons leads to the need for informed consent. But as the Belmont Report asserts, "[a]n agreement to participate in research constitutes a valid consent only if voluntarily given. This element of informed consent requires conditions free of coercion and undue influence." Is the payment for participation in a research study an "undue influence," that is "an excessive, unwarranted, inappropriate or improper reward"?(1) How large does the payment have to be before it is "excessive"? That is one concern.

The other concern is that even at levels that would not be deemed "excessive" for the general population, do financial incentives induce more people of limited economic means to volunteer to be research subjects than the population at large? If so, this disparity is a problem related to the ethical principle of justice. This principle asserts that it is wrong if one group of people bears most of the burden and/or risk for scientific research while another group receives most of the benefit.

Because of concern for respect for persons and justice, the FDA regulations(3)  and the Common Rule(2)  state that "[w]hen some or all of the subjects, such as children, prisoners, pregnant women, handicapped, or mentally disabled persons, or economically or educationally disadvantaged persons, are likely to be vulnerable to coercion or undue influence additional safeguards [should be] included in the study to protect the rights and welfare of these subjects."

From the information given in this case, it appears that Gary volunteers for the clinical trials only because he is in need of money, not because he is interested in supporting the research. The case also suggests that the $3000 offered by the second trial was so large that Gary was willing to risk joining the second trial at the same time that he was participating in the first. Would he have done that if the amount offered had been smaller? We don't know, and this is the sort of question that researchers and IRB members face frequently. What is the purpose of paying research subjects? How much is too much? And is any payment of subjects ethically justifiable? These are important questions to discuss, and the Dickert and Grady article(4)  as well as subsequent letters and articles in the literature are good resources.

Funding of Graduate Students

If graduate students are discussing this case, it will be difficult not to spend at least some time exploring this aspect of the case. The funding of graduate students' education is essential to their academic success -- in fact to their very survival -- but it has not been as frequently or thoroughly discussed as it should be. Discussions between professors and students have often been limited to purely academic topics.

This situation is changing, slowly, but now funding is a legitimate topic of discussion. Here are some examples. The Committee on Science, Engineering and Public Policy of the National Academy of Sciences, the National Academy of Engineering and the Institute of Medicine lists funding as one of the three logistical issues that faculty should discuss with predoctoral and postdoctoral candidates(5) . Particularly with regard to the sciences, Macrina observes that "[o]ne of the unique aspects of predoctoral mentoring is the degree to which the trainee is dependent upon the mentor. In many cases, this dependence is grounded in finances. . . ."(6) . For all disciplines at the University of Illinois, Urbana-Champaign, the Graduate College recommends that departments clearly communicate to graduate students the conditions for their financial support, if any, as part of its Best Management Practices for Graduate Program Improvement(7) . While few would assert that the university has an obligation to financially support all graduate students throughout their studies, most would agree that there is an obligation to clearly describe the support the university is willing to provide and candidly discuss all possible funding options with students. It is possible that Gary was not aware of other options open to him before he volunteered for the two clinical trials.

  • (1) a b c The National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research, The Belmont Report: Ethical Principles and Guidelines for the Protection of Human Subjects of Research, 1979. http://ohrp.osophs.dhhs.gov/humansubjects/guidance/belmont.htm.
  • (2) a b 45 CFR 46, http://ohrp.osophs.dhhs.gov/humansubjects/guidance/45cfr46.htm.
  • (3) a b 21 CFR 50 and 21 CFR 56, http://www.fda.gov/oc/gcp/preambles/53fr45678C.html.
  • (4) a b Dickert, Neal, and Christine Grady. "What's the Price of a Research Subject? Approaches to Payment for Research Participation," New England Journal of Medicine 341(3, 1999): 198-202.
  • (5)Committee on Science, Engineering and Public Policy of the National Academy of Sciences, the National Academy of Engineering, and the Institute of Medicine. Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering. Washington, D.C.: National Academy Press, 1997, pp. 30-32. http://www.nap.edu/readingroom/books/mentor.
  • (6)Macrina, Francis L. Scientific Integrity: An Introductory Text with Cases. Washington, D.C., ASM Press, 1995, p. 17.
  • (7)Graduate College, University of Illinois, Urbana-Champaign, "Best Management Practices for Graduate Program Improvement," http://www.grad.uiuc.edu/Pubs/bmp/index.html.

There are two major issues in this case: 1) the role of the faculty members on a student's dissertation committee in advising, directing and approving the student's research, and 2) the potential conflict between Sherry's roles as researcher and as academic counselor. These issues are intertwined in the case, and their separate consideration should help participants in your discussion delineate the ethical issues and evaluate possible courses of action.

The role of the faculty members who serve on a student's dissertation committee is frequently not well defined. Ideally, they should guide the student's work so that the student develops expertise in the field of study and becomes an independent researcher. As the student progresses in the program, the faculty members' role will change along a continuum from teachers supervising a student to colleagues working in partnership with a peer. Sometimes it is difficult for all to agree on where the committee-student relationship is at a given time, and as a result there can be differences in expectations and miscommunications.

Dissertation committee members act as gatekeepers to the degree, but they must sometimes serve as advocates for students, ensuring that they are treated fairly and that their educational needs are met. The faculty members who train students in research have a responsibility to make sure that the students learn the body of knowledge needed by professionals in that field of study. They also have a responsibility to the students, the scientific community, and any research subjects to ensure that the students' research is ethically sound and scientifically useful.(1)  The reason for using a committee is to avoid exploitation of the student, or advice from just one faculty member that might not represent the best experimental approach or that might not be in the student's best interests. Members of dissertation committees are intended to serve as checks on each other.

Normally, one of the members of the dissertation committee is the student' research advisor and as such has the primary responsibility for directing the student's research. In a situation like this case, in which research with human subjects is involved, the faculty research adviser is also responsible for seeing that all regulations for the protection of human subjects are followed. In the event of a violation of these regulations, both the student researcher and the faculty research adviser can be sanctioned.

So, for a number of reasons, both educational and legal, faculty members need to have a significant role in determining the design of research involving human subjects. It is the degree and style of the guidance from the members of the dissertation committee that are problematic in this case. Is the committee being unduly dictatorial in its insistence on a no-intervention control group? Have they seriously considered other research designs? Have they considered the ethical implications of the proposed experimental design? Has Sherry seriously researched and considered other experimental designs? Has she made a well-reasoned case for her design to the committee? Did she talk over the experimental design with her research adviser before presenting it to the entire committee? Are Sherry and the committee really listening to each other? We have very little information about the group dynamics involved in this scenario, but they are important considerations for Sherry as she tries to decide what to do next. Here is an opportunity to play with the details of the case. Propose different scenarios for Sherry's interaction with the committee (for example: The committee just rubber stamps whatever Professor A says is the way things should be done, or Sherry has not looked into and will not consider any design other than her proposed pre- and post-intervention GPA evaluation.), and then brainstorm ways that Sherry could try to break the impasse. For instance, assuming that Sherry has already done her homework on various experimental designs, she might choose one member of the committee to talk with one-on-one in order to work out a compromise with one faculty member as a starting point to convincing the entire committee. The various proposed courses of action could then be evaluated, and your discussion group could then consider plans on how to put the proposal into action.

In evaluating different experimental designs, it will be essential to consider whether they are consistent with ethical standards, and thus one must consider the potential conflict of interest each might introduce into Sherry's relationships with the students she counsels.

In the Belmont Report, the National Commission asserted that "It is important to distinguish between biomedical and behavioral research, on the one hand, and the practice of accepted therapy on the other, in order to know what activities ought to undergo review for the protection of human subjects of research. . . . Research and practice may be carried on together when research is designed to evaluate the safety and efficacy of a therapy."(2)  So it is not simply the coincidence of intervention and research by Sherry that is problematic. It is possible for Sherry to act in both the role of researcher and counselor to the students with whom she works, but the students must be aware of the dual nature of her role and must freely consent to being subjects in her experiment. Regardless of how the study is designed, Sherry will need to submit her protocol to her institutional review board and include her plans to obtain the students' informed and uncoerced consent, and protect their confidentiality. She will also need to allow the students to withdraw from the study at any time without negative consequences.(3)  Meeting these requirements will take thoughtful planning, especially ensuring that the students do not feel that they ought to agree to be part of Sherry's study if they want to take her study skills course.

However, the problematic part of the experimental design for Sherry and her dissertation committee has to do with the way in which she will evaluate the efficacy of her intervention with the study skills course. Sherry appears to want to minimize the effect of her research on the students with whom she works by simply looking at GPAs for students before and after the course. She is emphasizing the ethical concerns and is reluctant to offer any of her students anything less than what she believes to be the best intervention. The dissertation committee, on the other hand, seems to be emphasizing the scientific concern for producing the most clear-cut result possible from the experiment. Neither concern is irrelevant, and concern that the experiment yield meaningful results is also an ethical concern.(4)

What the committee proposes is the educational equivalent of a randomized clinical trial testing the effectiveness of an experimental treatment relative to a nontreatment or placebo control. The IRB Guidebook states that "the justifying pre-condition for ethical use of randomization is that a hypothesis (i.e., the stated experimental hypothesis that the experimental and control conditions have equally beneficial effect) can be reasonably entertained" and that "the use of placebos is generally unacceptable if there is an effective therapy that the subjects could be receiving for relief of severe symptoms or amelioration of a serious condition." In most cases today, the experimental treatment is evaluated relative to the conventional treatment rather than to no treatment at all. Also, one might argue that being in danger of being dismissed from college is a "serious condition,"(5) and a study skills course is an "effective therapy" that some of the students should not be randomly denied just so that a scientific study can be done. However, if it were clear that the proposed study skills course would be effective, then there would be no research question to be investigated. It may be necessary to review the literature again to determine the amount of uncertainty being tested in Sherry's proposal. Even if the hypothesis condition for randomization cannot be met, there may be other ways to identify a control population, although there may be associated problems for interpretation of the data (for example: at-risk students who self-select not to take the course, or at-risk students at a similar school that does not offer such a course).

In addition to the question of whether a nonintervention group is ethically justifiable, there is the potential conflict of roles and interest for Sherry. Is it acceptable to her employer that she experiment with the counseling and intervention that she does with the at-risk students she is to serve? Will she be able to be clear with the students and with herself as to when she is acting as counselor and when she is acting as researcher? Could the fact that she is conducting research connected with her work dissuade some students from seeking her help? The close linkage of her job and her research could also pose a conflict of interest. She may act or advise some students in such a way as to further her research goals over the best interests of the students. Further, bias could be introduced into her analyses of the research data because of interactions with some of the students. It may be too difficult to ensure that she is fulfilling her obligations as academic counselor and researcher, if she is doing research on the same students that she counsels.

Careful consideration of all these issues and design of the research protocol with the guidance of her committee members to minimize the difficulties will be essential, or her research may simply need to be separate from her employment.

  • (1)American Psychological Association, Section 6.06 Planning Research in Ethical Principles of Psychologists and Code of Conduct, Effective December 1, 1992; American Educational Research Association, Section VI, Guiding Standards: Students and Student Researchers in Ethical Standards of AERA, http://www.aera.net/about/policy/ethics.htm.
  • (2)National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research. The Belmont Report, Ethical Principles and Guidelines for the Protection of Human Subjects of Research. April 18, 1979, http://grants.nih.gov/grants/oprr/humansubjects/guidance/belmont.htm.
  • (3)Code of Federal Regulations, Title 45, Department of Health and Human Services, Part 46, Protection of Human Subjects,1991 and 1994, http://www.med.umich.edu/irbmed/FederalDocuments/hhs/HHS45CFR46.html#46.111.
  • (4)APA, Ethical Principles.
  • (5)Office for Protection from Research Risk, National Institutes of Health, Institutional Review Board Guidebook, 1993 http://grants.nih.gov/grants/oprr/irb.
Commentary On

This case focuses on the multiple relationships that are an essential part of the training of graduate students in the sciences, particularly in the laboratory sciences. The primary emphasis is on exploring the responsibilities of the faculty members who supervise and advise a student's dissertation research, as well as the responsibilities of the student himself. However, one can also use this case to initiate a discussion of the rights and responsibilities involved in the roles of rotating first-year graduate students, senior graduate students supervising less experienced students, and faculty members hoping to recruit new students to the laboratory.

Many universities are now beginning to articulate their expectations of faculty advisers,(1)  and discussion of this case represents an excellent opportunity to investigate the standards at your own institution. More generally, the National Academy Press has published two booklets that address the graduate student-faculty adviser relationship from both the student's and faculty member's points of view.(2)

Weil and Arzbaecher have stressed the need for regular communication in laboratories,(3) and it has been my observation that candid, thoughtful communication is critical for successful relationships between students and advisers. All observations and concerns need to be shared, but at the same time there must be restraint to avoid jumping to unwarranted conclusions. To practice such candor, both must feel that each is looking out for the other's best interests as well as his own. In this scenario, it does not seem that Roger trusts Dr. Hare to look out for Roger's best interests.

Discussion Questions:

  1. While the first few paragraphs of this case scenario do not present any major ethical problems, we discover by the end of the case that there have been some oversights that lead to trouble later. The situation described is not unusual and usually presents no problem. After all, a Ph.D. dissertation is supposed to demonstrate one's ability to carry out independent, original research, and the faculty members on the dissertation committee often will be less familiar with the details of the work than the student. In this case, one wonders if Roger was too independent too soon; if Hare did as much as he should have to learn about the techniques Roger was using (especially if he plans to have his research group continue using them); if the committee members were selected to give optimal guidance or to be a rubber stamp; and if an outside committee member or consultant who was more familiar with the techniques would have spotted potential problems earlier. Perhaps someone who knew more of the technical details would have been more willing to ask a question that would have led to earlier discovery of the problems, but perhaps not. Errors in procedures and interpretation can occur with no one at fault. That is part of doing research and trying to learn something new. What is important is taking reasonable precautions to avoid errors and oversights, and then acting once a problem is discovered. However, the case as a whole gives the impression that Hare is not sufficiently involved in the details of work in his lab and may not even make it a habit to review primary data with his students. Such poor lab management practices make it easier for those who work in the lab to cut corners and even falsify data. 
  2. Roger is being less than candid with Jessica when he fails to tell her that the two data sets are in direct conflict. However, one would think that Roger's dissertation would have been required reading for Jessica as she begins to work on a project based on his research. Had she read the dissertation, she would be aware of the discrepancy already. Roger is choosing not to share his initial concerns. There is a very real question in science concerning when one discusses one's first interpretations of preliminary results. Often it is more prudent to stop, consider the data more thoroughly, and do a few more experiments before going public with one's interpretation.
  3. At this point, Roger steps over the line to unacceptable behavior. He doesn't tell Hare about the conflict he has discovered, with Jessica's help, between his old data and those he has recently generated. He allows Jessica to continue to believe that she made some sort of error in her first experiments. Finally, Roger lies to Hare about the reason that Jessica switched projects, telling him that the first project idea was "likely to be fruitless." It seems that Roger is assuming that he made some error in the past and that if his mistake becomes known the consequences for him will be negative and serious. There are other possibilities. For instance, he is not allowing for the possibility that some unrecognized and uncontrolled variable in the experiments has changed since his last dissertation experiment, and he is precluding the opportunity to identify the variable and possibly learn more about the system he has been studying. In addition, he seems to fear a loss of credibility, respect, or even his degree. That does not seem likely in this scenario. Errors happen, but Roger's actions cause one to wonder what his view of Hare and the department are. He seems to have learned that errors are not permitted. At this point in your discussion, it would be a good idea to brainstorm Roger's options. What could he do? Whom might he contact? How could he present his new findings? Then you can discuss criteria for evaluating these options, and select what you consider to be the best one.
  4. I don't think that Jessica did anything wrong, but I wonder why she did not see the conflict with Roger's dissertation data. It could indicate that she is not fulfilling her responsibilities as a graduate student to read up on the background of the research she is doing and become an independent thinker. She is a first-year student and may not have made the transition to graduate study yet.
  5. Most of this point is addressed in my notes on Question 1, but I would like to have seen at least an ongoing consultation with a scientist who is experienced with the new techniques Roger brought to the lab. The responsibility to ensure that adequate expertise was represented on the committee or was available to Roger would rest first with Hare as Roger's research adviser, then with the committee members as other faculty members responsible for Roger's training, and finally with Roger himself. As a graduate student, he also has a responsibility for his own training and a responsibility to engage in self-evaluation to determine whether he needs consultation with others to guide his research.
  6. This question moves the discussion from consideration of the appropriate level of candor within a research group, to that expected in the wider scientific community. Preliminary, possibly unreliable results are not something that one talks about in the broader community. After all, the indication is that Roger did a whole body of work that supported the conclusions of his dissertation, and one does not discard those conclusions just because a novice researcher produces a disquieting result. Yes, one keeps this new result in mind and promptly follows up on it, but one need not broadcast the conflict at this point. One might argue that if Roger were asked a direct question to which Jessica's data were relevant, he should indicate that there was some recent uncertainty. However, I believe that most scientists would not consider this admission essential.
  7. Once Roger has his confirmatory results, he does need to make the scientific community aware that some uncertainty will need to be tracked down. He need not discard his whole body of work or his conclusions; the problem could simply be due to one of the reagents going bad with time. But it would be best to be honest about the current uncertainty. In that way, he will not run the risk of being seen as deceptive
  8. The standards for certainty are higher for a journal article than for a previously scheduled seminar. Roger probably could not control the timing of the post-doc interview relative to Jessica's experiments and then his opportunity to repeat Jessica's work. However, until it is in press, one can slow publication of a paper until any reasonable uncertainties have been cleared up. Considering that as the case says, Roger's results and conclusions did not entirely agree with the established framework of the phenomenon he was studying, it is in Roger's best interest to be as sure as possible that his data are accurately describe the phenomenon he is studying. However, even with the best practices, one can err because of uncontrolled variables or an unrecognized technical problem. Ethical scientific publication requires that one be thorough and honest in what one presents, not that one be right. Similarly, if the paper based on Roger's dissertation had already been published (probably with Hare as a coauthor) when Jessica did her experiments, then Roger and Hare have an obligation to identify the reason behind the conflicting data before they publish a correction. In fact, depending on what Roger determines is the cause of the conflicting data, a formal correction may not be necessary.(4)  The only exception to this considered approach to the correction of the literature would be if a delay in clearing up the conflict might endanger human lives, for instance, if the data were important in the design of a clinical trial.
  9. If Hare had intended to include Jessica's results in the paper to be published on Roger's dissertation, that suggests that these results are important to the conclusions of the paper and that Jessica would be included as a coauthor. Roger's actions then would also unfairly deprive Jessica of coauthorship, in addition to keeping her from a possibly productive line of research.

Key concepts: graduate student training and advising, student-faculty relationships, laboratory management.

  • (1)See, for example, the "Guiding Standards for Faculty Supervision of Graduate Students" of the Graduate College of the University of Illinois at Urbana-Champaign (http://www.grad.uiuc.edu/grad_handbook/supervision.html).
  • (2)Committee on Science, Engineering, and Public Policy; National Academy of Sciences; National Academy of Engineering; and Institute of Medicine, Careers in Science and Engineering: A Student Planning Guide to Grad School and Beyond (1996), (http://www.nap.edu/books/0309053935/html/index.html); Committee on Science, Engineering, and Public Policy; National Academy of Sciences; National Academy of Engineering; Institute of Medicine, Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering (1997), (http://books.nap.edu/books/0309063639/html/index.html) points of view.
  • (3)Vivian Weil and Robert Arzbaecher, "Relationships in Laboratories and Research Communities" in D. Elliott and J. E. Stern, eds., Research Ethics, A Reader (Hanover, N.H.: University Press of New England, 1997).
  • (4)See Cases 1 - 3 in Chapter 3 of Robin Levin Penslar, ed., Research Ethics: Cases and Materials (Bloomington: Indiana University Press, 1995) for examples of situations that might not warrant formal correction.
Commentary On

This case is very interesting and important because it highlights conflicts that researchers in both medical and social sciences might encounter between their responsibilities to the community as a whole, as personified in this case by the elders, and the individuals members of the community. The elders represent the practices and values that have been a traditional part of the Z community, what has made it "an isolated, closed society, separating itself from the general population," and the researchers in human genetics who work with the community wonder what impact their genetic screening program and research might have on those practices and values.

One of the community's values seems to be to have its elders protect and embody its interests. The case shows us several ways that this value is put into practice. The elders play a critical role in mate choice within the community. There is the suggestion that they do more than just act as a "go between" for the families of the young couple. It seems that they also must approve a potential match before they play this role and courtship is allowed. It is the elders who approach the genetic researchers for help, and it is they who tell the researchers what should be done and instruct the community about who should and should not be screened for the defective BCK allele.

Neither the elders' immediate reasons for asking the researchers to develop the carrier screening and diagnostic testing, nor their long-term goals are completely clear from the case. It appears that the genetic defect that causes the metabolic disorder of BCK has already been identified, and what the elders seek is a screening technique that can identify carriers of the defective allele as well as a diagnostic test to identify BCK-affected newborns so that the restricted diet can be initiated at birth. Does the birth of several BCK-affected infants in the community cause the elders to doubt the efficacy of the community's mate selection traditions? Are they looking for a way to confirm and/or improve the community's genealogies with regard to BCK? Do the elders want to give those already married a way to determine whether they have the potential to give birth to a BCK-affected child? Do they hope to trigger more interest in studying BCK by the medical community? Do the elders just want to ensure that affected infants are identified and put on the restricted diet as soon as possible by alerting the parents to their carrier status and informing them of the diagnostic test available to test newborns? We can't be sure of the elders' motives, but in some communities knowledge of carrier status for genetic diseases is held not by the individuals, but by the elders of the community who then tell the couples whether they are at risk of having an affected child.(1)

In this case, we are told that in the initial screening "testing is done for all who request it and the results are provided to those who inquire." However, we also learn that "the elders do not recommend testing prior to marriage." This scenario brings to mind several ethical concerns. Were the people aware that they were involved in medical research? (Presumably, this screening was a test of the newly developed screening method and so research rather than a standard procedure, but this reasoning could be questioned since the elders sought out the researchers, rather than the reverse, and we do not know if the researchers plan to publish their work.(2) ) Were those tested informed of the nature of the research, and did they give consent? Did some members of the community feel pressured to be tested or alternatively to avoid testing? In other words, was there coercion from the community and/or the elders, if not from the researchers? Based on the principle of respect for persons or autonomy, informed consent should be freely given, but here the researchers might have difficulty identifying or controlling any pressure put on members of the community, and one might question whether they should. The principles put forth by the Belmont Report and used as the basis for our regulations on human subjects research focus on the individual, not on the researcher's responsibilities to a certain, identifiable community as a whole.(3)  Should the researchers interfere with the manner in which such decisions are traditionally made within a community that may very well share medical costs in common?

In a similar way, consideration of how the information generated by the screening will be communicated and stored is complicated by the presence a strong community interest. In general, the IRB Guidebook recommends that subjects receive counseling regarding the genetic information they receive so as to minimize potential psychological harm.(4)  It is hard to predict how people will react to the news that they are carriers for a genetic disease, although unaffected themselves, and there is also the potential for stigmatization by the community if others learn of their carrier status. In addition, people need to understand the probability of false positive and false negative results in such tests. In this case, counseling does not appear to be provided for those tested, and it is unclear who will have access to the information. The case does state that "the results are provided to those who inquire" presumably about their own tests, but we don't know whether the researchers have agreed to make the information available to the elders as well, or if members of the community are expected to make their results known to the elders. The IRB Guidebook says that "in general, except where directly authorized by individual subjects, data may not be released to anyone other than the subject."(5)  But such an authorization could have been part of the form signed before testing, or the researchers could have concluded that this screening did not constitute "research" but was more like work done under contract to the community elders and so the elders should be given all the information. It seems to me that in order to respect the personhood and autonomy of all those tested and to be reasonably sure that the information is handled in a manner acceptable to all, a lot of discussion with those in the community needs to occur before the screening is done, regardless of whether or not this work is considered "research." In this way, the people will clearly understand what will happen to the information generated by the screening before they even decide whether to participate. "Before consenting to undergo genetic tests, whether new tests that are being developed, or already-established genetic tests, subjects should fully understand what it is they are going to learn about themselves, what they are not going to learn about themselves, and how reliable the information will be."(6)

At the end of the case, we see the researchers puzzling over allele frequencies from the initial screening that are much higher than expected based upon the frequency of children born with BCK in the community. Now the researchers are contemplating what would clearly be research aimed at trying to determine the frequency of the defective BCK allele throughout the Z community, including among those not married, and whether there is any evidence of nonrandom mating within the community. The lower frequency of BCK-affected births could be explained in a number ways including: the elders or community custom discouraging mating between two carrier families, a lower allele frequency in the entire community than among the subpopulation tested in the initial screening, or death in utero of a significant percentage of BCK-affected fetuses.

In order to do the research to answer these questions, the researchers would need the support and cooperation of the community elders, and probably would need to make some agreements with the elders on how the study will be done including who will be tested and who will have access to the information generated. The relationship with the elders established with the agreement would generate both practical and ethical obligations. First, there is the purely practical consideration that continued work in community Z requires maintaining a good relationship with the elders. Then as the Code of Ethics of the American Anthropological Association states "fieldworkers may develop close relationships with persons. . . with whom they work, generating an additional level of ethical considerations" concerning their obligations(7) . In many ways, this work is a type of anthropology, and it is instructive to note that the anthropology code of ethics states that "anthropological researchers must expect to encounter ethical dilemmas at every stage of their work," that "researchers have primary ethical obligations to the people. . . they study. . . [and] these obligations can supersede the goal of seeking new knowledge," and that "researchers who have developed close and enduring relationships (i.e., covenantal relationships) with either individual persons. . . or with hosts must adhere to the obligations of openness and informed consent, while carefully and respectfully negotiating the limits of the relationship." Could the research have an effect on the culture and practices of the community such as mate choice, and does this possibility preclude the work? Are the "people" to whom the researchers have an obligation best represented by the individuals being tested or by the community as a whole represented by the elders? What should the researchers do if the obligations to individuals and to the "host" elders conflict? These are concerns of a type that is just starting to be considered by researchers and ethicists and on which there is no consensus at this point.

It might be interesting in your discussion to consider the following extension of this case, which makes the potential conflict between obligations to the community and to individuals within the community more immediate: Let us assume that the researchers have worked out an agreement with the elders of the community for genetic screening of all those in the community who are willing to be tested. The agreement is that the results of the tests will be made available to those married individuals who request them, but that no results will be given to the elders or to community members who are not married. The individual results will be kept confidential by the researchers, although the aggregate results (such as allele frequencies) can be used in publications and will be made available to the community. Now suppose that two young, unmarried people from the community contact the researchers saying that they want to be tested for the defective BCK allele but only if they will be given their results. The young people say that the elders have not approved their courtship and the couple suspects it is because BCK is known to run in both of their families. They hope that if they can show the elders that one or both of them is not a carrier, the elders will change their minds and approve their courtship and marriage. What are the obligations of the researchers, some of whom are medical doctors, to these two young people? What are the researchers' obligations to the community? Note that this unmarried couple is asking the researchers to help them in their quest to overturn what is seen in the community as a decision that cannot be appealed.

  • (1)D. Abeliovich, A. Quint, N. Weinberg, G. Verchezon, I. Lerer, J. Ekstein, and E. Rubinstein, “Cystic Fibrosis Heterozygote Screening in the Orthodox Community of Ashkenazi Jews: The Dor Yesharim Approach and Heterozygote Frequency,” European Journal of Human Genetics 4 (6, 1996): 338-3341.
  • (2)National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research, The Belmont Report, Ethical Principles and Guidelines for the Protection of Human Subjects of Research, 1979, http://grants.nih.gov/grants/oprr/humansubjects/guidance/ belmont.htm.
  • (3)National Commission, Belmont Report; Code of Federal Regulations, Title 45, Department of Health and Human Services, Part 46, Protection of Human Subjects, 1991 and 1994, http://www.med.umich.edu/irbmed/FederalDocuments/hhs/HHS45CFR46.html#46.111; Office for Protection from Research Risk, National Institutes of Health, Institutional Review Board Guidebook, 1993, http://grants.nih.gov/grants/oprr/irb.
  • (4)OPRR, Guidebook.
  • (5)Ibid.
  • (6)Ibid.
  • (7)American Anthropological Association, Code of Ethics, 1998, http://www.aaanet.org/committees/ethcode.htm.
Commentary On

As one can quickly see from scanning this case, the central issues are trust and honesty. However, the case also raises several more peripheral yet important issues, including proper data management and the responsibilities of authors, collaborating researchers and faculty research advisers.

Trust is essential for science, particularly in collaborative research settings. In fact, in 1995 an entire issue of Science and Engineering Ethics was devoted to consideration of "Trustworthy Research." (Vol.1, No. 4, 1995) Trust, in turn, depends on honesty, the value listed first among the core values of science. (National Academy of Sciences et al., 1995, p. 21) If someone is seen as dishonest, we do not trust him and avoid him as a collaborator. In Part 1 of this case, we are asked to consider what can happen within a research group when trust is weakened; in Part 2 we look outward to consider the possible effects within the larger scientific community.

There is no glaring instance of misconduct in this case. At most, Peter can point to a lack of primary data in Sally's notebook for the second set of mice and report that she didn't save any of the cells for reanalysis. The failure to preserve relevant data has been termed a questionable research practice, something that may be detrimental to the research process, but is not misconduct. (National Academy of Sciences et al., 1992, Vol. 1, p. 28) However, when he considers the sum of Sally's actions, Peter comes to wonder if she might have falsified or even fabricated her results from the second group of mice so that they would be consistent with those of the first. If so, her action would be misconduct. Peter must decide whether and how to act on his misgivings. Exploring his options and identifying the factors he must weigh in making his decisions are the most obvious foci for discussion of this case.

The case also could be used to trigger a discussion of best practices for the conduct of collaborative laboratory research. The discussion group could identify what types of standard operating procedures for the laboratory could have been put into place before this incident so that the problem Peter faces would never have arisen. What if the lab agreed on norms for the number of subjects tested and/or replications to be done before publication of the results? What if there were a clearly articulated expectation that primary data will always be kept and preserved in a bound notebook? What if it were a standard part of preparing a manuscript to have all the authors sit down together to review the primary data, not just the graphs and tables generated from the data? That is exactly what some laboratories have done, but often only after a crisis. After two independent but concurrent incidents of misconduct in his laboratory, Lee Hood is reported to have "formalized the review process, so that each paper is now reviewed by three people inside the lab." In addition, "[t]here is considerably more emphasis on dealing with raw data, not merely a synopsis of the findings. And Hood now also requires everyone to keep a bound lab notebook." (Roberts, 1991, 1347)

Discussion Questions

Question 1

This question challenges us to consider what options are open to Peter, and how he might decide among them. Some possible options are 1) going ahead with the writing as if he had no misgivings,2) confronting Sally, or 3) telling Larson that he will not prepare the manuscript until the data can be verified. Surely discussion will generate other possibilities, such as Peter loudly and without warning declaring at a laboratory group meeting that he's certain Sally has fudged her data, that he's never going to work with her again, and that others had better watch out for her. Such a course of action is clearly inappropriate because it fails to show respect for the people involved, and it could have some very serious negative consequences for a number of people, including Peter. However, brainstorming that includes such inappropriate options can help us exercise our imaginations and improve our creativity in devising innovative appropriate solutions.

Next, one must evaluate the many possible courses of action and select among them. This process involves looking at the possibilities from a number of perspectives, considering ethical principles and obligations, predicting possible consequences and making reasoned judgments that take account of conflicting interests, principles and obligations. During this part of the discussion, it is important for participants to justify their positions, rather than falling back on "I think" or "I feel" statements. Discussants must be able to explain their reasoning so that the group can work toward a consensus on which possible courses of action are ethically acceptable, which are not, and why. Then the relative merits of the acceptable options can be evaluated to determine the best course of action. In the end, participants may differ on the best option because they will give different weights to conflicting obligations or principles.

In this scenario, Peter would need to consider the interests of such people as Sally, Larson, the rest of the lab members, and other scientists working in this field, in addition to his own. He needs to examine his obligations to these and other people, as well as the ethical principles on which these obligations are based. For instance, he has a basic obligation as a member of the human community to treat other people respectfully, as he would like to be treated. But he also has a responsibility to the scientific community in particular to honestly report the results of his research because of his respect for them, their time and their ability to draw their own conclusions from the data. These and other obligations may appear to be in conflict. Then Peter needs to examine the possible courses of action open to him to determine the possible consequences of each alternative, and evaluate how well each option fulfills his obligations and takes into account others' reasonable interests.

Discussion Questions

Questions 2 and 3

These questions focus the discussion on the responsibilities of the authors of scientific papers. For instance, how sure of your results do you need to be before you publish? In discussions of research ethics, we often focus on appropriate authorship and investigate criteria for determining who has the right to be listed as an author. The responsibilities of authorship, the flip side of this coin, often are not considered as thoroughly, probably because of the large variation in expectations.

Generally, an author is expected to take responsibility for the validity of the data presented in a paper, but there is some question as to whether one is responsible for all of the data presented, or only for the data one actually collected. In "Responsible Science," the NAS committee seems to argue for collective responsibility, saying that "the privilege of authorship should be based on a significant contribution . . . as well as a willingness to take responsibility for the defense of the study should the need arise." (National Academy of Sciences et al., 1992, Vol. 1, p.140) The American Society for Microbiology also comes down on the side of collective responsibility, stating that when publishing in its journals, "All authors of a manuscript must have agreed to its submission and are responsible for its content . . . ASM considers all authors responsible for the entire paper." (Journal of Bacteriology, 1998, Vol. , p. I-ii) In contrast, the Journal of the American Medical Association indicates that "[a]uthors may include explanation of each author's contribution and add a publishable footnote explaining specific contributions," presumably to indicate who is taking responsibility for what parts of the study (JAMA, Jan. 7, 1998, Vol. 279, p. 67). As the International Committee of Medical Journal Editors has asserted, "[a]ny part of an article critical to its main conclusions must be the responsibility of at least one author" (JAMA, 1997, p. 928), but, one infers, not necessarily all. Nature, where the characters in this scenario plan to publish, has not published its expectations concerning the criteria for and responsibilities of authors in its "Notes to Contributors" (Nature, 1997, p. 702), and we do not know the conventions in the characters' field, professional organization(s) or laboratory. Similarly, the conventions for the meaning of and responsibilities of first vs. last vs. internal authors vary considerably from discipline to discipline, and even from laboratory to laboratory.

Thus, even though he is the first author, Peter may be able to indicate via a footnote in the manuscript that he accepts responsibility only for the data he collected and not for the data from the cell function assays that Sally ran. Is that course of action ethically tenable? One's answer will vary depending on one's view of the responsibilities of authorship.

This case may trigger a discussion exploring the range of variation in the criteria for and responsibilities of authors that would inform students of the very real variation and point out the benefits of discovering the local conventions before they become authors.

Discussion Questions

Question 4

Now we pause to look at the actions of Larson, the faculty member who directs the laboratory in which Peter and Sally work, and who is the principal investigator on the grants that support the lab. In these roles, he is the person ultimately held responsible for the validity of the work done within the laboratory, and therefore is responsible for quality control. However, he is more than just the lab director: He is also the faculty adviser for Peter and Sally, two fourth-year graduate students working toward the Ph.D. As such, he is responsible for training them and helping them to develop the skills they will need to become independent investigators. Two of the ways in which he should do this are to model best practices in laboratory management, and to be explicit about how and why things are done.

One can argue that Larson acted properly when he told Sally that she would appear as the second author on Peter's manuscript if her data were informative, because he was providing explicit information as to what Sally might expect from the work she is being asked to do for Peter's project. He is also being consistent with the expectation that one is an author only if one has significantly contributed to the information presented in a paper. However, one could also argue that he is putting Sally in a difficult position, tempting her to fudge results so that her data will be seen as "informative."

That is particularly true when one notices indications that best practices are not the rule in this lab, so that tainted data may not be caught. This lab seems to have no standard procedures for recording and preserving primary data, or for storing samples for possible confirmatory analyses. Similarly, there are no routine mechanisms by which other members of the lab discuss primary data or review manuscripts. For instance, note that Larson never asks to see the raw data from Sally's analyses; he only looks at her graphical interpretations of the data. Peter didn't feel he could ask directly about the data, but instead went to look at Sally's notebook on the sly. He now seems to feel that publicly asking about the data will depart from routine lab practice sufficiently to stigmatize Sally. If a set of laboratory procedures had been in place and were routinely followed, Peter would not be in the quandary in which he now finds himself, and Larson would be publishing papers of higher quality.

Of course, this discussion does not address the very real problem of survival in the competitive atmosphere of contemporary scientific research. Many people use the time pressure as a justification for sloppy record keeping, lax laboratory oversight and over-interpretation of marginal data. However, consider all the time that the Hood lab spent sorting through the mess left after misconduct was discovered in their lab, and how much time and energy Peter is spending worrying about Sally's data rather than writing the manuscript and doing further experiments. Best practices include not falling prey to short-term expediency.

Discussion Questions

Questions 5 and 6

Now that the paper has been published and presumably discussed at scientific meetings, any actions Peter might take will occur in a far more public context. The basic ethical considerations are the same as they were when he wrote the paper, but now some of the possible consequences are different, and an additional course of action is now possible: retracting the paper.

If this were the only paper on the function of this gene and if understanding the gene had become important for human health between Part 1 and Part 2, then one might argue that Peter's obligation to investigate Sally's analyses had increased. However, other groups have already cast doubt on the Larson lab results so that Peter's failure to act will not endanger human lives or health. This is an example of the self-correcting nature of science to which many have referred. However, it is not without cost.

Note that it is possible that Sally really did get the sample labels correct, and the first two sets of analyses were a fluke. Alternatively, the assay conditions, mice or knockout alleles studied by the Larson lab may differ from those used by other labs in some way that affects the cell function analyses. Discrepancies between labs do not always indicate fraud, and Peter needs to be careful. Until the situation is clarified, however, the Larson lab will be perceived as either sloppy or dishonest.

References

  • American Society for Microbiology. "Instructions to Authors." Journal of Bacteriology 180 (January 1988): I-ii.
  • International Committee of Medical Journal Editors, "Uniform Requirements for Manuscripts Submitted to Biomedical Journals." Journal of the American Medical Association 277 (March 19, 1997): 927-934.
  • Journal of the American Medical Association. "Instructions for Authors." 279 (January 7, 1998): 64-74.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. On Being a Scientist, Responsible Conduct in Research. Washington, D.C.: National Academy Press, 1995.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Careers in Science and Engineering: A Student Planning Guide to Graduate School and Beyond. Washington, D. C.: National Academy Press, 1996. Available online at http://books.nap.edu/catalog/5129.html.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering. Washington, D. C.: National Academy Press, 1997. Available online at http://www.nap.edu/readingroom/books/mentor.
  • Nature. "Notes for Contributors." 390 (December 18/25, 1997): 702.
  • Roberts, Leslie. "Misconduct: Caltech's Trial by Fire." Science 253 (1991): 1344-1347.

The main issue raised by this case is the relationship between a graduate student and the student's faculty adviser. What should this relationship be like, and what can and should one do if the relationship goes sour? A secondary issue concerns how and when one should report misconduct by a faculty member.

Most will readily accept that misconduct is relevant to research ethics, but some will question whether the student-adviser relationship fits in this category. Because it concerns people's treatment of each other, many scientific societies and writers in the field of research ethics agree that treatment of graduate students is an issue in research ethics. A committee of the National Academy of Sciences included "Inadequately supervising research subordinates or exploiting them" among questionable research practices -- that is, "actions which violate traditional values of the research enterprise and that may be detrimental to the research process." (Responsible Science: Ensuring the Integrity of the Research Process, Vol. 1, p. 28, National Academy Press, 1992) With regard to relationships in research groups going sour, as is the situation in this case, Weil and Arzbaecher assert, "We can collect these ways of going astray under broader ethical questions about how to wield power responsibly and how to behave responsibly as one dependent on the power of others. As we proceed to point out the kinds of standards and practices that are needed, we thereby delineate role responsibilities in research groups. To fail to fulfill these role responsibilities would be to behave irresponsibly, that is, unethically."(Weil and Arzbaecher, p. 78)

In the past, it was often assumed that the student's research adviser would serve as the student's mentor as well. This assumption is still common in the natural sciences, but more and more people are using the term "mentor" as an honorific rather than as a description of an assigned role. For instance, Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering, states:

In a broad sense, a mentor is someone who takes a special interest in helping another develop into a successful professional. . .A fundamental difference between a mentor and an adviser is that mentoring is more than advising; mentoring is a personal as well as a professional relationship. An adviser might or might not be a mentor, depending on the quality of the relationship. (National Academy of Sciences et al., 1997, 15)

While the best situation may be to have one person fulfill both roles, that is not always possible for a number of reasons. Concerning the choice of an adviser, the National Academies' Student Planning Guide says, "The ideal person can not only guide your career, support your research, and help to find you a job, but can also serve as a close and caring mentor - a 'research uncle,' as one author puts it. Obviously, this is a rare combination, but one worth searching for." (National Academy of Science, 1996, 69) Often the personalities of the student and the adviser do not facilitate such a close relationship, and even when personalities are compatible another person may be a better mentor in a specialized area such as teaching or preparation of presentations. In fact, it has been asserted that "[n]o mentor can know everything a given student might need to learn in order to succeed. Everyone benefits from multiple mentors of diverse talents, ages, and personalities." (National Academy of Science et al., 1997, 5) That can be especially true when the student is a woman and the faculty adviser is a man, as is the situation with Hogan and Simpson. It has been observed that "[w]hile academic advisors are supposed to serve as formal mentors for women, they do not always do an adequate job. . . . Women often react by reaching beyond their official advisors to find other mentors among faculty from other disciplines, peers, or classmates," just as Hogan reaches out to Rodriguez in this scenario. (Bird et al. 1993, 8) And the National Academy of Sciences et al. suggest, "You might decide to seek several advisers to broaden the range of counsel available to you . That is particularly important for women and minority-group students, who might wish to have a woman or member of their minority group as a mentor." (National Academy of Sciences, 1996, 75) For these reasons, having a mentor who is not one's research adviser, having more than one mentor, or developing group mentoring opportunities are now being encouraged.

When a student's relationship with a mentor who is not the research adviser goes sour, the termination of the relationship can be difficult, but it will not usually have long-term negative consequences for the student. However, termination of a relationship with a research adviser can lead to a number of negative consequences including slowed progress toward one's degree, a change in the direction of one's research project, damaged reputations, and perhaps the need for a change to a different department or school. The Student Planning Guide offers the following advice:

What can you do if the relationship with your adviser is a poor one? If the two of you cannot work it out, you should try to find another professor who is qualified and willing to take you on. In general, it is best to make a change as soon as you see that the situation is unworkable. . .Only if it is late in your student career should you endure a difficult situation rather than try for a better one. The head of the graduate program or the departmental chair might be able to help you to decide what to do and who might help you." (National Academy of Sciences, 1996, 75)

The Guide also notes, "It is very important to remember that the education of a graduate student is the responsibility of an entire department, not just of a single adviser." (National Academy of Sciences, 1996, 70) However, not all departments acknowledge this responsibility, and the effects of changing advisers will depend on many factors including the department's attitude toward such changes, the details of the specific situation, and how the student and the advisers, old and new, negotiate their way through the change. If it is possible to be civil and rational throughout, the change may be beneficial to all concerned, but rumors, accusations, and recriminations can easily poison the atmosphere.

The secondary issue, the misconduct charge against Simpson, will be addressed in the discussion questions.

Discussion Questions

Question 1

This question explores the reasons why Hogan might not want to publicly accuse Simpson of plagiarism. Naturally she fears direct retribution and damage to her relationships within the department and the discipline. While I agree with the NAS panel that "every case of misconduct in science is serious and requires action," (Author? Responsible Science: Ensuring the Integrity of the Research Process, p. 31), I do not believe that the action must be either direct or immediate in this case. That is because the risk of potential harm to Hogan is so great while the risk of harm to others if she delays is minimal. Although Simpson goes too far when he assets that plagiarism is a harmless little transgression, it is true that it is not on the same level as publishing falsified data from a human clinical trial. Therefore, Hogan has some time to stop and carefully consider her actions.

First, she must be absolutely certain of what she saw, and she should have documentation; photocopies would be best. How much was plagiarized and where? Was it direct copying or a paraphrase without citation? Second, she needs to know her institution's regulations and the various routes by which she might make an accusation. Third, she needs to talk to a trusted faculty member like Rodriguez in confidence to check her reasoning and actions. Fourth, she needs to come up with as many creative possible courses of action as she can and then decide which is best. And fifth, she needs to design and carry out a plan of action. Steps four and five will probably involve consultations with the trusted faculty member.

These are steps that the discussion group can follow. The brainstorming to develop possible courses of action, and the investigation of institutional misconduct regulations and procedures might be the most valuable elements of the discussion. For instance, some may realize that it is not clear that Hogan needs to be directly involved in the accusation of misconduct at all. If Simpson publishes the book with the plagiarized material, then the author who was plagiarized could make the accusation, rather than Hogan. She need only make him/her aware of it, and that could even be done indirectly.

Question 2

Here we are asked to consider the conflict between Rodriguez's obligation to honor Hogan's request for confidentiality, and Rodriguez's obligation to her institution and the scientific community to report Simpson's plagiarism. As a faculty member and a member of the scientific community, Rodriguez has a responsibility to see that probable misconduct is reported to the proper authorities, but that does not have to be done immediately. It does not seem likely that the plagiarism will result in immediate, serious harm to anyone if it continues to go unreported for a while longer, and Rodriguez, like Hogan, needs to take time to learn about the facts of the situation and the local regulations, and to consider her options. She might even want to talk to a faculty friend in Simpson's department to sound out the situation there. Barring the risk of immediate harm to others, it is important that Rodriguez give Hogan time to develop her own plan for reporting the plagiarism, both because of her promise to Hogan to keep it confidential and because knowledge of the breached confidence would deter other students from seeking necessary advice in delicate matters. Thus, Rodriguez needs to respect Hogan's wish to develop her own plan for making the accusation, but Rodriguez does have an obligation to be sure that an accusation is made in a reasonable amount of time if the evidence for plagiarism is sound. A way needs to be found to make Simpson accountable while minimizing the possible harm to Hogan and herself, perhaps by having the accusation come from someone outside their institution.

Question 3

Some might argue that untruthful answers are never morally justified, but in this situation Hogan's untruthful response to the chair's question may be her best course of action, considering the possible consequences. However, that does not mean that Hogan has no responsibilities toward other graduate students, the academic community, or Simpson. Rather, it means that she may be able to fulfill those obligations through actions that pose less risk to herself.

Question 4

Many scientists believe that it is possible to have a successful mentoring relationship with a faculty member outside one's department or discipline. In discussions of mentoring with graduate students, I am learning of an increasing number of such successful pairings, particularly among students who have more than one mentor.

Questions 5 and 6

As written, the case indicates that Simpson's plagiarism leads Hogan to decide that she cannot continue to be advised by a person who knowingly engages in such unprofessional conduct. However, an adviser could engage in other types of unprofessional behavior that might make the continued relationship impossible for the student. These questions ask what a student could and should do in such a situation.

There are many reasons why the relationship between a student and his/her research adviser might go sour, short of unprofessional behavior. However, the basic advice is the same for almost all situations: Try to resolve the situation through improved communication and/or changes in procedures; if not, change advisers as soon as possible. What varies from situation to situation is whether the student should report the unprofessional conduct by the adviser, and to whom the report should be made. If the behavior is likely to be repeated with other graduate students and to have a deleterious effect on them as well, then the student has some obligation to report the behavior and so attempt to protect others. The report might be made to the graduate studies director of the department, the departmental chair, or some other senior faculty member who would have the standing to do something to change the adviser's behavior. Alternatively, a student might go to the university's graduate school administration, an advocacy office, or an ombudsperson, if one exists.

As discussed in the comments on Question 1, it is important to consider the person to be approached, the timing and the form of the complaint when projecting possible consequences and determining the best course of action. There is always the danger that the student, especially if she is a woman, will be viewed as a whiner and/or not tough enough for the academic world. The manner in which the complaint is made needs to be carefully considered to ensure that it is a factual report of observed incidents and not a formless recitation of grievances. In some cases it may be best to switch advisers first and report the unprofessional behavior later.

Question 7

I think that most will agree that Simpson is not qualified to train graduate students to become professionals in the field if he knowingly engages in plagiarism and thinks of it as typical behavior. The more interesting discussion would concern whether his behavior toward Hogan while her adviser would make him unsuitable to advise any graduate student. What are the minimal qualifications for an adviser? How can we help adequate advisers become great advisers?

References

  • S. J. Bird, C. J. Didion, E. S. Niewoehner and M.D. Fillmore. Mentoring Means Future Scientists: A Guide to Developing Mentoring Programs Based on the AWIS Mentoring Project. Washington, D. C.: Association for Women in Science, 1993.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Careers in Science and Engineering: A Student Planning Guide to Graduate School and Beyond. Washington, D. C.: National Academy Press, 1996. Available online at http://books.nap.edu/catalog/5129.html.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Adviser, Teacher, Role Model, Friend: On Being a Mentor to Students in Science and Engineering. Washington, D. C.: National Academy Press, 1997. Available online at http://www.nap.edu/readingroom/books/mentor.
  • National Academy of Sciences, National Academy of Engineering, Institute of Medicine. Responsible Science: Ensuring the Integrity of the Research Process, Vol. 1. Washington, D. C.: National Academy Press, 1992.
  • Weil, V., and R. Arzbaecher. "Relationships in Laboratories and Research Communities" in D. Elliott and J. E. Stern, editors. Research Ethics: A Reader. Hanover, N. H.: University Press of New England, 1997.

Overview

Part 1

Part 2

Overview

This case raises some very interesting and pertinent issues for those working in the sciences -- issues of intellectual property, intellectual turf, the training of graduate students, communication and cooperation vs. competition in science, and the evaluation of our scientific peers. In general, there are neither rules nor explicit guidelines from professional societies or universities that address these issues. At best, there are norms and/or precedents that senior scientists picked up somewhere and that they may pass on to their junior colleagues. Yet questions like "How do I give others appropriate credit for information they have shared with me informally?" are central to the lives and careers of scientists, and deserve more careful consideration. That is what this case seeks to facilitate.

While there are many "right" answers to the questions posed, some of which will be better than others, there are also "wrong" answers. Before diving in and trying to solve Eileen and Steve's problems, it is important to consider the criteria by which we judge the ethical rectitude of people's actions.

First, I would submit, the proposed course of action must demonstrate respect for the people affected by it. By respect, I do not mean just civility, but rather the respect for persons described by Kant:

Act only according to that maxim by which you can at the same time will that it should become a universal law. . . .

Act so that you treat humanity, whether in your own person or in that of another, always as an end and never as a means only. (Kant, 1785)

In other words, the proposed course of action, if it is to show respect for others, must be one that we would be happy to see everyone follow, and it must not treat people as things. The course of action needs to be consistent with the obligations of the people involved toward each other. These people need to have a voice in designing the solution to the problem, and all must be able to choose for themselves what they will do.

Second, the possible consequences of the proposed action must be considered. What are the good things that might result? What are the bad things that could happen? How are the potential benefits and harms distributed among the people involved? An ethically sound course of action should result in more benefit than harm. More than that, it should minimize the possible harms or risks, and ensure that the benefits and harms are equitably distributed. A course of action that has the potential of five benefits and two harms is better than one that could result in ten benefits and seven harms. A proposed course of action that exposes a single person to all the potential harms person, while reserving the benefits to the other two people involved, is not as good as a plan that has all three share equally in the risks and benefits.

Those doing research involving human subjects may recognize the criteria presented here as those underlying all the human subjects regulations: respect for persons, beneficence, and justice (National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research, 1978). These are some of the basic principles of ethics, and they provide excellent criteria by which to judge any action. Besides, if we use these three principles to guide our interactions with our human research subjects, shouldn't we expect a similar standard of conduct for our interactions with our colleagues?

This case presents a wonderful opportunity for the participants in the discussion to share their experiences, their knowledge of the norms in their disciplines and laboratories, and their ideas for how these situations should be handled. The experiences and norms will be quite varied, and will induce the participants to start evaluating the different conventions and to think of new solutions to the problem. At this point, the group can move on to evaluating the different potential courses of action suggested, and determining what should be done and why.

One additional comment: When discussing this case, someone will probably say, "Well, why doesn't Steve just join Eileen's lab? That would solve everything." However, this solution might not be feasible. Eileen may not have room for an additional student, or Steve may not be interested in plant population genetics. Perhaps Steve is interested in studying lizard species and has joined Bill's group because reptiles are the experimental animal of choice in Bill's lab.

Discussion Questions

Part 1

Question 1. Collaboration carries with it a tension. We usually see ourselves as selfless researchers exploring the world around us for the good of science and humanity; in that light, collaborations should be a good thing because they are frequently a more efficient use of resources. Yet, we judge each other based on personal achievement, individual inventiveness and insight. A member of a collaborative team is usually not as highly regarded as a solo researcher who has produced similar results and ideas solo. On the other hand, being a member of a collaboration is usually better for one's career than losing in a head-to-head competition between researchers to see who can publish first and so claim the discovery. Maybe it shouldn't work this way, but it often does.

The way that this question is written suggests that one may have an obligation to accept a proposed collaboration. That is an interesting idea that merits discussion. It is generally, although not universally, accepted among scientists that we have an obligation to make our discoveries known to others, usually through peer-reviewed publication. It is somewhat less generally accepted that once publication has occurred, one has an obligation to make available to other qualified scientists the unique research materials that were used in one's work such as biological strains (e.g. mice, plant seed stocks, tissue culture lines). Recently, the willingness to share unique research materials has been made a prerequisite for publication in a number of prestigious journals in the biological sciences. Collaboration, while usually viewed in a positive light, is not viewed as an obligation, and I'm not sure how one could argue that it is a general ethical duty of scientists toward each other.

However, the relationship between Eileen and Steve is special in some ways. She is a professor, a teacher, at a university, and he is a beginning graduate student in her department. Therefore, Eileen has greater obligation to help in Steve's training than that of a professor at another university or in a different department. In addition, at the brown bag lunch seminar, which Steve was probably required to attend as part of his training, Eileen exposed him to an idea that now seems to be coloring all his thoughts as he works to design a research project. She must take some responsibility for the consequences. Does that mean that Eileen must accept Steve's offer of collaboration? I don't think so, but it does suggest that she needs good reasons for refusing to collaborate.

Question 2. This question presents some possible reasons that Eileen might have for refusing and asks us to evaluate them. If we agree that Eileen does not have an absolute obligation to collaborate with Steve, then the questions are how strong is Eileen's obligation, and are her reasons for refusing sufficient to counter that obligation? Neither issue is easy to evaluate, but I would submit that if the cost of the collaboration to Eileen is high in terms of effort, time and money, it is more difficult to assert that Eileen should collaborate with Steve. Of course, the conflict between Steve's right to follow up on his plan for an experiment and have a good thesis project vs. Eileen's right to receive credit for originating her idea and control its public presentation still remains, but it should be possible to come up with a compromise that respects both of these claims. Determining whether they should participate in a proposed compromise plan is a question of the costs to each and their equitable distribution.

Discussion Questions

Part 2

Question 1. The ethical implications of "sitting on" an idea would be most fruitfully explored by looking at the expected consequences of continued "sitting" relative to letting another pursue the idea. The question points out that there may be potential harms or benefits to others besides the principals in the scenario, and that these should be considered as well. In the overall discussion, however, it is important to keep the principles of justice and individual rights in mind so that one does not just concentrate on ethical calculus.

Question 2. The means by which Eileen communicated her idea does make a difference because it bears on the issue of her right to receive credit for the idea. If it were a large, public forum, many in the field would know the idea was hers, and she would receive appropriate credit within that scientific community. That would be particularly true if there were some written record of her presentation of the idea in a technical note or poster abstract. Publication makes an idea available to all to pursue as they wish. Most researchers these days understand that anything presented in any form at a meeting may be pursued by anyone who learns of it.

In this case, the context for Eileen's presentation was an informal, in-house seminar. That makes the question much more difficult because keeping such for as open and free-wheeling as possible is beneficial to everyone; to achieve that goal, the participants must feel safe to share ideas that are not yet formally claimed as their own.

Question 3. I have never heard of a formal declaration of a "statute of limitations," but arriving at an understanding of this concept would be beneficial to all. In the past, I have usually heard this concept invoked by advisers who feel that their former students are talking too long to write up their research for publication. It is difficult to determine how long is "long enough," and I doubt that a single, interested colleague like Bill can do it fairly. It would require a group familiar with the experimental systems involved. In this case, one year may be far too short if Eileen gets only one growing season per year for her experimental plant and plans her experiments one to two seasons ahead.

Question 4. Eileen's actions differ from those of Dr. Igneous in at least one important way: She is being honest about her development of her idea and work to test it. She is not trying to mislead others, and she has already made an investment in the development of the idea. Dr. Igneous may not have generated any original ideas, but just claimed to be doing experiments in a number of areas to minimize the competition. The amount of work Eileen has already invested in the development and refinement of her ideal does matter because if she does not receive credit for originating this idea, the work already invested will represent a loss to her; a cost or harm to her as possible result of Steve's course of action. The amount of work she has done on the idea is also an indication of her determination to follow through with the testing and not just sit on the idea as Dr. Igneous did.

Question 5. Eileen's case for refusing Steve's offer to collaborate is strengthened if she has a clearly identifiable reason for her delay and if that delay has an identifiable end point. An example might be a heavy teaching load this year but a free semester in the next year.

Regardless, the situation for Steve is still not good, and some sort of creative solution would be best for all involved. Steve appears to be a grad student who has been captured by an idea. It is as if Eileen's model explains the observations he and others have made on his experimental population, and thinking about it has changed his view of all future investigations he has planned. He cannot ignore it. The idea has become part of the way in which he thinks about his research. Steve is not saying that Eileen should not get credit for the idea, but rather that he needs to be allowed to follow where consideration of it is taking him.

Question 6. Bill's role in this case is one that deserves some consideration. He is Eileen's departmental colleague and possibly friend, but he is presenting Eileen with what amounts to an ultimatum and then justifying his actions with the Dr. Igneous story. He is a very important player in this scenario, and it would be useful to explore what his obligations to the other people are, and what alternative courses of action he might have taken. Steve does need an advocate, and Bill is the logical person to fill this role, but Bill's course of action will not result in the best possible consequences for any of those involved.

Question 7. Bill and Eileen's argument certainly does indicate a tension between personal ownership and collaboration, between competition and cooperation, between individual recognition and the good of science. This is a very real tension in science and one we all try to balance. Collaborations are generally perceived as a good thing, but some may question the individual abilities of those who always work in collaboration with others. Collaboration with a more senior researcher can also result in the shadow effect; most assume that the major ideas and impetus for the work originated with the senior partner. Students need to be aware of the reality of these tensions, and of the need to work toward changing the culture of science so as to decrease them.

Brainstorming possible solutions to the situation as it stands at the end of Part II, followed by an evaluation of the various ideas generated, would be a good way to close a discussion of this case. This approach will help the participants think both creatively and critically if they find themselves in a similar situation. If the group wants to go further, they could try some role playing and work out what the characters in this case might actually say to each other as they try to implement the course of action the group has decided is best. Coming up with an equitable solution is one skill, and implementing it is another. Both require practice.

References

  • Kant, Immanuel. Groundwork of the Metaphysics of Morals (1785). Quoted in James Rachels. The Elements of Moral Philosophy, 2d ed. New York: McGraw-Hill, 1993.
  • National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research. The Belmont Report: Ethical Principles and Guidelines for the Protection of Human Subjects of Research. 1978.
Commentary On

Case Overview

The issues in this case are not unique to science. For instance, it was not unusual to wonder what one could do with a Ph.D. in English 20 or 30 years ago. At the time, a Ph.D. in biology was considered a virtual guarantee of a job at a university, but that is no longer true today, if it ever was.

The employment prospects for those holding doctorates in the sciences is a difficult topic for scientists at any level to broach, but it is one that faculty, students and post- doctoral fellows need to discuss candidly. This case could serve as a catalyst to open that discussion. Certainly the topic is no longer taboo: For one thing, it is hard to ignore. Many are finding it difficult to find jobs appropriate to their training, and some science Ph.D.s are completing three or more post-doctoral appointments before finding something more permanent. Twenty years ago, graduate students who leaned toward college teaching careers (rather than the expected, research university professorship) knew they needed to be quiet about their interest in teaching. Now many science departments offer graduate courses on how to teach college-level science, and job ads require teaching experience. Even the NIH, as well as AAAS and other scientific societies, have recognized that the traditional tenure-track position at a research university is not what awaits most of our graduate students, and they are making efforts to explore and educate scientists about other career paths.

This case forces us to consider the responsibilities and expectations of many with regard to employment after graduate school (including the scientific community as a whole, university science departments, individual senior scientists who train students, and the students and post-docs themselves). Do we see graduate training in the sciences as primarily education and inculcation into a profession, or as preparation for future employment? The responsibilities one ascribes to each of the involved parties will tend to vary depending on one's perception of the primary role of graduate education in a scientific discipline.

In many ways, the issues in this case resemble the need for informed consent in research with human subjects, particularly the ethical mandate that we respect other people as persons like ourselves; that we respect their right to make their own decisions and direct the course of their lives. Along with giving people the freedom to choose, what is critical in this situation, just as in research with human subjects, is the information on which the decision is based -- its validity, completeness and clear communication.

Prospective graduate students need honest information about the current status of the academic job market as well as the availability of so-called alternative career paths. During their graduate work, they should be kept informed, offered opportunities to inform themselves and to get the training and experience that may be necessary for nonacademic careers. Faculty members need to keep up with the status of the job market and the concerns of their students. They need to talk about these issues with their students and post-docs, and to support them in considering and preparing for careers other than the traditional research university professorship. I assert that the responsibility for the gathering and exchange of information lies with both the science faculty and our students, but each student must be free to make his/her own decisions.

Back to Top

Discussion Questions

  1. If one considers Bowman to be a mature individual capable of making his own decisions, one must conclude that Hill's approach was paternalistic and inappropriate. In fact, he lied to Bowman. In addition to considering alternative ways in which Hill could have handled the conversation with Bowman, it would be beneficial to look at what happened in the faculty meeting as well. (Discussion of this point may be delayed; see Question 4.) Hill proposed that the department limit the number of students accepted for graduate study, and his suggestion was rejected. What are some other strategies he might have suggested? How could he have improved on his introduction of this topic at the faculty meeting? What are some other things that he might do within his department? In your discussion, be sure to note that from what we can tell from the case, Hill is acting on limited information (his conversation with Jake at the restaurant).
  2. Devorak has a lot of things she could discuss with Bowman. The question is what she should say in this phone conversation. She feels the tension of potentially conflicting obligations to herself, the university, Bowman and Hill. The possible topics range from the real reason for Hill's refusal to take Bowman on as a grad student, through the current job market, all the way to how she prefers to do her research and the pressures to get tenure. For each of these topics she could tell the complete truth, give Bowman an idea of what the situation is, lie or omit the topic from the conversation all together. In determining what she should say to Bowman, the most important consideration is what Bowman needs to know to make an informed decision at this time. Devorak need not disclose every detail about all of these topics, and some things may be better communicated later -- in a face-to-face meeting, perhaps, but at least after Bowman and Devorak get to know each other a little better. Recall that this is only their second phone conversation. We don't know how much time Bowman has before he must decide on other offers for graduate study, or if Hill and Devorak's department has set a deadline. However, it seems unlikely that Bowman and Devorak must decide on the best course of action today, in this phone call. Thus, Devorak should not lie to Bowman, but she should communicate to him the basic situation in her lab, and the possible problem with future employment, as far as she knows it. She should not discuss Hill (see Question 3). It would probably be best for all concerned if she gave herself and Bowman some time before definite decisions were made.
  3. Devorak should not tell Bowman that Hill lied about his reasons for refusing to accept Bowman as a graduate student. This issue is between Hill and Bowman, and Hill needs to be given the opportunity to explain his actions and his reasons. She can and should urge Hill to explain the situation to Bowman, and she should discuss concerns about future employment with Bowman, but she should not presume to speak for Hill. These conclusions are based in part on professional loyalty, the fact that one faculty member tries to avoid interfering in the interactions between other faculty members and their students. The idea of autonomy is also relevant here. Hill was free to decide to lie to Bowman, and he should be free to decide how he wants to handle the consequences, unless failure to be honest with Bowman about Hill's actions threatens to harm Bowman. If Devorak discusses the job market with Bowman, not mentioning Hill, potential harm to Bowman should be minimized, and Hill will be able to talk with Bowman later.
  4. These questions are similar to the ones posed in the discussion of Question 1 regarding Hill's handling of the presentation of his concerns to the departmental faculty. An individual, faculty or student, can make a difference, but he/she needs to be savvy and well prepared, and then recruit others to the cause. A brain-storming session that includes the design and evaluation of action plans would be an excellent way to address these questions. Keep in mind possible involvement of other departments, the university as a whole and professional societies. Coming up with a plan of action for Hill and Devorak to follow in their department, or perhaps deciding on something that your discussion group will do to address the employment issue, would be a good way to conclude discussion of this case.

Case Overview

Confidentiality and Conflict of Interest

Phase 1

Phase 2

Phase 3

Case Overview

In this commentary, I will restrict my comments to the topic of the peer review of a manuscript submitted to a scientific journal for publication. As noted in the preceding commentary, there are now clearly articulated procedures designed to minimize problems with confidentiality and conflict of interest in the review of grant proposals submitted to the NIH or NSF. There is far less explicit guidance or uniformity on these issues in the context of manuscript review.

For example, in their book on Ethics and Policy in Scientific Publication, the Editorial Policy Committee of the Council of Biology Editors report journal editors' responses to a number of scenarios dealing with scientific publishing. One of these scenarios concerned a reviewer who, without the editor's knowledge, routinely asked members of a research seminar to carry out group reviews of the manuscripts he received because of the educational value of this exercise. While "most of the respondents felt that the reviewer's practice was wrong" and one wrote, "Using someone else's work as a teaching exercise before it has been published is a violation of confidentiality," some respondents "did not object to the reviewer's practice." One wrote: "I agree with the reviewer. Such a procedure can have educational value. The reviewer should take responsibility for the comments, if he or she has compiled and edited them. The author should be grateful for the additional feedback." (Editorial Policy Committee, Council of Biology Editors, 1990, pp. 88-89)

Because of this variability, it is important for scientists to know that confidentiality and conflict of interest can be problems when reviewing a manuscript, that they consider these issues in advance, that they develop their own standards and guidelines, and that they make sure that they check each journal's policies.

Back to Top

Confidentiality and Conflict of Interest

In the context of pre-publication review of research results, confidentiality does not require complete secrecy. After all, the authors have discussed the results among themselves and in their research group(s), they have submitted them to the editorial board of a journal, and they may very well have presented some or all of the results in talks at their institution(s) and at scientific meetings. Rather, what is important is the authors' control of the information. Authors have the reasonable expectation that they are the ones who control who knows what and when concerning their work, and that editors will not widely distribute their manuscript during the review process. This expectation then extends to the individual reviewer.

In this case study, conflict of interest refers to interests that the reviewer may have that could bias the reviewer's judgment. Conflict of interest in a scientific context has been said to refer to "any conflict between research or other professional scientific judgments and financial or personal interests where acting with disregard to that conflict by placing one's personal or financial interests ahead of professional interests compromises or detrimentally influences professional judgment in conducting or reporting research." (Werhane and Doering 1997, 169-170)

And, of course, the review process is part of reporting research. In this context, it is also important to remember that the perception of a conflict of interest can be as potentially harmful as an actual conflict whether or not one's professional judgment is influenced. Disclosure is critical, and management and/or avoidance are very important.

Back to Top

Phase 1

Optimally, the editor in this scenario would have contacted Slater before sending the manuscript, to determine whether Slater had time to do the requested review and whether any potential problems existed. However, despite the advent of E-mail, many editors just put the manuscript in the mail; the potential reviewer learns of the assignment only when he/she opens the envelope. So one could ask, when, in the process of looking over the manuscript, should Slater start to hear little alarm bells going off in his head? Regardless of how little he reads, his actions will be changed by what he finds in that envelope. Seeing the title and author(s), he knows that the competition is ahead of him and that Parker is at risk of being scooped. One can argue that he probably is already aware of the situation from meetings and the gossip network of science. Even if that is the case, seeing it in print is indisputable evidence. By the time Slater reads through the abstract, he can be sure that others have already done what he and Parker are working on. Reading any further will only further complicate the situation.

Because of potential problems with a conflict of interest, which, although it is not financial, could bias his judgement, he should contact the editor who sent him the manuscript as soon as possible, and preferably before carefully reading the entire manuscript. An alternative would be to insert a note declining the editor's request to review the manuscript because of a conflict of interest, taping the envelope shut to avoid temptation, and returning it to the editor immediately. However, this alternative seems a bit extreme, although not ethically objectionable. What Slater needs is to obtain an outside opinion concerning his ability to be objective in his review. The editor is the best person to consult because confidentiality will be maximized, and Slater also needs to disclose the potential conflict of interest. Slater should not show the manuscript to a colleague and ask whether he/she thinks he should go ahead with the review, and he can't rely on his own potentially biased judgment to determine whether he can be objective.

Of course, the editor knows that Slater works in the same area as the authors of the manuscript. If that were not the case, the editor never would have asked Slater to review it. However, the editor may not be aware of how close this work is to Slater's own, nor may the author(s), who might otherwise have asked that Slater not review the manuscript. The editor needs to be contacted and made aware of the details of the situation. It is entirely possible that the editor may still think that Slater is the best choice as one of the reviewers, but now the conflict has been disclosed and the editor can factor this information into his/her weighing of Slater's review. Editors are frequently scientists who do research in the same area as the authors and reviewers of the manuscripts they handle; they may have some personal knowledge of the reviewer's reputation and past behavior to draw on.

Back to Top

Phase 2

Slater has decided not to consult with the editor and just to go ahead with his review. He also decides to ask Parker's opinion of the manuscript since she is the one who has the greatest expertise. It might be interesting to modify the scenario a bit and explore whether changes in what Slater tells or shows Parker change one's conclusion regarding the appropriateness of his action. For instance, Slater could have asked Parker for her opinion on a small part of the paper without actually showing her the manuscript or telling her why he was asking. Or he could have shown her only a portion of the manuscript, perhaps the Results section but not Methods. As the case is written, he simply gave her the complete manuscript saying something like, "I've been asked to review this for the Journal of Cool Results. Since you probably know more about this topic than I do, would you look it over and give me your opinion of it?"

Question 1. It is not easy to determine whether Slater should have shown the manuscript to Parker in any circumstances, with or without the editor's permission. Conflict of interest is involved as well as confidentiality. As soon as Slater tells Parker about the manuscript or shows it to her, she has new information that will affect her behavior, just as was the case for Slater opening that envelope in Phase 1. It is likely that she will push harder to finish her experiments and prepare her own manuscript. However, while Slater was able to choose how far he read, and thus how much he knew about the competition, before he stopped and possibly contacted the editor, Parker has less freedom to choose. Unless she interrupts Slater while he is telling her about the manuscript (and few grad students would do that), she will know whatever he decides to tell her. Once Slater asks her to read the manuscript, it may be difficult for Parker to tell her adviser that she doesn't think she should read it, even assuming she is aware that there could be a problem.

Thus, Parker is pulled into an awkward situation through no fault of her own, but rather through Slater's decision. By showing the manuscript to Parker, Slater has placed her in a very difficult conflict of interest and has breached the confidentiality of the review process. How much is her judgment to be trusted in evaluating a manuscript that would scoop publication of her own dissertation work? How can she ignore the information she now has, including new ideas on techniques or materials to try in purifying her protein? The former is a conflict of interest concern, the later is a change in the situation due to the breach of confidentiality.

As mentioned in the overview of the case, the scientific and editorial community exhibits some disagreement regarding the propriety of consulting others while preparing a review. Given this uncertainty, it would be best to consult the editor first to determine the policy of the journal, and to seek his/her approval for the proposed course of action. At the very least, the editor needs to be informed of Parker's participation so that the editor can appropriately evaluate the review that he/she receives.

Back to Top

Question 2. If a graduate student is to be involved in a review, he/she needs to be made aware of the obligation of confidentiality incumbent upon prepublication reviewers. This information is not intended for dissemination during lunch with other students, nor over the Internet. Faculty members should be familiar with the conventions of confidentiality, and that would be the major difference between consulting a colleague rather than a grad student. Another difference would be that faculty members have higher standing in the scientific community and so might be better able to weather any storm associated with the review of this manuscript by Slater, a close competitor of the author(s).

Back to Top

Phase 3

Now we come to the question of what Slater and Parker should do with the insight into their own research insight they have gained by reading their competitors' manuscript. They have learned that another group managed to purify the recombinant form of the protein growth factor, a problem whose solution has eluded Parker so far. But note that the case does not say that Parker has tried everything she can think of. Rather, she is in the process of troubleshooting the protocol. It seems that she has a number of ideas that she is checking, and she may even be planning to try a technique similar to that reported in the manuscript. So it is not clear how critical the acquired information is to her progress. The case also indicates that the purification of the protein is not an experimental test of the model Slater and Parker propose, but is more analogous to the preparation of a necessary reagent for a critical experiment.

However, Slater and Parker did not suddenly and unexpectedly find themselves in possession of information concerning the purification protocol. They should have expected to find that information in the paper. In fact, were I in Parker's situation reading anything other than a prepublication manuscript, it is the first thing I would check. Besides, completing a thorough review would require them to read and evaluate the protocol. So the knowledge of the competitors' protocol should have been expected, and it was gained during the review process. Now what should they do with the information?

The case indicates that Parker and Slater successfully used the competitors' protocol as presented in the manuscript, and that they were then able to finish the work and publish in the Journal of Cool Results -- quite possibly ahead of the competitors whose manuscript they reviewed. Even assuming that their recommendation against publication was justified and that the other reviewer(s) agreed with their evaluation, there is certainly the appearance of impropriety here, and quite possibly more (see Question 2 below).

Question 1. Parker should not have used the competitors' protocol step-by-step for a number of reasons. First, on a practical level, doing so obligates her to cite the source for the technique, which is very difficult in this situation. Failing to cite the source would be improperly claiming someone else's work as her own, a type of plagiarism. Writing the citation as "Competitor, et al., unpublished results" will not work, because most journals will not allow citation of unpublished work without written permission. Second, using the protocol breaches the trust placed in a peer reviewer: Slater and Parker are deriving personal benefit from the use of privileged information. Finally, the competitor(s) and/or others familiar with their work will probably be asked to review Slater and Parker's manuscript, and they will recognize their protocol and figure out what happened.

Some might argue that because the purification protocol is not a test of the proposed model and because it has no substantive scientific relevance to Slater and Parker's conclusions, its use is not an issue. Others might argue that Parker could have planned to try a procedure very similar to the competitors' before seeing their manuscript, and thus she did not steal their ideas or gain an unfair advantage. What is most important for this question is to determine the criteria for evaluating possible courses of action. What are the probable consequences; the possible benefits and possible harms to Parker, the competitors and others who might be affected by this situation? What are Parker and Slater's obligations? What values and principles are important in such a situation, and which courses of action are consistent with them?

Back to Top

Question 2. This question is very practical: How can Slater and Parker possibly ignore what they now know? Is it even ethical to try to do so? Should Parker continue with her troubleshooting, trying everything but the particular column, for instance, that her competitors used in the hope that she will find something else that will work? Wouldn't that be a waste of time and money, probably federal money? Frequently, people will justify taking some questionable shortcut as the right thing to do because it will maximize the benefit for science or society. One must examine these situations closely to determine whether the "good of science" is not really in fact "the advancement of my scientific career."

As noted the situation for Slater and the competing group changed as soon as Slater opened that envelope and read the title of the manuscript. The more he and then Parker read, the greater the danger that trust in the peer review process would be damaged. That is why I conclude that the editor must be contacted as soon as possible and informed that Slater will be talking with the competing group either as the manuscript is on its way back to the editor, or as soon as the review is completed. This approach has potential problems, but I believe that the potential harms can be minimized by open, frank communication among the affected parties (see Question 4), and the editor needs to be informed of Slater's plan. One hopes that the editor will endorse it.

Question 4. Contact with the competitors seems to be required to avoid a potentially nasty situation when Slater and Parker eventually publish and the competitors wonder if some of their ideas were appropriated during the review process.

Contact could be initiated in a number of ways, and the two groups could agree to a number of arrangements. For instance, instead of being open about why she was contacting them, Parker could indicate that she heard through the grapevine that the competitors were working on a similar purification and ask them to give her information about their results -- for which she would cite them, of course. It seems that a more honest approach might work better, but that may depend on the history of interactions between these groups. Having made contact, the two groups might simply agree to cite each other, or perhaps they will agree to publish back-to-back in the Journal of Cool Results. They might even agree to a formal collaboration. A discussion and evaluation of various ways one might contact and cooperate with a competitor would be a useful way to conclude a discussion of this case.

References

  • Editorial Policy Committee, Council of Biology Editors. Ethics and Policy in Scientific Publication. Bethesda, Md.: Council of Biology Editors, Inc., 1990.
  • Werhane, P., and Doering, J. "Conflicts of Interest and Conflicts of Commitment" in D. Elliott and J. E. Stern, eds. Research Ethics: A Reader. Hanover, N. H.: University Press of New England, 1997, pp. 169-170.
Commentary On

Case Overview

An initial reading of this case might lead one see it as "simply" a case dealing with the issue of authorship. However, further reflection reveals that it hinges on the larger issue of the system of responsibility and reward in the laboratory, and how this system is communicated to and understood by all the laboratory members, including the PI. At an even more basic level, the essence of the problem here is a lack of communication.

In discussing the case, it would be very instructive to spend some time exploring the relevant obligations of all the major characters, and if, and how, the fulfillment of these obligations should be linked to rewards such as authorship. For instance, Smith had an obligation to carry out and document careful research that others could build on, and Johnson had a responsibility to supervise Smith's work and to review his results. Both failed in their responsibilities. Now we are asked to determine what happens to the reward, authorship. In Part 2, Johnson asserts that by failing to fulfill his obligation to the lab, Smith has given up his right to a co-authorship.

The responsibility-reward system will vary from lab to lab, yet is central to the scientific enterprise. In the discussion of this case, it will be thought-provoking to have participants share the systems in their labs, if they even know them, and then discuss what the linkage should be.

Another interesting aspect of this case is that a graduate student, Jill Green, has been put in the middle of a dispute over appropriate attribution for another's work. My experience indicates that this occurrence is not infrequent, but it is one that we do not usually discuss. A brain-storming session on what Jill might do, followed by an evaluation of the probable consequences of each suggestion, would be very valuable to graduate students who may find themselves in such a situation in the future. Green has an obligation to communicate honestly with all involved, but she must be savvy enough to do so without harming herself.

The case presents some ambiguous aspects that are interesting to play with. I list a few below.

How novel was the reagent the Smith said he used? What was the probability that any similarly trained chemist would have tried the same reagent? Was it likely that Green would have come up with the idea on her own, and Smith's only contribution was to save her time?

What type of information should Smith have had to "back up his claim?" How much is enough? What are the criteria? Who makes the determination?

Where is Smith now employed? Did he get the job based on a recommendation from Johnson? Is Smith doing research similar to that done in Johnson's lab? Is he in an academic position, possibly training graduate students?

Green noted that "Smith's experimental procedures were poorly written" and that "it was not possible to duplicate his work." Was this problem just sloppiness, or was it sloppiness that crossed the line into negligence by a man who claimed to be a professional scientist? Was there any indication of fraud? How would and should the determination of sloppiness vs. negligence vs. fraud affect the evaluation of other aspects of this case?

In this research group, would a temporary post-doc have been considered an employee or a colleague? It could have been a research group associated with a chemical company at which Green was doing her research and at which Smith was employed. If it was an academic research group, did Smith sign a release form concerning patents?

Back to Top

Discussion Questions

Phase 1

Question 1. Johnson and Green should have informed Smith of Green's results and their submission of the manuscript not because it is mandated by some professional code, but just as a matter of common courtesy to a colleague, even if he could have been considered an employee. Science depends on communication in all modes, not just the formal, published paper. Avoiding communicating with Smith may have seemed the easiest thing to do in the short run, but it can lead to more unpleasant consequences in the long run and is disrespectful of Smith as a person. If Smith had been listed as a co-author on the paper, he must be contacted. Authorship involves acceptance of responsibility for the contents of the paper, and Smith must be able to choose whether he will take on this responsibility.

Question 2. Smith's contribution could have been acknowledged in a variety of ways other than a co-authorship. Generally today, authorship represents recognition for a significant intellectual contribution to the published work. Some other possible modes of attribution are an acknowledgment, a footnote or a citation as an unpublished result. The criteria for these other forms of attribution are no more clearly formulated than is the definition of a significant contribution, and to make matters worse they vary from lab to lab, and from discipline to discipline. It may be useful to check the instructions to authors for a number of prominent journals in your field to see whether they provide guidelines. A comparative discussion of criteria and standard practice among the discussion participants will help everyone to look more critically at what they have accepted as the norm, and to consider what the criteria should be, and why.

Question 3. Without further information, it is not possible to determine whether Smith should have claimed to have solved the problem. One would need to know what items of documentation were in his notebooks and what the criteria for a solution were. However, among a focused group such as a research lab group or a class of beginning grad students, it would be very beneficial to discuss what sorts of documentation one should have, and how the monitoring of research progress should be carried out. In short, how could this problem have been avoided?

Back to Top

Discussion Questions

Phase 2

Question 1. Answering this question takes us back to the issues addressed under Question 2 in Phase 1 and in the overview above: What did Smith contribute? What was its the significance of his contribution? What would be the appropriate attribution for what Smith did? Once some of the uncertainties have been clarified by arbitrarily defining a few of the variables in the case, one could begin to discuss the appropriate way to acknowledge his contributions.

For example, let us assume that: 1) The reagent was not one that just any chemist would have thought to try in this situation, but it was not completely unknown. 2) Smith's notes were almost illegible, and included only one NMR analysis run on the products of the critical reaction, but it didn't look as if he had falsified or fabricated anything in his notes. In this situation, I would conclude that Smith's contribution, while not significant enough to warrant an authorship, does require some form of acknowledgment, probably in the form of an acknowledgment at the end of the paper.

Question 2. One might not expect a patent lawyer to raise ethical arguments. However, it would be a good idea to raise the issue of patents, particularly in the field of chemistry. What arguments could be made for including Dr. Smith on the patent?

Notice that in the text of the case, Johnson seems reluctant to acknowledge Smith in any way partly because he feels it would obligate him to include Smith on a patent application. I doubt that that would be the case, but many people worry unnecessarily about the ramifications of their actions on the distribution of royalties from possible patents. Practice concerning patents has varied with time and institution; it is best to consult with those concerned with patents to determine the relevant policy is. Researchers should have some basic information about this issue. In fact, this case provides an excellent opportunity to ask an official who deals with patents to join the discussion.

In either a commercial company or an academic university, the institution is the entity that makes the patent application. Scientists may share in the royalties, depending on the practice of the institution. It is my experience that university scientists sign patent wavers along with other employment papers when they start work; I have observed quite a bit of variation among heads of laboratories in the royalties distribution concerning inclusion of graduate students and post-docs who worked on the project. It is not unusual for the royalties to be split between the PI and the university. Many PIs feel that they have fulfilled their obligation to others in their labs if they use the royalties to fund further work in the lab.

With regard to the case involving Smith, Johnson and Green, it should be noted that patents are awarded for practical applications, not ideas. Thus, it is the tested machine or the process that is patented, not an unproven idea. Smith's contribution was, at best, an idea of a reagent to use; Green worked out the details of the reaction conditions. Smith does not have a good case for being included on the patent, but Green does.

Question 3. It is difficult for me to see how having Smith leave his place of employment and return to the lab would solve any of the problems in this case, unless he were being invited back to try to replicate results in his notebook and so prove that they were not fabricated. I suppose that one possible compromise that the parties in this case could have reached was to have Smith try to get the other reaction conditions written in his notes to work, document his results and then have them included with Green's in a revised manuscript on which Smith would be a co-author. This strategy seems awkward, but possible.

It might be useful to change this question into an opportunity to brainstorm possible solutions to the problem as it now exists for Smith, Johnson and Green, and then to investigate the probable consequences of each.

Back to Top

Overview

This case challenges us to consider one of the fundamental questions in ethics: How ought people to treat each other? Yes, it's a human subjects case that emphasizes the issue of informed consent. However, if we only talk about "human subjects," we can too easily forget what this term really means, and that is people, like ourselves. Here, the subjects are 10,000 women who were initially recruited for the study when they were pregnant five to ten years ago. Those considering the ethical issues in the case are two senior researchers who have collaborated on the study since its beginning. In their discussions in this case, Smith seems to concentrate on the research subject aspect of those who participated in the study while Jones tends to view them more as individual people.

At one level, this case can be discussed as a means to clarify what the regulations are and how one goes about fulfilling them. That is a necessary part of educating researchers in the proper procedures to follow in human subject research, and should not be overlooked. However, this approach leads one to discuss rules and legalities more than ethics and can result in protocols that, while technically correct, do not truly treat the subjects respectfully and/or fail to consider all the possible consequences of the study. To get at the ethical issues, one must move from just considering the letter of the regulations to their spirit; to exploring the ethical principles that underlie all those rules. For research involving human subjects, these are the three principles delineated in the Belmont Report:

  1. Respect for persons involves a recognition of the personal dignity and autonomy of individuals. . .
  2. Beneficence entails an obligation to protect persons from harm by maximizing anticipated benefits and minimizing possible risks of harm.
  3. Justice requires that the benefits and burdens of research be distributed fairly. (National Commission, 1979)

If researchers got into the habit of looking beyond the rule itself to consider the reason for the rule, and then asked whether their protocols and procedures were consistent with the underlying principles, a lot of the difficulties that have come up in human subjects research could be avoided.

Discussion Questions

Part 1

The question asks whether the researchers are obligated to recontact the participants for their consent before doing a genetic analysis on the specimens donated by the participants for the previous study. The short answer is "yes," both for legal and ethical reasons.

I use the term "previous" study because it has "been five years since the end of data collection," and I assume that the analysis of these data and presentation of the researchers' interpretations at meetings and in journals would have already occurred. Thus, the primary work on the study is probably completed, but the samples are still in the freezer. I make this point because, acting on a point made by Wallace (1982) and noted in the IRB Guidebook (NIH 1993), most current IRBs would require that the experimental protocol specify what would happen to the data and specimens at the conclusion of the study. That does not seem to have been the case here, and probably was not required for most older studies. The fate of the data and specimens should be a concern of all researchers carrying out research using human subjects.

Why all this concern about a few blood samples and a few proposed "look-see" preliminary tests? After all, scientists are trained to keep samples just in case something comes up in the future such as a question about the previous work, or an opportunity to extend the investigation previously conducted. Most scientists have freezers and storage cabinets bulging with materials from earlier research. In this case, it is important to note that those 10,000 blood samples that Smith has in her freezer are not the same as the thousands of slices from a sediment core drilled out of a lake bed that a geologist across campus may have carefully stored in her cabinet. Smith's samples come from human beings, and therefore the issues of respect for persons and confidentiality need to be considered as well as scientific merit when new analyses are contemplated.

The women who participated in the study have a right to have a say in what is done with these specimens taken from their bodies. It is interesting to note that some universities have gone so far as to broaden the official, federal definition of human subjects to include living individuals "as well as human embryos, fetuses, cadavers, and any human tissue or fluids." (Indiana University, 1997) This definition emphasizes that these research materials are different from others scientists may use in their work, like the sediment core noted above, and that these samples must be treated with respect. The concept of respect for persons comes largely out of the writings of Immanuel Kant who asserted that one should:

Act so that you treat humanity, whether in your own person or in that of another, always as an end and never as a means only. (Kant, 1785)

In this case, Smith is in danger of forgetting that the participants in the study are people, not just a means to advance her research. I'm sure that she has good intentions, but she can't skip asking permission of the study participants just because she believes the end will be beneficial for science and society.

Genetic analyses were not listed in the original consent form, and they yield information that is fundamentally different from that yielded by the analyses listed. The original analyses looked at characteristics that were environmental in nature; things that could be changed by treatment or modification of the women's behavior. The genetic analyses could yield information about characteristics of the women over which they have no control, and that could be passed on to their children. This information could cause harm if it became known to either the women themselves, or to others with whom they interact such as employers or insurance companies. Even for oneself, genetic self-knowledge is not always a benefit. It can be a burden if no available corrective treatment is available, and particularly if the knowledge is unwanted and is forced upon a person. What Smith proposes is not just an extension of the previous analyses. The women have a right to decide whether the genetic analyses are to be done on their samples, particularly in light of possible consequences.

The issue of confidentiality will be discussed further in the commentary on later questions in this case.

Back to Top

Part 2

Question 1. In response to Jones' objections, Smith suggests that they just destroy the list that links the study participants' names with the other data collected in the study. That would seem to solve the problem of confidentiality since no one could find out the participants' names after the list was destroyed. In addition, it would seem to qualify their work for an exemption from the need for informed consent based on the federal regulations. This suggestion raises several issues.

First, the proposed course of action is disingenuous. It's not the case that the researchers never had the participants' names. They have simply destroyed them because it seemed the easiest short-term solution. This approach is not the best way to enhance the public's trust of scientists. Also, as Jones points out, those names might be needed later if they decide to pursue a follow-up study.

Second, destroying the names does not really eliminate the linkage between those samples in the freezer and the women who donated them. The relevant regulations state that an exemption from the requirement for informed consent may be granted

. . . if the information is recorded by the investigator in such a manner that subjects can not be identified, directly or through identifiers linked to the subjects. (45 CFR 46.101(b)(4)).

However, that is not the situation here because of the other data collected by these researchers. What the regulations refer to as "other identifiers" could still be used to link the genetic results to the women. Information that might facilitate the linkage are such items as maternal age, date of delivery, days preterm, baby's weight, location of interview and delivery hospital. It would not take too much work by a private investigator to identify the women to whom these data referred. Thus, confidentiality must still be a concern of the researchers, and they do not qualify for an exemption from the requirement for informed consent.

Question 2. This question is largely addressed in the commentary for Question 1 above, but there is an additional reason why the researchers should contact the participants for their consent before doing the genetic analysis: The genetic analysis poses potential risks that differ from those presented by the earlier analyses. If confidentiality is broken somehow, the women could be burdened with self-knowledge that they did not want; they may lack support in dealing with this information; and their new knowledge could have a negative impact on them and their children. Depending on the genetic markers that Smith selects for her work, the information derived may simply enable molecular geneticists to estimate the risk of premature delivery for the participants and their daughters.

However, if genetic markers with known associations with other diseases or conditions were used, the information could be even more damaging. For instance, if Smith checks for the alleles at the BRCA1 locus, one of the so-called breast cancer genes, she may or may not gather data that will help predict if a woman is in danger of giving preterm birth, but she will have information that bears on the current health of these women and their blood relatives. This information could be devastating to a woman, especially if she did not expect to receive it, and could be detrimental to her future employment and insurability if it became known to others. Because of these considerations, the new informed consent form would need to be different from the original one in order to address the potential risks associated with genetic analyses. The women should also be asked to indicate if they want to be informed of any information the researchers may gather that could have an impact on the participants' health. They have a right to determine what information will be gathered about them, and to control who will have access to the information, including themselves.

Question 3. Asking whether the researchers have an obligation to inform the women of the results of the genetic analysis assumes that the researchers went ahead with the analyses without obtaining additional consent from the study participants. Of course, it would be best to avoid the dilemma posed by discovering that Mary Brown, for instance, has an 80 percent chance of breast and/or ovarian cancer, but not knowing how she will react if you send her a letter containing this information. This scenario could be avoided by getting additional consent, before doing the analyses, in which one also asks the women if they want to be informed of the results of the genetic tests.

Assuming that they didn't ask, Smith and Jones must now consider the possible consequences of their decision to contact Mary Brown with the bad news. Here it is valuable to take a closer look at the principle of beneficence.

In the Belmont Report, the committee discussed what they meant by the principle of beneficence as it applies to research involving human subjects.

In this document, beneficence is understood . . . as an obligation. Two general rules have been formulated as complementary expressions of beneficent actions in this sense: (1) do not do harm and (2) maximize possible benefits and minimize possible harms. (NIH, 1993)

Note that the first rule listed is "Do not do harm." Many philosophers assert that one has a greater obligation to avoid doing harm than to attempt to do good (Frankena, 1973), and that is an important consideration in this case. What "goods" might the researchers do by carrying out the analyses and then contacting a participant if they discovered something they felt she should know? What "harms" might occur as a result of this course of action? In the "goods" column, a list might include advancing scientific knowledge for the benefit of society, and aiding in the health care of study participants. In the "harms" column would go such things as psychological harm to the participants who are told bad news, and risk to future employment and insurability if confidentiality is broken. Many of the potential benefits are to science, society and the researchers, while almost all of the potential harms are to the study participants themselves. Thus, one must consider both the principle of beneficence and the principle of justice in answering this question. In my analysis, the potentials for harm and injustice were so much greater than the potential benefits that I would argue that the woman should not be informed of the results of the genetic analysis unless they have said they wanted to know.

Back to Top

Part 3

In a situation like this, a vague statement on a consent form is of benefits to no one, although it might be technically in compliance with the regulations. One must question whether the researchers are really demonstrating a respect for the participants when the consent process is so vague. In addition, the vague wording, while sanctioning the genetic analyses Smith wants to do, does nothing to help her resolve the dilemma of how to respond if data are found that could have an impact on the participants' health. For the benefit of the participants, the researchers and the enterprise of science, it is best to be as through and forthright as possible in the consent form and in the recruiting interviews with the potential subjects.

Back to Top

References

  • Kant, Immanuel. Groundwork of the Metaphysics of Morals (1785), as quoted in James Rachels. The Elements of Moral Philosophy, 2d ed. New York: McGraw-Hill, 1993.
  • National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research. The Belmont Report: Ethical Principles and Guidelines for the Protection of Human Subjects of Research. 1979.
  • National Institutes of Health, Office of Extramural Research, Office for Protection from Research Risks. Protecting Human Research Subjects: Institutional Review Board Guidebook, Second Edition. 1993.
  • National Institutes of Health, Office of Extramural Research, Office for Protection from Research Risks, "Summary of the Belmont Report" in Protecting Human Research Subjects, Institutional Review Board Guidebook, 2d ed. 1993.
  • U. S. Department of Health and Human Services. "Protection of Human Subjects." Code of Federal Regulations Title 45, Part 46 (rev. 1991).
  • Wallace, R. Jay. "Privacy and the Use of Data in Epidemiology." In Tom L. Beauchamp, Ruth R. Fallon, R. Jay Wallace, Jr, and LeRoy Walters, eds. Ethical Issues in Social Research. Baltimore: Johns Hopkins University Press, 1982, pp. 274-291.