DNA Technology in Forensic
DNA Technology in Forensic
DNA Technology in Forensic
DETAILS
CONTRIBUTORS
GET THIS BOOK Committee on DNA Technology in Forensic Science, National Research Council
SUGGESTED CITATION
Visit the National Academies Press at NAP.edu and login or register to get:
Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.
DNA Technology in
Forensic Science
ii
iii
Former Members
C. THOMAS CASKEY (Resigned December 21, 1991), Baylor College of
Medicine, Houston, Texas
MICHAEL W. HUNKAPILLER (Resigned August 17, 1990), Applied
Biosystems Inc., Foster City, California
iv
BOARD ON BIOLOGY
vi
PREFACE vii
Preface
In recent years, advances in the techniques for mapping and sequencing the
human genome have contributed to progress in both basic biology and medicine.
The applications of these techniques have not been restricted to biology and
medicine, however, but have also entered forensic science. Today, methods
developed in basic molecular biology laboratories can potentially be used in
forensic science laboratories in a matter of months.
On the basis of its study of the mapping and sequencing of the human
genome (reported in 1988), the Board on Biology and several federal agencies
recognized the potential of DNA typing technology for forensic science. In
particular, the Federal Bureau of Investigation, the preeminent organization in the
United States for the development and application of forensic techniques, initiated
an effort to develop and evaluate DNA typing in forensic applications in the
mid-1980s. The first case work was performed in December 1988. Several
private-sector laboratories entered the field early, and state government crime
laboratories also began to offer services in DNA typing. However, as DNA typing
entered the courtrooms of this country, questions appeared about its reliability
and methodological standards and about the interpretation of population
statistics.
By the summer of 1989, a crescendo of questions concerning DNA typing
had been raised in connection with some well-publicized criminal cases, and calls
for an examination of the issues by the National Research Council of the
National Academy of Sciences came from the scientific and legal communities.
As a response, this study was initiated in January 1990.
Because of the broad ramifications of forensic DNA typing, a number
PREFACE viii
of federal agencies and one private foundation provided financial support for this
study: the Federal Bureau of Investigation, the National Institutes of Health
National Center for Human Genome Research, the National Institute of Justice,
the National Science Foundation, the State Justice Institute, and the Alfred P.
Sloan Foundation.
Many persons offered assistance to the committee and staff during this
complex study. In particular, the following deserve recognition and praise for
their efforts: John Hicks, Federal Bureau of Investigation; Elke Jordan and Eric
Juengst, National Institutes of Health National Center for Human Genome
Research; James K. Stewart, Charles B. DeWitt, Bernard V. Auchter, and Richard
Laymon, National Institute of Justice; John C. Wooley, National Science
Foundation; David I. Tevelin, State Justice Institute; and Michael S. Teitelbaum,
Alfred P. Sloan Foundation.
I also thank the many experts who offered their advice to the committee
during its briefings and open meetings. The names of those who offered
testimony are given in the appendix. Additionally, I want to thank the many who
wrote to me or to the National Research Council and provided valuable data and
suggestions to the committee; much was gained from their input. We also
acknowledge the efforts of Robert Kushen, Columbia Law School, in assisting
Judge Weinstein. I also thank Della Malone, my secretary, for her help
throughout. The committee thanks the reviewers of our report for many valuable
comments and suggestions. Although the reviewers are anonymous to us, I
personally want to thank them for their constructive comments and suggestions.
The staff of the Board on Biology deserve special praise for their efforts
during the many months of intense activity. Oskar R. Zaborsky, Study Director
and Director of the Board on Biology, deserves recognition for his administrative
and technical contributions and for handling many complex matters. Marietta
Toal, Administrative Secretary, served the committee well in logistics and the
preparation of the report. The committee also thanks Mary Kay Porter for her
assistance. Norman Grossblatt edited the report.
Last but not least, I thank my colleagues on the committee who served so
well and unselfishly to address key issues from the perspective of their special
expertise and to prepare this report in a timely fashion.
DNA typing for personal identification is a powerful tool for criminal
investigation and justice. At the same time, the technical aspects of DNA typing
are vulnerable to error, and the interpretation of results requires appreciation of
the principles of population genetics. These considerations and concerns arising
out of the felon DNA databanks and the privacy of DNA information made it
imperative to develop guidelines and safeguards for the most effective and
socially beneficial use of this powerful tool. We hope that our efforts will
enhance understanding of the issues and serve to
PREFACE ix
bring together people of good will from science, technology, law, and ethics. We
hope that our report will serve well the sponsors and the general public.
Victor A. McKusick
Chairman
Committee on DNA Technology in Forensic Science
On April 14, 1992, The New York Times printed an article on this report.
That article seriously misrepresented the findings of the committee; in an article
on April 15, the Times corrected the misrepresentation. To avoid any potential
confusion engendered by the April 14 article, the committee provides the
following clarifying statement:
We recommend that the use of DNA analysis for forensic purposes,
including the resolution of both criminal and civil cases, be continued while
improvements and changes suggested in this report are being made. There is no
need for a general moratorium on the use of the results of DNA typing either in
investigation or in the courts.
We regard the accreditation and proficiency testing of DNA typing
laboratories as essential to the scientific accuracy, reliability, and acceptability of
DNA typing evidence in the future. Laboratories involved in forensic DNA
typing should move quickly to establish quality-assurance programs. After a
sufficient time for implementation of quality-assurance programs has passed,
courts should view quality control as necessary for general acceptance.
The Committee
CONTENTS xi
Contents
SUMMARY 1
1 INTRODUCTION 27
Background 27
Genetic Basis of DNA Typing 32
Technological Basis of DNA Typing 36
Population Genetics Relevant to the Interpretation of DNA Typing 44
Characteristics of an Optimal Forensic DNA Typing System 48
References 49
CONTENTS xii
CONTENTS xiii
GLOSSARY 167
PARTICIPANTS 179
INDEX 179
SUMMARY 1
Summary
SUMMARY 2
technology, management of DNA typing data, and legal, societal, and ethical
issues surrounding DNA typing. The techniques of DNA typing are fruits of the
revolution in molecular biology that is yielding an explosion of information
about human genetics. The highly personal and sensitive information that can be
generated by DNA typing requires strict confidentiality and careful attention to
the security of data.
DNA, the active substance of the genes, carries the coded messages of
heredity in every living thing: animals, plants, bacteria, and other
microorganisms. In humans, the code-carrying DNA occurs in all cells that have a
nucleus, including white blood cells, sperm, cells surrounding hair roots, and
cells in saliva. These would be the cells of greatest interest in forensic studies.
Human genes are carried in 23 pairs of chromosomes, long threadlike or
rodlike structures that are a person's archive of heredity. Those 23 pairs, the total
genetic makeup of a person, are referred to as the human "diploid genome." The
chemistry of DNA embodies the universal code in which the messages of heredity
are transmitted. The genetic code itself is spelled out in strings of nucleotides of
four types, commonly represented by the letters A, C, G, and T (standing for the
bases adenine, cytosine, guanine, and thymine), which in various combinations of
three nucleotides spell out the
SUMMARY 3
codes for the amino acids that constitute the building blocks of proteins. A gene,
the basic unit of heredity, is a sequence of about 1,000 to over 2 million
nucleotides. The human genome, the total genetic makeup of a person, is
estimated to contain 50,000-100,000 genes.
The total number of nucleotides in a set of 23 chromosomes—one from each
pair, the "haploid genome"—is about 3 billion. Much of the DNA, the part that
separates genes from one another, is noncoding. Variation in the genes, the
coding parts, are usually reflected in variations in the proteins that they encode,
which can be recognized as "normal variation" in blood type or in the presence of
such diseases as cystic fibrosis and phenylketonuria; but variations in the
noncoding parts of DNA have been most useful for DNA typing.
Except for identical twins, the DNA of a person is for practical purposes
unique. That is because one chromosome of each pair comes from the father and
one from the mother; which chromosome of a given pair of a parent's
chromosomes that parent contributes to the child is independent of which
chromosome of another pair that parent gives to that child. Thus, the different
combinations of chromosomes that one parent can give to one child is 223, and the
number of different combinations of paired chromosomes a child can receive from
both parents is 246.
The substitution of even one nucleotide in the sequence of DNA is a
variation that can be detected. For example, a variation in DNA consisting of the
substitution of one nucleotide for another (such as the substitution of a C for a T)
can often be recognized by a change in the points at which certain biological
catalysts called "restriction enzymes" cut the DNA. Such an enzyme cuts DNA
whenever it encounters a specific sequence of nucleotides that is peculiar to the
enzyme. For example, the enzyme HaeI cuts DNA wherever it encounters the
sequence AGGCCA. A restriction enzyme will cut a sample of DNA into
fragments whose lengths depend on the location of the cutting sites recognized by
the enzyme. Assemblies of fragments of different lengths are called "restriction
fragment length polymorphisms" (RFLPs), and RFLPs constitute one of the most
important tools for analyzing and identifying samples of DNA.
An important technique used in such analyses is the "Southern blot,"
developed by Edwin Southern in 1975. A sample of DNA is cut with a restriction
enzyme, and the fragments are separated from one another by electrophoresis
(i.e., they are separated by an electrical field). The fragments of particular interest
are then identified with a labeled probe, a short segment of single-stranded DNA
containing a radioactive atom, which hybridizes (fuses) to the fragments of
interest because its DNA sequence is complementary to those of the fragments (A
pairing with T, C pairing with G). Each electrophoretic band represents a separate
fragment of DNA, and a given person will have no more than two fragments
derived from a partic
SUMMARY 4
ular place in his or her DNA—one representing each of the genes that are present
on the two chromosomes of a given pair. The forms of a given gene are referred
to as alleles. A person who received the same allele from the mother and the
father is said to be "homozygous" for that allele; a person who received different
alleles from the mother and the father is said to be "heterozygous." Many RFLP
systems are based on change in a single nucleotide. They are said to be
"diallelic," because there are only two common alternative forms. And there are
only three genotypes: two kinds of homozygous genotypes and a heterozygous
genotype. Another form of RFLP is generated by the presence of variable number
tandem repeats (VNTRs). VNTRs are sequences, sometimes as small as two
different nucleotides (such as C and A), that are repeated in the DNA. When such a
structure is subjected to cutting with restriction enzymes, fragments of varied
length are obtained.
It was variation of the VNTR type to which Alex Jeffreys in the United
Kingdom first applied the designation "DNA fingerprinting." He used probes that
recognized not one locus, but multiple loci, and "DNA fingerprinting" has come
to refer particularly to multilocus, multiallele systems. A locus is a specific site of a
gene on a chromosome. In the United States, in particular, single-locus probes are
preferred, because their results are easier to interpret. "DNA typing" is the
preferred term, because "DNA fingerprinting'' is associated with multilocus
systems. Discriminating power for personal identification is achieved by using
several—usually at least four—single-locus, multiallelic systems.
The entire procedure for analyzing DNA with the RFLP method is
diagrammed in Figure 2.
After the bands (alleles) are visualized, those in the evidence sample and the
suspect sample are compared. If the bands match in the two samples, for all three
or four enzyme-probe combinations, the question is: What is the probability that
such a match would have occurred between the suspect and a person drawn at
random from the same population as the suspect?
Answering that question requires calculation of the frequency in the
population of each of the gene variants (alleles) that have been found, and the
calculation requires a databank where one can find the frequency of each allele in
the population. On the basis of some assumptions, so-called Hardy-Weinberg
ratios can be calculated. For a two-allele system, the ratios are indicated by the
expressions p2 and q2 for the frequency of the two homozygotes and 2pq for
heterozygotes, p and q being the frequencies of the two alleles and p + q being
equal to 1. Suppose that a person is heterozygous at a locus where the frequencies
of the two alleles in the population are 0.3 and 0.7. The frequency of that
heterozygous genotype in the population would be 2 × 0.3 × 0.7 = 0.42. Suppose,
further, that at three
SUMMARY 5
other loci the person being typed has genotypes with population frequencies of
0.01, 0.32, and 0.02. The frequency of the combined genotype in the population
is 0.42 × 0.01 × 0.32 × 0.02 or 0.000027, or approximately 1 in 37,000.
SUMMARY 6
typing can rely on methods of detection that do not use radioactive substances.
Furthermore, the technique of PCR amplification permits the use of very small
samples of tissue or body fluids—theoretically even a single nucleated cell.
The PCR process (Figure 3) is simple; indeed, it is analogous to the process
by which cells replicate their DNA. It can be used in conjunction with various
methods for detecting person-to-person differences in DNA.
It must be emphasized that new methods and technology for demonstrating
individuality in each person's DNA are being developed. The present methods
explained here will probably be superseded by others that are more efficient,
error-free, automatable, and cost-effective. Care should be taken to ensure that
DNA typing techniques used for forensic purposes do not become "locked in"
prematurely, lest society and the criminal justice system be unable to benefit fully
from advances in science and technology.
TECHNICAL CONSIDERATIONS
The forensic use of DNA typing is an outgrowth of its medical diagnostic
use—analysis of disease-causing genes based on comparison of a patient's DNA
with that of family members to study inheritance patterns of genes or comparison
with reference standards to detect mutations. To understand the challenges
involved in such technology transfer, it is instructive to compare forensic DNA
typing with DNA diagnostics.
DNA diagnostics usually involves clean tissue samples from known
sources. Its procedures can usually be repeated to resolve ambiguities. It involves
comparison of discrete alternatives (e.g., which of two alleles did a child inherit
from a parent?) and thus includes built-in consistency checks against artifacts. It
requires no knowledge of the distribution of patterns in the general population.
Forensic DNA typing often involves samples that are degraded,
contaminated, or from multiple unknown sources. Its procedures sometimes
cannot be repeated, because there is too little sample. It often involves matching
of samples from a wide range of alternatives in the population and thus lacks
built-in consistency checks. Except in cases where the DNA evidence excludes a
suspect, assessing the significance of a result requires statistical analysis of
population frequencies.
Each method of DNA typing has its own advantages and limitations, and
each is at a different state of technical development. However, the use of each
method involves three steps:
SUMMARY 7
FIGURE 3
Polymerase chain reaction (PCR). Courtesy, Perkin-Elmer Cetus Instruments.
SUMMARY 8
whether the types match and thus whether the samples could have
come from the same source.
3. If the types match, statistical analysis of the population frequencies
of the types to determine the probability that a match would have
been observed by chance in a comparison of samples from different
persons.
Before any particular DNA typing method is used for forensic purposes,
precise and scientifically reliable procedures for performing all three steps must
be established. It is meaningless to speak of the reliability of DNA typing in
general—i.e., without specifying a particular method.
Despite the challenges of forensic DNA typing, it is possible to develop
reliable forensic DNA typing systems, provided that adequate scientific care is
taken to define and characterize the methods.
Recommendations
SUMMARY 9
SUMMARY 10
SUMMARY 11
a common gene pool. Under that assumption, the procedure for calculating the
population frequency of a genotype is straightforward:
• Count the frequency of alleles. For each allele in the genotype, examine a
random sample of the population and count the proportion of matching
alleles—that is, alleles that would be declared to match according to the
rule that is used for declaring matches in a forensic context.
• Calculate the frequency of the genotype at each locus. The frequency of a
homozygous genotype a1/a1 is calculated to be pa12, where pa1denotes
the frequency of allele a1. The frequency of a heterozygous genotype
a1/a2 is calculated to be 2pa1pa2, where pa1 and pa2 denote the
frequencies of alleles a1 and a2. In both cases, the genotype frequency is
calculated by multiplying the two allele frequencies, on the assumption
that there is no statistical correlation between the allele inherited from
one's father and the allele inherited from one's mother. When there is no
correlation between the two parental alleles, the locus is said to be in
Hardy-Weinberg equilibrium.
• Calculate the frequency of the complete multilocus genotype by
multiplying the genotype frequencies at all the loci. As in the previous
step, this calculation assumes that there is no correlation between the
genotypes at the individual loci; the absence of such correlation is called
linkage equilibrium. Suppose, for example, that a person has genotype
a1/a2, b1/b2, c1/c1. If a random sample of the appropriate population
shows that the frequencies of alleles a1, a2, b1, b2, and c1 are
approximately 0.1, 0.2, 0.3, 0.1, and 0.2, respectively, then the
population frequency of the genotype would be estimated to be [2(0.1)
(0.2)][2(0.3)(0.1)][(0.2)(0.2)] = 0.000096, or about 1 in 10,417.
SUMMARY 12
SUMMARY 13
forensic DNA typing to undergo much change over the next decade—
including the introduction of different types of DNA
polymorphisms, some of which might have different properties from
the standpoint of population genetics.
4. It is desirable to provide a method for calculating population
frequencies that is independent of the ethnic group of the subject.
SUMMARY 14
markers tested so far cannot be taken to mean that such does not exist for other
markers. Preservation of population DNA samples in the form of immortalized
cell lines will ensure that DNA is available for determining population
frequencies of any DNA pattern as new and better techniques become available,
without the necessity of collecting fresh samples. It will also provide samples for
standardization of methods across laboratories.
Because of the similarity in DNA patterns between relatives, databanks of
DNA of convicted criminals have the ability to point not just to individuals but to
entire families—including relatives who have committed no crime. Clearly, this
raises serious issues of privacy and fairness. It is inappropriate, for reasons of
privacy, to search databanks of DNA from convicted criminals in such a fashion.
Such uses should be prevented both by limitations of the software for searching
and by statutory guarantees of privacy.
The genetic correlation among relatives means that the probability that a
forensic sample will match a relative of the person who left it is considerably
greater than the probability that it will match a random person.
Especially for a technology with high discriminatory power, such as DNA
typing, laboratory error rates must be continually estimated in blind proficiency
testing and must be disclosed to juries.
Recommendations
SUMMARY 15
groups and the highest of these values or 10% (whichever is the larger)
should be used. Data on at least three major "races" (e.g., Caucasians,
blacks. Hispanics, Asians, and Native Americans) should be analyzed.
• Any population databank used to support DNA typing should be openly
available for scientific inspection by parties to a legal case and by the
scientific community.
• Laboratory error rates should be measured with appropriate proficiency
tests and should play a role in the interpretation of results of forensic
DNA typing.
STANDARDS
Critics and supporters of the forensic uses of DNA typing agree that there is a
lack of standardization of practices and a lack of uniformly accepted methods for
quality assurance. The deficiencies are due largely to the rapid emergence of DNA
typing and its introduction in the United States through the private sector.
As the technology developed in the United States, private laboratories using
widely differing methods (single-locus RFLP, multilocus RFLP, and PCR) began
to offer their services to law-enforcement agencies. During the same period, the
FBI was developing its own RFLP method, with a different restriction enzyme
and different single-locus probes. The FBI's method has become the one most
widely used in public forensic-science laboratories. Each method has its own
advantages and disadvantages, databanks, molecular-weight markers, match
criteria, and reporting methods.
Regardless of the causes, practices in DNA typing vary, and so do the
educational backgrounds, training, and experience of the scientists and
technicians who perform the tests, the internal and external proficiency testing
conducted, the interpretation of results, and approaches to quality assurance.
It is not uncommon for an emerging technology to go without regulation
until its importance and applicability are established. Indeed, the development of
DNA typing technology has occurred without regulation of laboratories and their
practices, public or private. The committee recognizes that standardization of
practices in forensic laboratories in general is more problematic than in other
laboratory settings; stated succinctly, forensic scientists have little or no control
over the nature, condition, form, or amount of sample with which they must
work. But it is now clear that DNA typing methods are a most powerful adjunct
to forensic science for personal identification and have immense benefit to the
public—so powerful, so complex, and so important that some degree of
standardization of laboratory procedures is necessary to assure the courts of
high-quality results. DNA typing is capable, in principle, of an extremely low
inherent rate of false results, so the risk of error will come from poor laboratory
SUMMARY 16
practice or poor sample handling and labeling; and, because DNA typing is
technical, a jury requires the assurance of laboratory competence in test results.
At issue, then, is how to achieve standardization of DNA typing laboratories
in such a manner as to assure the courts and the public that results of DNA typing
by a given laboratory are reliable, reproducible, and accurate.
Quality assurance can best be described as a documented system of activities
or processes for the effective monitoring and verification of the quality of a work
product (in this case, laboratory results). A comprehensive quality-assurance
program must include elements that address education, training, and certification
of personnel; specification and calibration of equipment and reagents;
documentation and validation of analytical methods; use of appropriate standards
and controls; sample handling procedures; proficiency testing; data interpretation
and reporting; internal and external audits of all the above; and corrective actions
to address deficiencies and weight their importance for laboratory competence.
Recommendations
Although standardization of forensic practice is difficult because of the
nature of the samples, DNA typing is such a powerful and complex technology
that some degree of standardization is necessary to ensure high standards.
SUMMARY 17
SUMMARY 18
SUMMARY 19
Recommendations
• In the future, if pilot studies confirm its value, a national DNA profile
databank should be created that contains information on felons convicted
of particular violent crimes. Among crimes with high rates of
recidivism, the case is strongest for rape, because perpetrators typically
leave biological evidence (semen) that could allow them to be
identified. Rape is the crime for which the databank will be of primary
use. The case is somewhat weaker for violent offenders who are most
likely to commit homicide as a recidivist offense, because killers leave
biological evidence only in a minority of cases.
• The databank should also contain DNA profiles of unidentified persons
made from biological samples found at crime scenes. These would be
samples known to be of human origin, but not matched with any known
persons.
• Databanks containing DNA profiles of members of the general
population (as exist for ordinary fingerprints for identification purposes)
are not appropriate, for reasons of both privacy and economics.
• DNA profile databanks should be accessible only to legally authorized
persons and should be stored in a secure information resource.
• Legal policy concerning access and use of both DNA samples and DNA
databank information should be established before widespread
proliferation of samples and information repositories. Interim protection
and sanctions against misuse and abuse of information derived from
DNA typing should be established immediately. Policies should
explicitly define authorized uses and should provide for criminal
penalties for abuses.
• Although the committee endorses the concept of a limited national DNA
profile databank, it doubts that existing RFLP-based technology
provides an appropriate wise long-term foundation for such a databank.
We expect current methods to be replaced soon with techniques that are
sim
SUMMARY 20
SUMMARY 21
expert evidence should be handled before and during a trial to ensure prompt and
effective adjudication apply to all evidence and all experts.
In the United States, there are two main tests for admissibility of scientific
information through experts. One is the Frye test, enunciated in Frye v. United
States. The other is a "helpfulness" standard found in the Federal Rules of
Evidence and many of its state counterparts. In addition, several states have
recently enacted laws that essentially mandate the admission of DNA typing
evidence.
The test for the admissibility of novel scientific evidence enunciated in Frye
v. United States is still probably the most frequently invoked test in American
case law. A majority of states profess adherence to the Frye rule, although a
growing minority have adopted variations on the helpfulness standard suggested
by the Federal Rules of Evidence.
Frye predicates the admissibility of novel scientific evidence on its general
acceptance in a particular scientific field: "While courts will go a long way in
admitting expert testimony deduced from a well-recognized scientific principle or
discovery, the thing from which the deduction is made must be sufficiently
established to have gained general acceptance in the particular field in which it
belongs." Thus, admissibility depends on the quality of the science underlying the
evidence, as determined by scientists themselves. Theoretically, the court's role in
this preliminary determination is narrow: it should conduct a hearing to determine
whether the scientific theory underlying the evidence is generally accepted in the
relevant scientific community and whether the specific techniques used are
reliable for their intended purpose.
In practice, the court is much more involved. The court must determine the
scientific fields from which experts should be drawn. Complexities arise with
DNA typing, because the full typing process rests on theories and findings that
pertain to various scientific fields. For example, the underlying theory of
detecting polymorphisms is accepted by human geneticists and molecular
biologists, but population geneticists and other statisticians might differ as to the
appropriate method for determining the population frequency of a genotype in the
general population or in a particular geographic, ethnic, or other group. The
courts often let experts on a process, such as DNA typing, testify to the various
scientific theories and assumptions on which the process rests, even though the
experts' knowledge of some of the underlying theories is likely to be at best that
of a generalist, rather than a specialist.
The Frye test sometimes prevents scientific evidence from being presented
to a jury unless it has sufficient history to be accepted by some subspecialty of
science. Under Frye, potentially helpful evidence may be excluded until
consensus has developed. By 1991, DNA evidence had been considered in
hundreds of Frye hearings involving felony prosecutions in
SUMMARY 22
more than 40 states. The overwhelming majority of trial courts ruled that such
evidence was admissible, but there have been some important exceptions.
In determining admissibility according to the helpfulness standard under the
Federal Rules of Evidence, without specifically repudiating the Frye rule, a court
can adopt a more flexible approach. Rule 702 states that, "if scientific, technical
or other specialized knowledge will assist the trier of fact to understand the
evidence or to determine a fact in issue, a witness qualified as an expert by
knowledge, skill, experience, training, or education, may testify thereto in the
form of an opinion or otherwise."
Rule 702 should be read with Rule 403, which requires the court to
determine the admissibility of evidence by balancing its probative force against
its potential for misapplication by the jury. In determining admissibility, the court
should consider the soundness and reliability of the process or technique used in
generating evidence; the possibility that admitting the evidence would
overwhelm, confuse, or mislead the jury; and the proffered connection between
the scientific research or test result to be presented and particular disputed factual
issues in the case.
The federal rule, as interpreted by some courts, encompasses Fryeby making
general acceptance of scientific principles by experts a factor, and in some cases a
decisive factor, in determining probative force. A court can also consider the
qualifications of experts testifying about the new scientific principle, the use to
which the technique based on the principle has been put, the technique's potential
for error, the existence of specialized literature discussing the technique, and its
novelty.
With the helpfulness approach, the court should also consider factors that
might prejudice the jury. One of the most serious concerns about scientific
evidence, novel or not, is that it possesses an aura of infallibility that could
overwhelm a jury's critical faculties. The likelihood that the jury would abdicate
its role as critical fact-finder is believed by some to be greater if the science
underlying an expert's conclusion is beyond its intellectual grasp. The jury might
feel compelled to accept or reject a conclusion absolutely or to ignore evidence
altogether. However, some experience indicates that jurors tend not to be
overwhelmed by scientific proof and that they prefer experiential data based on
traditional forms of evidence. Moreover, the presence of opposing experts might
prevent a jury from being unduly impressed with one expert or the other.
Conversely, the absence of an opposing expert might cause a jury to give too
much weight to expert testimony, on the grounds that, if the science were truly
controversial, it would have heard the opposing view. Nevertheless, if the
scientific evidence is valid, the solution to those possible problems is not to
exclude the evidence, but to ensure through instructions and testimony that the
jury is equipped to consider rationally whatever evidence is presented.
SUMMARY 23
Recommendations
SUMMARY 24
SUMMARY 25
whose interests are to count and whether some people's interests should be given
greater weight than others'. For example, there are the interests of the accused, the
interests of victims of crime or their families in apprehending and convicting
perpetrators, and the interests of society. Whether the interests of society in
seeing that justice is done should count as much as the interests of the accused or
the victim is open to question.
A major issue is the preservation of confidentiality of information obtained
with DNA technology in the forensic context. When databanks are established in
such a way that state and federal law-enforcement authorities can gain access to
DNA profiles, not only of persons convicted of violent crimes but of others as
well, there is a serious potential for abuse of confidential information. The
victims of many crimes in urban areas are relatives or neighbors of the
perpetrators, and these victims might themselves be former or future perpetrators.
There is greater likelihood that DNA information on minority-group members,
such as blacks and Hispanics, will be stored or accessed. However, it is important
to note that use of the ceiling principle removes the necessity to categorize
criminals (or defendants in general) by ethnic group for the purposes of DNA
testing and storage of information in databanks.
The introduction of a powerful new technology is likely to set up
expectations that might be unwarranted or unrealistic in practice. Various
expectations regarding DNA typing technology are likely to be raised in the
minds of jurors and others in the forensic setting. For example, public perception
of the accuracy and efficacy of DNA typing might well put pressure on
prosecutors to obtain DNA evidence whenever appropriate samples are available.
As the use of the technology becomes widely publicized, juries will come to
expect it, just as they now expect fingerprint evidence.
Two aspects of DNA typing technology contribute to the likelihood of its
raising inappropriate expectations in the minds of jurors. The first is a jury's
perception of an extraordinarily high probability of enabling a definitive
identification of a criminal suspect; the second is the scientific complexity of the
technology, which results in laypersons' inadequate understanding of its
capabilities and failings. Taken together, those two aspects can lead to a jury's
ignoring other forensic evidence that it should be considering.
As large felon databanks are created, the forensic community could well
place more reliance on DNA evidence, and a possible consequence is the
underplaying of other forensic evidence. Unwarranted expectations about the
power of DNA technology might result in the neglect of relevant evidence.
The need for international cooperation in law enforcement calls for
appropriate scientific and technical exchange among nations. As in other areas of
science and technology, dissemination of information about DNA
SUMMARY 26
typing and training programs for personnel likely to use the technology should be
encouraged. It is desirable that all nations that will collaborate in law-
enforcement activities have similar standards and practices, so efforts should be
furthered to exchange scientific knowledge and expertise regarding DNA
technology in forensic science.
Recommendations
INTRODUCTION 27
1
Introduction
BACKGROUND
Characterization, or ''typing," of blood, semen, and other body fluids has
been used for forensic purposes for more than 50 years.1 It began with blood
groups, such as those of the ABO system, and later was extended to serum
proteins and red-cell enzymes and in some forensic applications, particularly
paternity testing, to human leukocyte antigens (HLA), which are associated with
tissue types. The genetically determined person-to-person variation revealed by
such typing was used mainly to include or exclude suspects, that is, to determine
whether a person showed a combination of genetically determined characteristics
consistent with having been the source of an evidence sample in a criminal case
or having been the father of a child in a paternity case. Except when HLA testing
was used, the chance that a randomly chosen person would be excluded by the
tests was about 98%; that left a 2% chance that the test would "include" an
innocent person.
In the last decade, methods have become available for deoxyribonucleic acid
(DNA) typing, that is, for showing distinguishing differences in the genetic
material itself. Advances in DNA technology in the 1970s paved the way for the
detection of variation (polymorphism) in specific DNA sequences and shifted the
study of human variation from the protein products of DNA to DNA itself. By
analyzing a sufficient number of regions of DNA that show much person-to-
person variability, one can reduce the probability of a chance match (inclusion)
of two persons to an extremely low
INTRODUCTION 28
level. Indeed, the probability can, in principle, be made so low that DNA typing
becomes not simply a method for exclusion or inclusion, but a means of absolute
identification.
The potential applicability of DNA typing to forensic samples was
demonstrated during the mid-1980s by laboratories in the United Kingdom,
United States, and Canada. Their work established that DNA was present in
forensic samples in sufficient quantity for testing (see Table 1.1) and that it
survived in a state that allowed it to be typed. In the publications in 1985 by
Jeffreys and colleagues,2,3 the term ''DNA fingerprint" carried the connotation of
absolute identification. The mass-media coverage that accompanied the
publications fixed in the general public's minds the idea that DNA typing could
be used for absolute identification. Thus, the traditional forensic paradigm of
genetic testing as a tool for exclusion was in a linguistic stroke changed to a
paradigm of identification. (See Box 1 for a contrasting of dermatoglyphic
fingerprints with "DNA fingerprints.")
Forensic DNA typing, first used in casework in 1985 in the United
Kingdom, was initiated in the United States in late 1986 by commercial
laboratories and in 1988 by the Federal Bureau of Investigation (FBI) and is now
being used by dozens of state and local crime laboratories. Because of its great
potential benefits for criminal and civil justice, but also because of the
possibilities for its misuse or abuse, forensic DNA typing has been subjected to
special scrutiny. Important questions have been asked about reliability, validity,
and confidentiality:4-6
TABLE 1.1 DNA Content of Biological Samples
Type of Sample Amount of DNAa
Blood 20,000-40,000 ng/ml
stain 1 cm2 in area ca. 200 ng
stain 1 mm2 in area ca. 2 ng
Semen 150,000-300,000 ng/ml
postcoital vaginal swab 0-3,000 ng
Hair:
plucked 1-750 ng/hair
shed 1-12 ng/hair
Saliva 1,000-10,000 ng/ml
Urine 1-20 ng/ml
a The amount of DNA is given in nanograms (ng); 1 ng = one-billionth of a gram (10-9 g).
INTRODUCTION 29
FINGERPRINTS IN PERSONAL
IDENTIFICATION:DIFFERENCES BETWEEN DNA TYPING
AND DERMATOGLYPHICS
Because of the central role of dermatoglyphic fingerprints in human
identification, arising out of the personal uniqueness of the patterns, it is
useful to compare and contrast traditional fingerprints with "DNA
fingerprints".
Fingerprints were described as an individualizing characteristic as early
as 1892. The use of fingerprints in forensic science (and in relation to
chromosonal abnormalities, such as Down syndrome, and other clinical
disorders) was developed empirically without reference to the specific
genetic basis of patterns. Ridge count is a polygenetic or multifactoral trait.
The close correlation for ridge counts with that expected for an almost
exclusively genetic trait, when "identical" and fratenal twins and other
relatives are compared, supports polygenetic inheritance.
In the forensic application, minutiae in the fingerprint patterns, not ridge
counts, are used for personal identification. The minutiae result from random
nongenetic events during embryonic development of the fingerpads. As a
consequence, the patterns even of "identical" twins are distinguishable.
Indeed, it appears that the fingerprint pattern of each human being is
unique.
The distinction between two types of fingerprints is illustrated by prints
from "identical" twins shown here. The dermatoglyphic fingerprints shown in
figure B-a (Twin A) and figure B-B (Twin B) are distinguished by the
patterns of loops and whorls.
Twin A Twin B
Finger Right Hand
Thumb (1) Whorl Whorl
2 Whorl Central pocket Whorl
3 Ulnar loop Whorl
4 Whorl Whorl
5 Ulnar loop Whorl
Finger Left Hand
Thumb (1) Whorl Whorl
2 Whorl Ulnar loop
3 Ulnar loop Ulnar loop
4 Whorl Ulnar loop
5 Ulnar loop Ulnar loop
However, typing of DNA from the blood of these twins in three
laboratories showed a match for all tests. One laboratory, testing for
variation in four chromosones, estimated the population frequency of
INTRODUCTION 30
INTRODUCTION 31
INTRODUCTION 32
INTRODUCTION 33
INTRODUCTION 34
INTRODUCTION 35
number tandem repeats), the number of repetitions of a sequence can vary from
person to person. VNTRs are a leading form of variation used currently in
forensic DNA typing. The repeating unit can be as small as a dinucleotide—e.g.,
the (TG)n polymorphism—or as large as 30, or even more, nucleotides. Tandem
repeats are not limited to noncoding segments of DNA, although they are found
less frequently in coding segments.
The two main types of variation—single-nucleotide differences and VNTRs
—are both potentially recognizable by change in the lengths of fragments that
result when DNA is cut with a restriction enzyme. Variation in the lengths of
fragments can result from a change in the cluster of four, five, or six nucleotides
that is the specific cutting site of the particular restriction enzyme (Figure 1-2).
Or the variation can result, not from a change in the cutting site of the enzyme,
but from the existence of different numbers of tandem repeats between two
cutting sites. Figure 1-3 diagrams the major characteristics of the two forms of
variation and their use in DNA typing.
INTRODUCTION 36
INTRODUCTION 37
DNA with a single restriction enzyme; each fragment has a distinct sequence and
length. For analysis of RFs to demonstrate RFLPs, the fragments are separated
electrophoretically on the basis of size. Electrophoresis, typically performed on
agarose or acrylamide gels, results in large fragments at one end and small
fragments at the other; the small fragments migrate farthest in the electric field.
The fragments are then denatured (i.e., rendered single-stranded), neutralized, and
transferred from the gel to a nylon membrane, to which they are fixed; this
facilitates detection of specific RFLPs and VNTRs.
INTRODUCTION 38
INTRODUCTION 39
INTRODUCTION 40
Because the extension products of one primer bind the other primer in
successive cycles, there is in principle a doubling of the target sequence in each
cycle. However, the efficiency of amplification is not 100%, and the yield from a
30-cycle amplification is generally about 106-107 copies of the
INTRODUCTION 41
INTRODUCTION 42
INTRODUCTION 43
INTRODUCTION 44
INTRODUCTION 45
FIGURE 1-8 Distribution of gene for blood group B in Europe. Similar wide
variation is observed with genes for blood groups A and O. From Mourant et
al.,38 p. 266.
INTRODUCTION 46
INTRODUCTION 47
INTRODUCTION 48
root of the frequency of the homozygotes for the particular allele or half the
summed frequencies of the heterozygotes, which, in the case of the five-allele
system, will be of four types for each allele.
Population Substructure
It is intuitively obvious that relatives have genes in common. Thus, the
chance that DNA typing will yield a match with a suspect when the evidence
sample in fact came from a brother or other relative is considerable, especially if
only a few loci are tested.
The U.S. population is a conglomerate of many different population groups,
which might be viewed as extended families derived from all parts of the world.
The major ethnic groupings—white, black, Hispanic, Asian, etc.—are each
composites of many different subpopulations, which might have quite different
frequencies of the alleles used in forensic DNA typing. Allele frequencies
estimated from sampling of an overall ethnic group represent weighted averages.
Some of the component subpopulations might have allele frequencies quite
different from the mean values of the whole population. Further discussion of this
problem and recommendations for handling it are given in Chapter 3.
INTRODUCTION 49
of the genome. Even if a locus is only modestly polymorphic, its use in DNA
typing could have other advantages, such as complete unambiguity of scoring;
used in combination, such loci could demonstrate that the chance of a random
match is extremely low.
It must be emphasized that new methods and technology for demonstrating
individuality in each person's DNA continue to be developed. The methods
outlined in this chapter are likely to be superseded in efficiency, automatability,
economy, and other features by new methods. Care must be taken to ensure that
DNA typing techniques used for forensic purposes do not become "locked in"
prematurely. Otherwise, society and the criminal justice system will not be able to
derive maximal benefit from advances in the science and technology.
REFERENCES
1. Gaensslen RE. Sourcebook in forensic serology, immunology, and biochemistry. Washington,
D.C.: U.S. Government Printing Office, 1983.
2. Jeffreys AJ, Wilson V, Thein SL. Individual-specific "fingerprints" of human DNA. Nature.
316:75-79, 1985.
3. Gill P, Jeffreys AJ, Werrett DJ, Forensic application of DNA "fingerprints." Nature. 318:577-579,
1985.
4. Evett IW, Werrett DJ, Gill P, Buckleton JS. DNA fingerprinting on trial. Nature. 340:435, 1989.
5. Lander ES. DNA fingerprinting on trial. Nature. 339:501-505, 1989.
6. U.S. Congress, Office of Technology Assessment. Genetic witness: forensic uses of DNA tests.
OTA-BA-438. Washington, D.C.: U.S. Government Printing Office, 1990.
7. Thompson L. A smudge on DNA fingerprinting? The Washington Post, Monday, June 26, 1989.
8. Barinaga M. DNA fingerprinting: pitfalls come to light, Nature. 339:89, 1989.
9. Roychoudhury AK, Nei M. Human polymorphism genes: world distribution, New York: Oxford
University Press, 1988.
10. Cooper DN, Smith BA, Cooke HJ, Niemann S, Schmidtke J. An estimate of unique DNA
sequence heterozygosity in the human genome. Hum Genet. 69:201-205, 1985.
11. Kirby LT. DNA fingerprinting: an introduction, New York: Stockton Press, 1990.
12. Farley MA, Harrington JJ, eds.: Forensic DNA technology. Chelsea, Michigan: Lewis Publishers,
1991.
13. U.S. Department of Justice, Federal Bureau of Investigation. Proceedings of the international
symposium on the forensic aspects of DNA analysis. Washington, D.C.: U.S. Government
Printing Office, 1991.
14. Southern EM. Detection of specific sequences among DNA fragments separated by gel
electrophoresis. J Mol Biol.98:503-527, 1975.
15. Botstein D, White RL, Skolnick M, Davis RW. Construction of a genetic linkage map in man
using restriction fragment length polymorphisms. Am J Hum Genet. 32:314-331, 1980.
16. Helminen P, Ehnholm C, Lokki ML, Jeffreys A, Peltonen L. Application of DNA "fingerprints" to
paternity determinations. Lancet. 1:574-576, 1988.
17. Jeffreys AJ, Wilson V, Thein SL, Weatherall DJ, Ponder BAJ. DNA "fingerprints" and
segregation analysis of multiple markers in human pedigrees. Am J Hum Genet. 39:11-24,
1986.
INTRODUCTION 50
18. Higuchi R, yon Beroldingen CH, Sensabaugh GF, Erlich HA. DNA typing from single hairs.
Nature. 332:543-546, 1988.
19. Jeffreys AJ, Wilson V, Neumann R, Keyte J. Amplification of human minisatellites by the
polymerase chain reaction: towards DNA fingerprinting of single cells. Nucleic Acids Res.
16:10953-10971, 1988.
20. Erlich IIA, ed. PCR technology: principles and applications for DNA amplification. New York:
Stockton Press, 1989.
21. Rose EA. Applications of the polymerase chain reaction to genomc analysis. FASEB J.5:46-54,
1991.
22. Amheim N, Levenson CH. Polymerase chain reaction. Chem Eng News. 68:36-47, October 1,
1990.
23. Mullis KB. The unusual origin of the polymerase chain reaction. Sci Am.262:56-65, 1990.
24. Conner B J, Reyes AA, Morin C, Itakura K, Teplitz RL, Wallace RB. Detection of' sickle-cell
beta-S-globin allele by hybridization with synthetic oligonucleotides. Proc Natl Acad Sci
LISA. 80:278-282, 1983.
25. Chehab FF, Kan YW. Detection of specific DNA-sequences by fluorescence amplifica-tion--a
color complementation assay. Proc Natl Acad Sci USA. 86:9178-9182, 1989.
26. Newton CR, Graham A, Heptinstall LC, Porvell SI, Summers JC, Markham AF. Analysis of any
point mutation in DNA—the amplification refractory mutation system (ARMS). Nucleic
Acids Res. 17:2503-2516, 1989.
27. Wu DY, Ugozzoli L, Pal BK, Wallace RB. Allele-specific enzymatic amplification of beta-globin
genomic DNA for diagnosis of sickle-cell anemia. Proc Natl Acad Sci LISA. 86:2757-2760,
1989.
28. Nickerson DA, Kaiser R, Lappin S, Steward J, Hood L, Landegren U. Automated DNA
diagnostics using an ELISA-based oligonucleotide ligation assay. Proc Natl Acad Sci LISA.
87:8923-8927, 1990.
29. Kasai K, Nakamura Y, White R. Amplification of a variable number of tandem repeats (VNTR)
locus (pMCT118) by the polymerase chain reaction (PCR) and its application to forensic
science. J For Sci. 35:1196-1200, 1990.
30. Myers RM, Maniatis T, Lerman I,S. Detection and localization of single base changes by
denaturing gradient gel electrophoresis. Methods Enzymol. 155:501-527, 1987.
31. Cotton RGH, Rodriques NR, Campbell RD. Reactivity of cytosine and thymine in single-base-
pair mismatches with hydroxylamine and osmium-tetroxide and its application to the study
of mutations. Proc Natl Acad Sci USA. 85:4397-4401, 1988.
32. Nakamura Y, Leppert M, O'Connell P, Wolff R, Holm T, Culver M, Martin C, Fujimoto E, Hoff
M, Kumlin E, White R. Variable number of tandem repeat (VNTR) markers for human gene
mapping. Science. 235:1616-1622, 1987.
33. Saiki R, Bugawan TL, Horn GT, Mullis KB, Erlich HA. Analysis of enzymatically amplified [3-
globin anti HLA-DQ DNA with allele-specific oligonucleotide probes. Nature.
324:163-166, 1986.
34. Saiki R, Walsh PS, Levenson CH, Erlich HA. Genetic analysis of amplified DNA with
immobilized sequence-specific oligonucleotide probes. Proc Natl Acad Sci USA.
86:6230-6234, 1989.
35. liagelberg E, Gray IC, Jeffreys AJ. Identification of the skeletal remains of a murder by DNA
analysis. Nature. 352:427-428, 1991.
36. Jeffreys A, MacLeod A, Tamaki K, Nell D, Monckton D. Minisatellite repeat coding as a digital
approach to DNA typing. Nature. 354:204-209, 1991.
37. Hartl DL, Clark AC. Principles of population genetics. 2nd ed. Sunderland, Massachusetts:
Sinauer Associates, 1989.
38. Mourant AE, Kopec AC, Domaniewska-Sobczak K. The distribution of the human blood groups
and other polymorphisms. 2nd ed. Oxford: Oxford University Press, 1976.
2
DNA Typing: Technical Considerations
"DNA typing" is a catch-all term for a wide range of methods for studying
genetic variations. Each method has its own advantages and limitations, and each
is at a different state of technical development. Each DNA typing method
involves three steps:
Before any particular DNA typing method is used for forensic purposes, it is
essential that precise and scientifically reliable procedures be established for
performing all three steps. This chapter discusses the first two—laboratory
analysis and pattern comparison—and Chapter 3 focuses on statistical analysis.
There is no scientific dispute about the validity of the general principles
underlying DNA typing: scientists agree that DNA varies substantially among
humans, that variation can be detected in the laboratory, and that DNA
comparison can provide a basis for distinguishing samples from different
persons. However, a given DNA typing method might or might not be
scientifically appropriate for forensic use. Before a method can be ac
Scientific Foundations
The forensic use of DNA typing is an outgrowth of its medical diagnostic
use—analysis of disease-causing genes based on comparison of a patient's DNA
with that of family members to study inheritance patterns of genes or with
reference standards to detect mutations. To understand the challenges involved in
such technology transfer, it is instructive to compare forensic DNA typing with
DNA diagnostics.
DNA diagnostics usually involves clean tissue samples from known
sources. It can usually be repeated to resolve ambiguities. It involves comparison
of discrete alternatives (e.g., which of two alleles did a child inherit from a
parent?) and thus includes built-in consistency checks against artifacts. It requires
no knowledge of the distribution of patterns in the general population.
Forensic DNA typing often involves samples that are degraded,
contaminated, or from multiple unknown sources. It sometimes cannot be
repeated, because there is too little sample. It often involves matching of samples
from a wide range of alternatives present in the population and thus lacks built-in
consistency checks. Except in cases where the DNA evidence
systems in common forensic use, in which a single sample might yield somewhat
different allele sizes on repeat measurements.1 It is easy to determine whether two
samples match in the former case (assuming that the patterns have been correctly
identified), but the latter case requires a match criterion—i.e., an objective and
quantitative rule for deciding whether two samples match. For example, a match
criterion for VNTR systems might declare a match between two samples if the
restriction-fragment sizes lie within 3% of one another.
The match criterion must be based on the actual variability in measurement
observed in appropriate test experiments conducted in each testing laboratory.
The criterion must be objective, precise, and uniformly applied. If two samples lie
outside the matching rule, they must be declared to be either ''inconclusive" or a
"nonmatch." Considerable controversy arose in early cases over the use of
subjective matching rules (e.g., comparison by eye) and the failure to adhere to a
stated matching rule.
Experiential Foundation
Before a new DNA typing method can be used, it requires not only a solid
scientific foundation, but also a solid base of experience in forensic application.
Traditionally, forensic scientists have applied five steps to the implementation of
genetic marker systems:2,3
Choice of Probes
A DNA probe used in forensic applications should have the following
properties:
• It should recognize a single human locus (or site), preferably one whose
chromosomal location has been determined.
• It should detect a constant number of bands per allele in most humans.
• It should be characterized in the published literature, including its
typical range of alleles, and its tendency to recognize DNA from other
species.
• It should be readily available for scientific study by any interested
person.
a clean result that shows the correct pattern for a particular hybridization, the
result of the test hybridization should be discounted.
Single-Band Patterns
Sometimes, only a single band will be detected when two distinct alleles are
present. That might occur because the second allele is so small that it has
migrated off the end of the gel, because the second allele is similar in size to the
first allele and thus is not resolved, or because the second allele is much larger,
and larger fragments are preferentially lost in partially degraded samples.
When only a single band is found, the interpretation should always include
the possibility that a second band has been missed—i.e., that the pattern is
actually of a heterozygote, not a homozygote. (For statistical interpretation, the
frequency of a single-band pattern should be taken to be the sum of the
frequencies of all patterns containing this band. This is approximately twice the
allele frequency of the band.) In some cases, it could be important to interpret the
absence of a second larger fragment—e.g., when two samples match in a smaller
band, but the questioned sample lacks a second larger band. That could arise
either because the samples are from different persons or because the samples
come from the same person but the questioned sample is partially degraded.
Ideally, to distinguish these alternatives, one should determine whether a second
larger band could have been detected in the questioned sample by hybridizing the
membrane with a single-copy probe that detects an even larger monomorphic
fragment—i.e., one that is constant in all humans. In contrast, it would not be
sufficient simply to estimate the degree of degradation from the ethidium bromide
staining pattern of the sample.
Anomalous Bands
A sample might show more than two bands for various reasons. E.g., the
hybridization conditions were improper and caused the probe to hybridize to
incorrect fragments; the probe was contaminated with another sequence, which
caused it to recognize other fragments; the membrane was incompletely stripped
after a previous use, so a pattern seen on the previous hybridization is still being
detected; the restriction digestion did not proceed to completion, so the region
recognized by the probe is present in incompletely cut fragments of multiple
sizes; or the sample actually contains a mixture of multiple DNAs. The last
example is extremely important to recognize, because it can bear importantly on a
case. Whenever extra bands are observed, their origin should be determined.
The following clues provide a partial decision tree:
Reporting of Anomalies
Examiners should document their interpretations of samples thoroughly in
writing. They should note all observed bands and any questionable densities that
they do not consider to be bands. Anomalous bands should be explained on the
basis of appropriate control experiments of the sorts described above.
Measurement of Fragments
Molecular-weight measurements of fragments should initially be made by
comparing band positions with known molecular-weight standards run in separate
lanes on the same gel (so-called external molecular-weight stan
Match Criteria
Current RFLP-based tests use VNTR probes that have dozens of closely
spaced alleles. On the one hand, the high degree of polymorphism increases the
power of the test to detect differences among persons. On the other hand, the
large number of alleles increases the complexity of matching samples, because
gels have little ability to resolve nearby alleles (which can differ by as little as 9
basepairs, so that, for practical purposes, the distribution of alleles can appear to
be continuous).
Because of the limited resolution, two samples from a single person will
often lead to slightly different measurements—e.g., 3.00 and 2.45 kilo-bases (kb)
in one case, 3.03 and 2.40 kb in another. To decide whether two samples match,
each laboratory must have a match criterion.6 The match criterion should provide
an objective and quantitative rule for deciding whether two patterns match—e.g.,
all fragments must lie within 2% of one another. When samples fall outside the
match criterion, they should be declared to be "inconclusive" or "nonmatching."
The match criterion must be based on reproducibility studies that show the
actual degree of variability observed when multiple samples from the same
person are separately prepared and analyzed under typical forensic
conditions. Some testing laboratories originally used matching rules that were
based on the average spacing of fragment sizes in each region of the gel, rather
than on actual studies of reproducibility. Other laboratories used purely visual
matching criteria. Both are inadequate. Each testing laboratory must carry out its
own reproducibility studies, because reproducibility varies among laboratories.
The precise match criterion of each laboratory should be made freely available to
all interested persons and should be stated in forensic reports.
The match criterion is also used in the calculation of allele frequencies. To
determine the probability that a matching allele was found by chance, one counts
the number of matching alleles in an appropriately chosen reference population.
For the calculation to be valid, the same match criterion must be applied in
screening the population databank and in comparing the forensic samples. Some
testing laboratories originally used less stringent rules for declaring a match
between forensic samples and more stringent rules for determining the frequency
of matching alleles in the databank; the effect was an overstatement of the
probability of obtaining a match by chance.
Some have advocated that testing laboratories, instead of using a match
criterion, should report a likelihood ratio—the ratio of the probability that the
measurements would have arisen if the samples came from the same person to the
probability that they would have arisen if they came from different persons. No
testing laboratories in the United States now use that approach. The committee
recognizes its intellectual appeal, but recommends against it. Accuracy with it
requires detailed information about the joint distribution of fragment positions,
and it is not clear that information about a match could be understood easily by
lay persons.
A laboratory's level of reproducibility can increase or decrease over time.
Reproducibility should be measured not only when a laboratory first implements
DNA typing, but continually on the basis of actual casework, as well as external
proficiency testing (see Chapter 4). One easy way is to record the fragment
measurements from the control samples of known DNA included on the
membrane and regularly examine the variability in these measurements. A
drawback of that approach is that the control pattern might become too well
known to the examiners. A slight variation would eliminate the problem.
Examiners would continue to use a fixed known control sample on every
membrane, but would also be given a blind control sample as a bloodstain to
analyze with each case. The latter sample would be randomly selected from a
collection of a few dozen known samples. The examiners would not know its
specific identity, but only a code number. They would compare the blind control
sample against the known patterns, to determine whether it matched to the
expected extent. Such an internal test of reproducibility would provide continuing
internal measurement of a
Retention of Sample
Scientifically, the best way to resolve ambiguity is often to repeat the
experiment. The U.S. justice system guarantees opposing sides the right to have
repeat experiments performed by experts of their choice, whenever that is
possible. Accordingly, testing laboratories should measure DNA samples before
analysis (with accurate devices, such as fluorometers, as well as with ethidium-
stained "yield" gels) and should use only the quantity of DNA required for
reliable Southern blot analysis. When they can, they should retain enough of a
sample to repeat the entire analysis.
Amplification Conditions
The quality and specificity of amplification with PCR depends on the
amplification conditions: the amplification cycling program (temperatures,
mixture (e.g., primer, nucleotide, polymerase, and magnesium concentrations),
times, and number of cycles), the composition of the amplification and the
amount and nature of the target DNA in the sample (single-stranded or
ensure that differential amplification does not occur should be defined and
documented.
Quantitative analysis of mixed samples with PCR might be problematic.
Suppose that PCR amplification reveals four alleles in a sample, and alleles 1 and 2
give a stronger signal than alleles 3 and 4. A conclusion that the two stronger
alleles correspond to one contributor with genotype 1/2 and the two weaker
alleles to a contributor with genotype 3/4 would be justified only if one had
demonstrated that the amplification and detection process yielded signals that
were directly proportional to the initial quantities of the alleles. If the locus were
subject to differential amplification, the conclusion might be unjustified. This
underscores the importance of characterizing possible differential amplification.
Ideally, primer pairs should amplify only the desired target locus. However,
nonspecific amplification can be seen, if one amplifies for extended cycle
numbers. Limits on the cycle number might be required as a safeguard against
nonspecific products.
Amplification Inhibition
Some forensic samples contain factors that inhibit amplification, either by
binding to the target DNA or by inhibiting the polymerase. In particular,
amplification inhibition is often seen with DNA from older bloodstains. It can
usually be remedied by re-extracting the DNA to remove the inhibiting factor, by
diluting the offending DNA, or by increasing the concentration of polymerase.
There is no evidence that any of those procedures affects typing adversely.
Nevertheless, the nature of inhibiting factors and the mechanism of the inhibition
effect deserve additional study. Each PCR system should be thoroughly
characterized on a range of simulated and known forensic samples, to document
any effect on reliability.
Contamination
One of the most serious concerns regarding PCR-based typing is
contamination of evidence samples with other human DNA. PCR is not
discriminating as to the source of the DNA it amplifies, and it can be exceedingly
sensitive. Potentially, amplification of contaminant DNA could lead to spurious
typing results. Three sorts of contamination can be identified, as set forth below;
each has its own solutions.
sperm DNA and the DNA of vaginal epithelial cells separately. That
allows the genetic contribution of the male and female to be
distinguished. However, there is one important caveat: if the sperm
fraction shows a genotype that matches that of the victim, one cannot
conclude that this represents the genotype of the perpetrator, inasmuch
as it could be due to residual vaginal epithelial cells. The problem should
disappear as PCR-based assays for more loci become available. For
other mixtures, such separation is not possible. For example, it is not
possible to separate the DNA contributed by different persons in mixed
bloodstains or in sexual-assault samples that involve two or more
perpetrators. Mixed samples are a reality of the forensic world that must
be accommodated in interpretation and reconstruction. As a rule, mixed
samples must be interpreted with great caution. Their interpretation
should always be based on results from multiple PCR assays, so that one
can check for consistency across various loci. Interpretations based on
quantity can be particularly problematic—e.g., if one saw two alleles of
strong intensity and two of weak intensity, it would be improper to
assign the first pair to one contributor and the second pair to a second
contributor, unless it had been firmly established that the system was
quantitatively faithful under the conditions used.
• Contamination from handling in the field and laboratory. It is
conceivable that DNA can be transferred to evidence samples or reaction
solutions through handling, either from the person doing the handling or
in transfer from other evidence samples. There are no hard data on the
amounts of DNA transferred by physical contact, but there are anecdotal
reports of experimenters who contaminated their PCR mixtures with
their own DNA. It is difficult to assess the likelihood of this sort of
contamination. Steps should be taken to minimize it, such as handling
samples with gloves and preparing solutions and processing samples in
separate areas. Contamination of solutions can be recognized with
appropriate positive-control and blank-control amplifications, which
should be used routinely. When a stain composed of blood, semen, or
other biological material is analyzed with PCR, it is important to analyze
unstained materials next to the stain with PCR as a control for
contamination.
• PCR product carryover contamination. The most serious problem is
contamination of evidence samples and reaction solutions with PCR
products from prior amplifications. Such products can contain a target
sequence at a concentration a million times greater, and even a relatively
small quantity could swamp the correct signal from the evidence
sample. Even the simple act of flipping the top of a plastic tube might
aerosolize enough DNA to pose a problem.
Many research and diagnostic laboratories have been afflicted with the
problem of PCR carryover. Contamination risks can be minimized by strict
adherence to sterile technique; the use of separate work areas for sample
processing, solution preparation, amplification, and type testing; the use of
separate pipettes in each area (pipettes are a major source of carryover
contamination); and maintenance of a one-way flow of materials from the
evidence-storage area to the sample-preparation area to the type-testing area.
Those precautions focus primarily on preventing PCR carryover
contamination. But it has become clear that carryover products from the PCR
reaction to another must also be eliminated. One way is to use the nucleotide
dUTP in place of dTTP in all PCR reactions.10 PCR products made in this
manner can be selectively destroyed by the enzyme uracil N-glycolase (UNG),
which excises dUTP. Accordingly, all evidence samples would be treated before
PCR amplification with UNG, to destroy contamination from previous PCR
reactions. The method holds promise, although it has not yet been extensively
tested in practice. Methods of detecting and preventing contamination from one
PCR reaction to another in forensic laboratories are generally still in their early
stages, and additional development should be encouraged.
As with contamination due to handling, carryover contamination can be
signaled by the appearance of product in blank controls and of mixed or
inappropriate types in samples and positive controls. Such controls should be
used rigorously. Moreover, it should be remembered that the controls are useful
for monitoring general contamination in the laboratory, not the accuracy of a
particular experiment. If a blank control is positive in one experiment, it indicates a
potential problem not just for that experiment, but for any experiment performed
at about the same time—even in a laboratory contaminated with PCR carryover,
blank controls do not necessarily become contaminated on every occasion. It will
be wise to repeat all work with samples that have never been exposed to the
PCR-typing laboratory.
In view of the problem of contamination due to handling and carryover,
laboratories must incorporate contamination control into their standard operating
procedures. And outbreaks of contamination and the steps taken to correct the
problem should be documented.
One of the best safeguards against contamination is to have DNA typing
independently performed in two laboratories, each starting with a piece of the
unprocessed evidence sample. Given the inexpensiveness of typing, serious
consideration should be given to independent replication of results—at least
during the early stages of this technology.
Use of Kits
One commercial kit for forensic PCR analysis has been marketed. Other
such kits will probably be ready for commercial markets soon. The committee
sees a potential for introduction of unreliable kits and the misuse of kits. The
existence of a kit suggests ease of use and low chance of technical error. The
committee believes that nonexpert laboratories will run a significant chance of
error in using kits. We therefore recommend that a standing committee (discussed
later in this chapter) consider the issue of regulatory approval of kits for
commercial use in forensic DNA analysis. Even though no precedent exists for
regulation of tests in forensic applications, we believe that it might be necessary
for a government agency to test and approve kits for DNA analysis before their
actual forensic use.
loci. It generates a large quantity of relatively pure product that can be analyzed
with much greater precision than Southern blots, even down to the nucleotide
level. At the same time, it poses even more serious issues of proficiency, control,
and technology transfer than RFLP typing.
In summary, it is well established that one can greatly amplify a locus with
authenticity and that one can reliably detect alleles or sequence variation at the
amplified locus with any of a number of techniques. PCR analysis is extremely
powerful in medical technology, but it has not yet achieved full acceptance in the
forensic setting. The theory of PCR analysis, even though it is the analysis of
synthetic DNA, as opposed to the natural sample, is scientifically accepted and
has been accepted by a number of courts. However, most forensic laboratories
have invested their energy in development of RFLP technology and have left the
development of forensic PCR technology to a few other laboratories. Thus, there
is no broad base of experience in the use of the technique in identity testing.
Forensic PCR-based testing is now limited for the most part to analysis of
genetic variation at the DQ locus in the HLA complex. Potential ambiguities in
typing results cannot yet be checked by studying a number of other loci in the
same DNA sample. That shortcoming will be rectified with the addition of new
PCR markers for forensic analysis. However, it is clear that analysis of the DQ
locus with PCR can often provide useful information during the investigative
phase in the forensic setting.
In general, further experience should be gained with respect to PCR in
identity testing. Information on the extent of the contamination problem in PCR
analysis and the differential amplification of mixed samples needs to be further
developed and published. A great deal of this information can be obtained when a
number of polymorphic systems are available for PCR analysis. Ambiguous
results obtained with a number of polymorphic markers will signal contamination
or mixtures of DNA in a sample.
Quantification of PCR results needs to be explored, to make the results more
reliable. Laboratories that gain experience with PCR should determine the
relationship between cycle number and percentage of contaminating DNA easily
detected for each system used. Control primers that amplify small amounts of
DNA reliably and robustly need to be added to test amplifications. In general,
information derived from new polymorphic loci under standardized conditions
with easily quantifiable results or end points is needed. Considerable advances in
the use of PCR in forensic analysis can be expected soon; the method has
enormous promise.
used in some court cases, and other methods are being developed in scientific and
commercial research laboratories. Typing methods will continue to be replaced
with ever more sophisticated approaches for some time to come. These
developments hold great promise for increasing the sensitivity and reliability of
forensic DNA typing.
The rapidity of development creates a need to balance two competing
societal objectives. On the one hand, new technologies should be made available
quickly. On the other hand, forensic typing methods should not be used until their
soundness is established both in principle and in practice. The problem involves
both technology and technology transfer. Forensic DNA typing is drawing
methods from the cutting edge of molecular genetics, but must apply them to
quite different circumstances.
The committee believes that the field of forensic science would be best
served by the creation of a National Committee on Forensic DNA Typing
(NCFDT) to provide advice on scientific and technical issues as they arise.
NCFDT would consist primarily of molecular geneticists, population geneticists,
forensic scientists, and additional members knowledgeable in law and ethics. Its
charges would be to provide guidance about the power and limitations of DNA
typing methods, to identify potential problems and their solutions, to provide
guidance about whether new technologies are ready for practical use in the
forensic laboratory, and to provide advice concerning the regulation of kits for
forensic DNA typing. In addition (as discussed in Chapter 3), NCFDT would
provide advice on population genetics and statistical interpretation.
Such a committee could play a critical role in smoothing the acceptance of
DNA typing technologies in the courtroom while ensuring their reliability.
Although NCFDT would have no formal regulatory authority, we anticipate that
substantial influence would derive from its stature and the quality of its advice, so
that courts could look to its recommendations in making their decisions.
The present committee recommends that NCFDT be convened under the
auspices of an appropriate government agency. Because its task is fundamentally
scientific, we feel that the agency should be one whose primary mission is
scientific, rather than related to law enforcement. To avoid any appearance of
conflict of interest, an agency that uses forensic DNA typing itself would be
unsuitable. Two excellent choices would be the National Institutes of Health
(NIH) or the National Institute of Standards and Technology (NIST). NIH has
extensive experience in molecular biology, population genetics, and laboratory
practice. NIST has less direct experience in those fields, but has considerable
experience in evaluating technologies. Regardless of which agency convenes
NCFDT, we believe that the effort should have broad government support from
NIST, NIH, the National Science Foundation, the National Institute of Justice, the
Federal
Bureau of Investigation, and the State Justice Institute. NCFDT should also have
broad support from the American Society of Crime Laboratory Directors, the
Genetics Society of America, and the American Society of Human Genetics.
The creation of an expert advisory committee is a somewhat unusual step
for forensic science. However, we feel that it is the appropriate way to ensure
that the field can incorporate new developments promptly while maintaining high
standards.
SUMMARY OF RECOMMENDATIONS
which it will perform reliably and for strict contamination measures and
other control procedures
• The committee strongly recommends the establishment of a National
Committee on Forensic DNA Typing under the auspices of an
appropriate government agency, such as NIH or NIST, to provide expert
advice primarily on scientific and technical issues concerning forensic
DNA typing.
REFERENCES
1. Balazs I, Baird M, Clyne M, Meade E. Human population genetic studies of five hypervariable
DNA loci, Am J Hum Genet. 44:182-190, 1989.
2. Culliford BJ. Determination of phosphoglucomutase types in bloodstains. J Forensic Sci Sociol.
7:131-133, 1967.
3. Culliford BJ. The examination and typing of blood stains in the crime laboratory. Washington,
D.C.: U.S. Government Printing Office, 1971.
4. Sensabaugh GF. Biochemical markers of individuality, pp. 338-415 in: Saferstein R, ed. Forensic
science handbook. Englewood Cliffs, New Jersey: Prentice-Hall, 1982.
McNally L, Baird M, McElfresh K, Eisenberg A, Balazs I. increased migration rate observed in DNA
from evidentiary material precludes the use of sample mixing to resolve forensic cases of
identity. Appl Theor Electrophoresis. 5:267-272, 1990.
6. Thompson WC, Ford S. The meaning of a match: sources of ambiguity in the interpretation of a
DNA print in forensic DNA technology, pp. 93152 in: Farley M, Harrington J, eds. Forensic
DNA technology. Chelsea, Michigan: Lewis Publishing, 1991.
7. Amheim N, Levenson CH. Polymerase chain reaction. C&E News68:36-47, October 1, 1991).
8. Erlich HA, Gelfand D, Sninsky J J. Recent advances in the polymerase chain reaction. Science.
252:1643-1651, 1991.
9. Comey CT, Jung JM, Budowle B. Use of formamide to improve amplification of HLA DQa
sequences. Biotechniqucs. 10:60-61, 1991.
10. Longo MC, Berninger MS, Harley JL. Use of uracil DNA glycosylase to control carryover
contamination in polymerase chain reactions. Gene.93:125, 1990.
11. Saiki RK, Walsh PS, Levenson CH, Erlich HA. Genetic analysis of amplified DNA with
immobilized sequence-specific oligonucleotide probes. Proc Natl Acad Sci USA.
86:6230-6234, 1989.
3
DNA Typing: Statistical Basis for
Interpretation
Can DNA typing uniquely identify the source of a sample? Because any two
human genomes differ at about 3 million sites, no two persons (barring identical
twins) have the same DNA sequence. Unique identification with DNA typing is
therefore possible provided that enough sites of variation are examined.
However, the DNA typing systems used today examine only a few sites of
variation and have only limited resolution for measuring the variability at each
site. There is a chance that two persons might have DNA patterns (i.e., genetic
types) that match at the small number of sites examined. Nonetheless, even with
today's technology, which uses 3-5 loci, a match between two DNA patterns can
be considered strong evidence that the two samples came from the same source.
Interpreting a DNA typing analysis requires a valid scientific method for
estimating the probability that a random person might by chance have matched
the forensic sample at the sites of DNA variation examined. A judge or jury could
appropriately weigh the significance of a DNA match between a defendant and a
forensic sample if told, for example, that ''the pattern in the forensic sample
occurs with a probability that is not known exactly, but is less than 1 in 1,000" (if
the database that shows no match with the defendant's pattern is of size 1,000).
To say that two patterns match, without providing any scientifically valid
estimate (or, at least, an upper bound) of the frequency with which such matches
might occur by chance, is meaningless.
Substantial controversy has arisen concerning the methods for estimating
To estimate the frequency of a particular DNA pattern, one might count the
number of occurrences of the pattern in an appropriate random population
sample. If the pattern occurred in 1 of 100 samples, the estimated frequency
would be 1%, with an upper confidence limit of 4.7%. If the pattern occurred in 0
of 100 samples, the estimated frequency would be 0%, with an upper confidence
limit of 3%. (The upper bound cited is the traditional 95% confidence limit,
whose use implies that the true value has only a 5% chance of exceeding the
upper bound.) Such estimates produced by straightforward counting have the
virtue that they do not depend on theoretical assumptions, but simply on the
sample's having been randomly drawn from the appropriate population.
However, such estimates do not take advantage of the full potential of the genetic
approach.
Unlike many of the technical aspects of DNA typing that are validated by
daily use in hundreds of laboratories, the extraordinary population-frequency
estimates sometimes reported for DNA typing do not arise in research or medical
applications that would provide useful validation of the frequency of any
particular person's DNA profile. Because it is impossible or impractical to draw a
large enough population to test calculated frequencies for any particular DNA
profile much below 1 in 1,000, there is not a sufficient body of empirical data on
which to base a claim that such frequency calculations are reliable or valid per
se. The assumption of independence must be strictly scrutinized and estimation
procedures appropriately adjusted if possible. (The rarity of all the genotypes
represented in the databank can be demonstrated by pairwise comparisons. Thus,
in a recently reported analysis of the FBI database, no exactly matching pairs of
profiles were found in five-locus DNA profiles, and the closest match was a
single three-locus match among 7.6 million basepair comparisons.)13
The multiplication rule has been routinely applied to blood-group
frequencies in the forensic setting. However, that situation is substantially
different: Because conventional genetic markers are only modestly polymorphic
(with the exception of human leukocyte antigen, HLA, which usually cannot be
typed in forensic specimens), the multilocus genotype frequencies are often about
1 in 100. Such estimates have been tested by simple empirical counting. Pairwise
comparisons of allele frequencies have not revealed any correlation across loci.
Hence, the multiplication rule does not appear to lead to the risk of extrapolating
beyond the available data for conventional markers. In contrast, highly
polymorphic DNA markers exceed the informative power of protein markers, so
multiplication leads to estimates that are less than the reciprocal of the size of the
databases.
• Count the frequency of alleles. For each allele in the genotype, examine a
random sample of the population and count the proportion of matching
alleles—that is, alleles that would be declared to match according to the
rule that is used for declaring matches in a forensic context. This step
requires only the selection of a sample that is truly random with
reference to the genetic type; it does not appeal to any theoretical
models.
c1. If a random sample of the appropriate population shows that the frequencies
of a1, a2, b1, b2, and c1 are approximately 0.1, 0.2, 0.3, 0.1, and 0.2,
respectively, then the population frequency of the genotype would be estimated to
be [2(0.1)(0.2)][2(0.3)(0.1)][(0.2)(0.2)] = 0.000096, or about 1 in 10,417.
studies on a large mixed population, such as a racial group, and use statistical
tests to detect the presence of substructure; (2) derive theoretical principles that
place bounds on the possible degree of population substructure; and (3) directly
sample different groups and compare the observed allele frequencies. The third
offers the soundest foundation for assessing population substructure, both for
existing loci and for many new types of polymorphisms under development.
In principle, population substructure can be studied with statistical tests to
examine deviations from Hardy-Weinberg equilibrium and linkage equilibrium.
Such tests are not very useful in practice, however, because their statistical power
is extremely low: even large and significant differences between subgroups will
produce only slight deviations from Hardy-Weinberg expectations. Thus, the
absence of such deviations does not provide powerful evidence of the absence of
substructure (although the presence of such deviations provides strong evidence
of substructure).
The correct way to detect genetic differentiation among subgroups is to
sample the subgroups directly and to compare the frequencies. The following
example is extreme and has not been observed in any U.S. population, but it
illustrates the difference in power. Suppose that a population consists of two
groups with different allele frequencies at a diallelic locus:
A a
Group I 0.5 0.5
Group II 0.9 0.1
If there is random mating within the groups, Hardy-Weinberg equilibrium
within the groups will produce these genotype frequencies:
AA Aa aa
Group I 0.25 0.50 0.25
Group II 0.81 0.18 0.01
Suppose that Group I is 90% of the population and Group II is 10%. In the
overall population, the observed genotype frequencies will be
AA = (0.9)(0.25) + (0.1)(0.81) = 0.306
Aa = (0.9)(0.50) + (0.1)(0.18) = 0.468
aa = (0.9)(0.25) + (0.1)(0.01) = 0.226
If we were unaware of the population substructure, what would we expect
under Hardy-Weinberg equilibrium? The average allele frequencies will be
A = (0.9)(0.5) + (0.1)(0.9) = ).54
a = (0.9)(0.5) + (0.1)(0.1) = 0.46
which would correspond to the Hardy-Weinberg proportions of
AA = (0.54)(0.54) = 0.2916
Aa = 2(0.54)(0.46) = 0.4968
aa + (0.46)(0.46) = 0.2116
Even though there is substantial population substructure, the proportions do
not differ greatly from Hardy-Weinberg expectation. In fact, one can show that
detecting the population differentiation with the Hardy-Weinberg test would
require a sample of nearly 1,200, whereas detecting it by direct examination of
the subgroups would require a sample of only 22. In other words, the Hardy-
Weinberg test is very weak for testing substructure.
The lack of statistical power to detect population substructure makes it
difficult to detect genetic differentiation in a heterogeneous population. Direct
sampling of subgroups is required, rather than examining samples from a large
mixed population.
Similarly, population substructure cannot be predicted with certainty from
theoretical considerations. Studies of population substructure for protein
polymorphisms cannot be used to draw quantitative inferences concerning
population substructure for VNTRs, because loci are expected to show different
degrees of population differentiation that depend on such factors as mutation rate
and selective advantage. Differences between races cannot be used to provide a
meaningful upper bound on the variation within races. Contrary to common
belief based on difference in skin color and hair form, studies have shown that the
genetic diversity between subgroups within races is greater than the genetic
variation between races.15 Broadly, the results of the studies accord with the
theory of genetic drift: the average allele frequency of a large population group
(e.g., a racial group) is expected to drift more slowly than the allele frequencies
of the smaller subpopulations that it comprises (e.g., ethnic subgroups).
In summary, population differentiation must be assessed through direct
studies of allele frequencies in ethnic groups. Relatively few such studies have
been published so far, but some are under way.16 Clearly, additional such studies
are desirable.
utable to any single locus would be 1/400 (1/20 × 1/20). In any case, it seems
reasonable not to attach much greater weight to any single locus.
Some legal commentators have pointed out that frequencies should properly
be based on the population of possible perpetrators, rather than on the population
to which a particular suspect belongs.17,18Although that argument is formally
correct, practicalities often preclude use of that approach. Furthermore, the ceiling
principle eliminates the need for investigating the perpetrator population, because
it yields an upper bound to the frequency that would be obtained by that
approach.
Some have proposed a Bayesian approach,19,20,21 to the presentation of DNA
evidence. However, this approach, focusing on likelihood ratios, does not avoid
the kinds of population genetic problems discussed in this chapter. The
committee has not tried to assess the relative merits of Bayesian and frequentist
approaches, because, outside the field of paternity testing, no forensic laboratory
in this country has, to our knowledge, used Bayesian methods to interpret the
implications of DNA matches in criminal cases.
the scene of a crime, its DNA pattern can be determined and compared
with a databank. If the unidentified sample perfectly matches a sample in
the convicted-criminal databank at enough loci, the probable perpetrator
is likely to have been found. However, a different outcome could occur:
the sample might match no entry perfectly, but match some entry at
about one allele per locus. Depending on the number of loci studied, one
could have a compelling case that the source of the sample was a first-
degree relative (e.g., brother) of the convicted criminal whose entry was
partially matched. (In practice, four loci would not suffice for this
conclusion, but 10 might.) Such information could be sufficient to focus
police attention on a few persons and might be enough to persuade a
court to compel a blood sample that could be tested for exact match with
the sample.
To put it succinctly, DNA databanks have the ability to point not just to
individuals but to entire families—including relatives who have committed no
crime. Clearly, this poses serious issues of privacy and fairness. As we discuss
more fully later (Chapter 5), it is inappropriate, for reasons of privacy, to search
databanks of DNA from convicted criminals in such a fashion. Such uses should
be prevented both by limitations on the software for search and by statutory
guarantees of privacy.
1. First, the testing laboratory should check to see that the observed
multilocus genotype matches any sample in its population database.
Assuming that it does not, it should report that the DNA pattern was
compared to a database of N individuals from the population and no
match was observed, indicating its rarity in the population. This
simple statement based on the counting principle is readily
understood by jurors and makes clear the size of the database being
examined.
2. The testing laboratory should then calculate an estimated population
frequency on the basis of a conservative modification of the ceiling
principle, provided that population studies have been carried out in
at least three major "races" (e.g., Caucasians, blacks, Hispanics,
Asians, and Native Americans) and that statistical evaluation of
Hardy-Weinberg equilibrium and linkage disequilibrium has been
carried out (with methods that accurately incorporate the empirically
determined reproducibility of band measurement) and no significant
deviations were seen. The conservative calculation represents a
reasonable effort to capture the actual power of DNA typing while
reflecting the fact that the recommended population studies have not
yet been undertaken. The calculation should be carried out as
follows.
a The upper 95% confidence limit is given by the formula p + 1.96 , where p is the
observed frequency and N is the number of chromosomes studied.
claimed to be proprietary. If scientific evidence is not yet ready for both scientific
scrutiny and public re-evaluation by others, it is not yet ready for court.
SUMMARY OF RECOMMENDATIONS
Although mindful of the controversy concerning the population genetics of
DNA markers, the committee has decided to assume that population substructure
might exist for currently used DNA markers or for DNA markers that will be
used in the future. The committee has sought to develop a recommendation on the
statistical interpretation of DNA typing that is appropriately conservative, but at
the same time takes advantage of the extraordinary power of individual
identification provided by DNA typing. We have sought to develop a
recommendation that is sufficiently robust, but is flexible enough to apply not
only to markers now used, but also to markers that might be technically
preferable in the future. We point out that in using conservative numbers in the
interpretation of DNA typing results, any loss of statistical power is often offset
through typing of additional loci. The committee seeks to eliminate the necessity
to consider the ethnic background of a subject or of the group of potential
perpetrators.
al funds, in view of the benefits for law enforcement in general and for
the convicted-offender databanks in particular.
• The ceiling principle should be used in applying the multiplication rule
for estimating the frequency of particular DNA profiles. For each allele
in a person's DNA pattern, the highest allele frequency found in any of
the 15-20 populations or 5% (whichever is larger) should be used.
• In the interval (which should be short) while the reference samples are
being collected, the significance of the findings of multilocus DNA
typing should be presented in two ways: 1) If no match is found with any
sample in a total databank of N persons (as will usually be the case),
that should-be stated, thus indicating the rarity of a random match. 2) In
applying the multiplication rule, the 95% upper confidence limit of the
frequency of each allele should be calculated for separate U.S. "racial"
groups and the highest of these values or 10% (whichever is the larger)
should be used. Data on at least three major "races" (e.g., Caucasians,
blacks, Hispanics, Asians, and Native Americans) should be analyzed.
• Any population databank used to support DNA typing should be openly
available for scientific inspection by parties to a legal case and by the
scientific community.
• Laboratory error rates should be measured with appropriate proficiency
tests and should play a role in the interpretation of results of forensic
DNA typing.
REFERENCES
1. Devlin B, Risch N, Roeder K. No excess of homozygosity at loci used for DNA fingerprinting,
Science. 249:1416-1420, 1990.
2. Cohen JE, Lynch M, Taylor CE, Green P, Lander ES. Forensic DNA tests and Hardy-Weinberg
equilibrium. (Comment on Devlin et al. Science. 249:1416-1420, 1990.) Science.
253:1037-1039, 1991.
3. Devlin B, Risch N, Roeder K. (Response to Cohen et al. Science. 253:1(137-1039, 1991). Science.
253:1039-1041, 1991.
4. Lander ES. Research on DNA typing catching up with courtroom application. (Invited Editorial.)
Am J Hum Genet. 48:819-823, 1991.
5. Wooley JR. A response to Lander: The courtroom perspective. Am J Hum Genet. 49:892-893,
1991.
6. Caskey CT. Comments on DNA-based forensic analysis. (Response to Lander. Am J Hum Genet.
48:819, 1991.) Am J Hum Genet. 49:893-905, 1991,
7. Chakraborty R. Statistical interpretation of DNA-typing data. (Letter.) Am .I Hum Genet.
49:895-897, 1991.
8. Daiger SP. DNA fingerprinting. (Letter.) Am J Hum Genet. 49:897, 1991.
9. Lander ES. Lander reply. (Letter.) Am J Hum Genet. 49:899-903, 1991.
10. Lewontin RC, Hartl DL. Population genetics in forensic DNA typing. Science. 254:1745-1750,
1991.
11. Chakraborty R, Daiger SP. Polymorphisms at VNTR loci suggest homogeneity of the white
population of Utah. Hum Biol. 63:571-588, 1991.
12. Chakraborty R, Kidd K. The utility of DNA typing in forensic work. Science. 254:17351739,
1991.
13. Risch N, Devlin B. On the probability of matching DNA fingerprints. Science. 255:717720, 1992.
14. Weir B. Independence of VNTR alleles defined as fixed bins. Genetics, in press.
15. Lewontin RC. The apportionment o£ human diversity. Evol Biol. 6:381-398, 1972.
16. Deka R, Chakraborty R, Ferrell RE. A population genetic study of six VNTR loci in three
ethnically defined populations. Genomics. 11:83-92, 1991.
17. Cavalli-Sforza LL, Bodmer WF. The genetics of human populations. San Francisco: W.H.
Freeman, 1971.
18. Lempert R. Some caveats concerning DNA as criminal identification evidence: with thanks to the
Reverend Bayes. Cardozo Law Rev. 13:303-341, 1991.
19. Evett I, Werrett D, Pinchin R, Gill P. Bayesian analysis of single locus DNA profiles. Proceedings
of the International Symposium on Human Identification 1989. Madison, Wisconsin:
Promega Corp., 1991).
20 Berry DA. Influences using DNA profiling in forensic identification and paternity cases. Star Sci.
6:175-205, 1991.
21. Berry DA, Evett IW, Pinchin R. Statistical inferences in crime investigation using DNA profiling. J
Royal Star Soc. [Series C - Applied Statistics], in press.
22. Budowle B, Giusti AM, Waye JS, Baechtel FS, Fourney RM, Adams DE, Presley LA, Deadman
HA, Monson KL. Fixed-bin analysis for statistical evaluation of continuous distributions of
allelic data from VNTR loci, for use in forensic comparisons. Am J Hum Goner.
48-841-855, 1991.
23. DiLonardo AM, Darlu P, Baur M, Orrego C, King MC. Human genetics and human rights:
identifying the families of kidnapped children. Am J Forensic Med Pathol.5:339-347, 1984.
24. King MC. An application of DNA sequencing to a human rights problem. Pp. 117-132in:
Friedmann T, ed. Molecular Genetic Medicine. Vol. 1. New York: Academic Press, 1991.
25. California Association of Crime Laboratory Directors, DNA Committee, Reports to the Board of
Directors: 1, August 25, 1987; 2, November 19, 1987; 3, March 28, 1988, 4, May 18, 1988;
5, October 1, 1988; 6, October 1, 1988.
26. Lander ES. DNA fingerprinting on trial. (Commentary.) Nature. 339:501-505, 1989.
27. The fallibility of forensic DNA testing: of proficiency in public and private laboratories. Part I.
The private sphere. Sci Sleuth Rev. 14121:10, 1990.
28. Dausset J, Cann H, Cohen D, Lathrop M, Lalouel J-M, White R, Centre d'Etude du
Polymorphisme Humain (CEPH): collaborative genetic mapping of the human genome.
Genomics. 6:575-577, 1990.
29. Jeffreys A, MacLeod A, Tamaki K, Nell D, Monckton D. Minisatellite repeat coding as a digital
approach to DNA typing. Nature. 354:204-209, 1991.
4
Ensuring High Standards
Critics and supporters of the forensic use of DNA typing agree that there has
been a lack of standardization of practices and uniformly accepted methods for
quality assurance. The lack is due largely to the rapid emergence of DNA typing
and its introduction in the United States through the private sector.
As the technology developed in the United States, private laboratories using
widely differing methods (single-locus RFLP, multilocus RFLP, and PCR) began
to offer their services to law-enforcement agencies. During the same period, the
FBI was developing its own RFLP method, with yet another restriction enzyme
and different single-locus probes. Its method has become the one most widely
(albeit not exclusively) used in public forensic-science laboratories, as a result of
the FBI's national offering of free extensive training programs to forensic
scientists. Each method has its own advantages and disadvantages, databanks,
molecular-weight markers, match criteria, and reporting methods. In some
courts, there have been differences of opinion as to the reliability, acceptability,
and applicability of the various methods and particularly the degree of their
specificity or discriminating power.
Regardless of the causes, practices in DNA typing vary and so do the
educational backgrounds, training, and experience of the scientists and
technicians who perform these tests, the internal and external proficiency testing
conducted, the interpretation of results, and approaches to quality assurance.
It is not uncommon for an emerging technology to go without regulation
until its importance and applicability are established. Indeed, the de
Certification of Individuals
In the certification approach, a certifying body dictates that fulfillment of
specified education, training, and experiential requirements be demonstrated by
documentation and examination. An examination can also include a laboratory
practicum. Individual certification has many advantages, but it is not adequate. A
person does not perform DNA typing tests in isolation. A person's ability to
produce high-quality results consistently depends heavily on the procedures,
reagents, equipment, management, and attitudes in the work environment. It is
impossible to separate a person from his or her organization and physical setting.
In addition, personal certification can be expensive and usually requires funding
support from the employing institution.
We recommend that the National Institute of Justice, given its interest in
training, develop training programs in association with the American Society of
Crime Laboratory Directors. Such a cooperative effort would allow continuity of
candidate selection, training, examination, and certification.
Laboratory Accreditation
Accreditation is a more comprehensive approach to regulation. It requires
that a laboratory demonstrate that its management, operations, individual
personnel, procedures and equipment, physical plant and security, and personnel
safety procedures all meet standards. Laboratory accreditation programs can be
voluntary or mandatory. Although voluntary programs can have a positive effect,
they suffer from the limitations that laboratories need not comply, that standard-
setting need not be open to public scrutiny, and that accreditation might be
contingent on membership in a professional organization. Accreditation programs
required by federal or state law provide a greater level of assurance.
Licensing of Laboratories
Licensing involves vesting, by the federal government or a state
government, of a regulatory body with the responsibility and authority to
establish a series of requirements that a laboratory must satisfy if it is to be
allowed to operate in a defined jurisdiction or to present evidence in its courts.
The licensing approach does not suffer the disadvantage of being voluntary. State
or federal laws can place sanctions on a laboratory that is not licensed by the
specified body. A potential drawback is that the development of such a program
can be time-consuming and expensive. In addition, licensing can be
anticompetitive and can discourage innovation. In the
case of state (as opposed to federal) regulation, there is the potential problem that
different practices will be mandated in different jurisdictions, although this can be
avoided if a few states take the lead and others follow suit by adopting similar
regulations or recognize other states' licensing. The most efficient path seems to
be for a federal licensing procedure to standardize the process for the nation and
avoid state licensing, which can be restrictive and redundant. The federal license
would be most cost-efficient and provide the field with a mechanism of quality
assurance.
the membership requirements for ASCLD. There has, however, been no listing of
laboratories that failed examination, had deficiencies, or were advised to
discontinue providing services, and that constitutes a flaw in voluntary laboratory
accreditation as carried out by ASCLD-LAB.
The New York state legislature has developed legislation to create a
licensing program that was recommended by the Governor's Select Commission
on DNA Typing, which was chaired by the state commissioner of criminal justice
and consisted of lawyers, molecular biologists, forensic scientists, and population
geneticists. It proposes that New York administer a licensing program with the
advice of a scientific review board and a DNA advisory committee. Those bodies
would consist of independent scientists in molecular biology and population
genetics with no ties to forensic laboratories, through financial relationships,
extensive collaboration, or provision of extensive testimony. This pioneering
effort is consistent with the leadership role that New York has assumed in other
laboratory testing—it is the national leader in the regulation of cytogenetics, for
example. Both houses of the New York legislature passed a bill establishing a
regulatory framework but it later was withdrawn. A similar bill has been
reintroduced in the current legislative session.
A third approach is reflected in legislation recently introduced in the U.S.
Congress "to provide financial assistance to state and local governments wishing
to upgrade their crime laboratories with DNA genetic testing capability"
contingent on their adherence to particular standards. The proposed law, entitled
the DNA Proficiency Testing Act (H.R. 339, introduced by Rep. Frank Horton),5
would promote increased quality assurance by providing grants from the
Department of Justice (DOJ) Bureau of Justice Assistance to state and local
forensic laboratories and mandating the FBI to publish "advisory standards for
testing the proficiency of forensic laboratories" and to carry out certification for
DNA proficiency-testing programs. The Bureau of Justice Assistance would
make grants to state or local forensic laboratories. Within 6 months of enactment,
the FBI would publish standards for testing the proficiency of laboratories that
conduct DNA typing. The standards would "specify criteria to be applied to each
procedure used by forensic laboratories to conduct analyses of DNA." The FBI
would revise the standards from time to time, as necessary.
Under H.R. 339, the FBI could approve a DNA proficiency-testing program
offered by a private organization, if the program satisfied the proficiency-testing
standards and the testing laboratory were prepared to issue to each forensic
laboratory that participated in the program a document that certified the
participation and specified the period for which the proficiency test applied to the
forensic laboratory. The FBI would have to withdraw approval for a DNA
proficiency-testing program if the program or testing laboratory failed to satisfy
the proficiency-testing standards.
SUMMARY OF RECOMMENDATIONS
REFERENCES
1. Technical Working Group on DNA Analysis Methods (TWGDAM). Guidelines for a quality
assurance program for DNA restriction fragment length polymorphism analysis . Crime Lab
Dig. 16(2):40-59, 1989.
2. Technical Working Group on DNA Analysis Methods (TWGDAM). Guidelines for a proficiency
testing program for DNA restriction fragment length polymorphism analysis. Crime Lab
Dig. 17(3):50-60, 1990.
3. Technical Working Group on DNA Analysis Methods (TWGDAM). Statement of the Working
Group on Statistical Standards for DNA Analysis. Crime Lab Dig. 17(3):53-58, 1990.
5
Forensic DNA Databanks and Privacy of
Information
DNA typing in the criminal-justice system has so far been used primarily for
direct comparison of DNA profiles of evidence samples with profiles of samples
from known suspects. However, that application constitutes only the tip of the
iceberg of potential law-enforcement applications. If DNA profiles of samples
from a population were stored in computer databanks (databases), DNA typing
could be applied in crimes without suspects. Investigators could compare DNA
profiles of biological evidence samples with a databank to search for suspects.
In many respects, the situation is analogous to that of latent fingerprints.
Originally, latent fingerprints were used for comparing crime-scene evidence with
known suspects. With the development of the Automated Fingerprint
Identification Systems (AFIS) in the last decade, the investigative use of
fingerprints has dramatically expanded. Forensic scientists can enter an
unidentified latent-fingerprint pattern into the system and within minutes compare
it with millions of people's patterns contained in a computer file. In its short
history, automated fingerprint analysis has been credited with solving tens of
thousands of crimes.1
This chapter examines whether similar databanks of DNA profiles should be
created and, if so, how and when.
• Latent fingerprints are found at crime scenes much more commonly than
are body fluids that contain DNA. Latent-fingerprint analysis can be
useful in a wide range of crimes, including many murders, rapes,
assaults, robberies, and burglaries. However, the probative value of
latent fingerprints is often limited to establishing that a suspect was
present at a location—and that does not automatically imply guilt. DNA
analysis will be useful in more limited settings. DNA analysis will be
useful primarily in rapes (because semen is often recovered) and
murders (those in which either the perpetrator's blood was spilled at the
crime scene or the victim's blood stained the perpetrator's personal
effects—only the former will assist in identifying an unknown suspect).
Where it exists, DNA evidence will often be more probative than
fingerprints, in that the presence of body fluids is harder to attribute to
innocuous causes. That is especially true in rape cases, in which positive
identification of semen in the vagina is virtual proof of intercourse
(although it leaves open the issue of whether it was consensual).
Consequently, the potential utility of a DNA profile databank must be
evaluated in terms of the particular crimes to which it is primarily
suited.
• Fingerprints have a defined physical pattern independent of the method
of visualization, whereas DNA profiles are derived patterns that can be
constructed with various protocols (e.g., different restriction enzymes to
cut the DNA and different probes to examine different loci) that produce
completely different patterns that cannot be readily interconverted. The
advance of DNA technology will see the development of new protocols
that offer technical advantages but produce different and incompatible
patterns.
address, and genetic counselor's name and address is stored with the samples.
That information is useful for local, independent bookkeeping and record
management. But it is also ripe for statistical or correlative disclosure. Just the
existence of a sample from a person in a databank might be prejudicial to the
person, independently of any DNA related information. In some laboratories, the
donor cannot legally prevent outsiders' access to the samples, but can request its
withdrawal. A request for withdrawal might take a month or more to process. In
most cases, only physicians with signed permission of the donor have access to
samples, but typically no safeguards are taken to verify individual requests
independently. That is not to say that the laboratories intend to violate donors'
rights; they are simply offering a service for which there is a recognized market
and attempting to provide services as well as they can. Much has been written on
statistical databank systems and associated security issues.3
Guidelines for release of DNA samples and disclosure of DNA typing
information must be designed to safeguard the rights of persons who, for one
reason or another, get involved in a DNA typing (see Chapter 7 for further
discussion) without burdening law-enforcement agencies and civil investigative
authorities with unnecessarily protective policies.
The need for safeguards of DNA information has not been completely lost
on lawmakers considering databank legislation. Some state legislation has
addressed the issue. For example, the Virginia law4 establishing a DNA profile
databank for convicted offenders states that
Although that is a good start, state laws should state explicitly the types of
uses that can be authorized. In particular, in addition to the points made in the
opinion just quoted, investigation of DNA samples or stored information for the
purpose of obtaining medical information or discerning other traits should be
prohibited, and violations should be punishable by law. Several states incorporate
some of those specific protections into their statutes establishing DNA profile
databanks. However, the committee urges all states to be systematic in defining
authorized uses of information in DNA databanks.
METHODOLOGICAL STANDARDIZATION
Because of the incompatibility between DNA typing methods, federal, state,
and local laboratories that wish to use a national DNA profile databank must all
adopt a single standardized method for analyzing samples—both databank
specimens and evidence specimens. Accordingly, the development of a national
DNA databank has the potential advantage of acting as a driving force for
standardization in forensic DNA typing, but the potential disadvantage of
ossifying a rapidly moving technology.
It is broadly agreed that current RFLP typing methods constitute simply an
initial approach that will be replaced in the next few years by procedures that are
much easier to automate, much less expensive, and more informative. Premature
development of a national databank based on current RFLP typing methods runs
the risk of perpetuating a "dinosaur" technology in the face of better techniques.
The committee believes that it is too early to launch a comprehensive
national DNA profile databank. However, it is appropriate to carry out pilot
programs based on RFLP technology with the FBI and states that have active
DNA typing efforts. The initial efforts should help to define the problems and
issues that will be encountered in the fashioning of a comprehensive program.
Such projects should be explicitly viewed as preliminary, with the clear
expectation that the databank will be supplanted in the next several years by
better methods.
Before even pilot projects can be begun, the degree of interlaboratory
reproducibility—which is essential to the success of a databank—should be
thoroughly documented. So far, there have been only a few interlaboratory-
reproducibility studies to compare the ability of different laboratories to measure
the same DNAs accurately under different circumstances. The National Institute
of Standards and Technology (NIST), in concert with the Federal Bureau of
Investigation (FBI) and the Technical Working Group on
murders and 100,000 forcible-rape cases per year. It is estimated that 30% of all
murder cases and 70% of all rape cases are never closed by arrest (John Hicks,
personal communication, 1990). It should also be pointed out that only an
estimated 50% of rapes are in fact even reported.
The second question is also difficult to answer, but it is clear that crimes of
most types will not afford the opportunity to recover relevant biological evidence
that will allow the police to identify an unknown suspect—i.e., the perpetrator's
own body fluids. They include larcenies, burglaries, and assaults, for which
ordinary fingerprints are frequently found. The major exception is rape, for which
semen samples can be recovered in many cases and might provide prima facie
evidence of sexual intercourse. In a small minority of homicides, blood, hair, or
tissue samples from the perpetrator are left at the scene of the crime (e.g., because
of a fight at the scene).
A DNA profile databank would thus be valuable primarily in investigating
forcible rape, although the databank would be useful for some other
investigations. State legislatures considering setting up such databanks should
weigh the benefits in terms of solved rape cases and the costs in terms of
collecting samples from persons likely to commit rapes (primarily, it seems,
convicted sex offenders). Initial state efforts to develop DNA profile databanks
were indeed aimed at sex offenders. Interestingly, some states rapidly expanded
their programs to include all convicted offenders—without explicit weighing of
the potential benefits of possessing such persons' patterns for solving crimes and
the potential costs.
The above discussion justifies the development of a databank of DNA
profiles of unknown subjects (open cases) and of offenders convicted of violent
sex crimes. Such a databank would provide law enforcement with a powerful tool
in linking sexual-assault cases through DNA profiles and tracking the activities
of serial rapists. In light of recidivism and the continuing increase in reported
rapes in this country, a databank of convicted sex offenders would provide
investigators with a logical first place to look for assistance in solving unknown-
offender sexual-assault cases.
SAMPLE STORAGE
Another difficult issue is the storage and maintenance of DNA samples
themselves (or any reusable products of the typing process), as opposed to DNA
profiles. In principle, retention of DNA samples creates an opportunity for misues
—i.e., for later testing to determine personal information. In general, the
committee discourages the retention of DNA samples.
However, there is a practical reason to retain DNA samples for short
periods. Because DNA technology is changing so rapidly, we expect the profiles
produced with today's methods to be incompatible with tomorrow's methods.
Accordingly, today's profiles will need to be discarded and replaced with profiles
based on the successor methods. It would be extremely expensive and inefficient
to have to redraw blood samples for retyping. We are therefore persuaded that
retention of samples after typing should be permitted for the short term—only
during the startup phase of DNA profile databanks. As databanks become
established and technology stabilizes somewhat, samples should be destroyed
promptly after typing.
RULES ON ACCESSIBILITY
Computer security should be ensured through use of the best available
practices and technologies. Access to the databank should be limited to a small
number of legally authorized persons and should be limited to what is required
for specific official investigations. All instances of access should be audited and
archived. An excellent discussion of computerized audit-trail systems is
available.8
If the computer system and associated databank are to be made available for
remote access by cooperating state and federal agencies, such as by telephone or
networked by other means, the access mechanism (i.e., the network switch)
should be made available only for specific, authorized remote-access sessions;
that is, the system should not be continuously available to remote users. This type
of limited access can be achieved either administratively or physically; it is a
simple and inexpensive means of safeguarding sensitive information and is
common practice in many national security situations. For example, secure
computers are virtually never connected to unsecured computers at national
defense laboratories; when newspaper headlines make statements that computers
at these facilities have been breached, it has been the case that the computers
were unsecured and not connected to the secure computers. In many cases, these
unsecured computers have telecommunication connections available to
employees for routine use, but they do not contain security information.
State Level
According to a recent FBI survey,9 27% of 177 forensic science laboratories
responding indicated that they have legislative authority or a mandate to
construct a databank for their own jurisdictions to match DNA profiles. An
additional 38% believed that such authority or mandate was likely by 1991.
According to an Office of Technology Assessment survey conducted in late
1989,10 at least 17 states had passed or were considering legislation creating
statewide DNA databanks. The persons to be included in the databanks range from
sex offenders to all felons. Since the time of that survey,
the number has no doubt increased. Therefore, it is obvious that many state
legislatures recognize the potential benefit of a DNA databank as an important
investigative tool and that such databanks will become a reality. Many states are
already collecting samples in earnest, although at this writing no databanks are
operative.
Federal Level
The FBI and TWGDAM have proposed the creation of a national DNA
profile databank system, including one statistical and three investigative
databanks. The statistical databank would include DNA profiles of randomly
selected unrelated persons and would be built collaboratively and maintained by
the FBI for use by all forensic laboratories. The investigative databanks would
contain DNA profiles of body fluids from the scenes of crimes for which suspects
have been identified, convicted offenders, and bodies, body parts, and bone
fragments of unidentified persons. In the proposed national DNA profile databank
system, individual law-enforcement agencies (forensic laboratories) would
contribute DNA profiles (without personal information) to a centralized
databank, but retain absolute control of their own case records. The national
databanks would reference the sources of the profiles, but case data would be
secured and controlled by the state and local agencies.
In the national program, the FBI would play the lead role. It would
coordinate quality assurance with a technical advisory group to implement
appropriate guidelines; coordinate with other agencies that have a law-
enforcement interest in the development of the databank; provide hardware and
software for the databank server and for state access to the databank; provide
hardware to store and back up the databank server; provide training for states in
forensic DNA technology, quality control, and databank access; determine
formats for databank input and output; update index with new state and federal
submissions; assemble population data for all probes used and calculate and
disseminate population frequencies; and modify the system to accommodate new
DNA typing methods.
State and local agencies would be responsible for performing DNA analyses
of samples with consensus methods; submitting new information in a specified
format for incorporation into the databanks; guaranteeing the quality of their new
submissions; providing hardware and software for state image-analysis
workstations for telephone access to centralized index; maintaining centrally
indexed case files for as long as they remain in the index; and providing relevant
information from case files that are indexed centrally to other law-enforcement
agencies, which subscribe when requested.
Just as the Department of Defense keeps dental records and fingerprints
(with the FBI) of American soldiers, it is seeking funding to collect blood
samples from each soldier and establish a DNA profile databank. When a soldier
is killed and cannot be identified with usual methods, a sample of tissue, blood,
or bone marrow from the remains would be subjected to DNA analysis for
comparison with entries in the databank. There are 3.3 million active and reserve
members of the armed forces. Given the costs associated with the current
technology, a DNA databank of such scope would not be amenable to RFLP
analysis. The Armed Forces Institute of Pathology therefore proposes to begin
collecting and storing samples while working on the development of a DNA
analysis method, which when perfected will be much less expensive and time-
consuming than existing RFLP methods.
A databank of military personnel could also offer ancillary forensic
applications: criminal investigations conducted by criminal investigation
divisions of the armed forces could be aided in the same manner as those of other
law-enforcement agencies, identification of subjects for security purposes could
be enhanced, and identification of urine samples from disputed sources for drug
testing could be verified.
The present committee has not been asked to comment on this program; we
simply acknowledge its existence.
FIGURE 5-1 Hypothetical national information resource. Data flow starts with
forensic laboratories in various states that provide raw data. Data reduction
process provides information to national information resource databank. Merged
and reduced data are provided only to authorized users.
at the state or regional level. Thus, all the information concerning the sample,
experimental, and result contexts would be stored at the state or regional level;
only data associated with the result context would be accessed at the national
level. The box labeled ''Data Reduction Process" in the center of the figure ideally
represents a standardized method for DNA typing that all laboratories use.
The hypothetical national information resource shown in Figure 5-1does not
necessarily represent a physical entity, but could be simply a view of derived data
from all the various regional databanks. A view would be achieved by having
access software sitting "on top" of the various state or regional databanks. The
software would have distinctly different requirements for level of access to data
in the databank. For example, "outside views" would only need access to VNTR
profiles and some arbitrary identification number; no further information at this
first level of access would be required for initial identification searching.
SUMMARY OF RECOMMENDATIONS
There are a number of scenarios that illustrate the point that the databank
need not be limited to persons convicted of specified crimes.
REFERENCES
1. Wilson T. Automated fingerprint identification systems. Law Enfore Technol. 1986. Aug-
Sep:17-20, 45-48.
2. Dalenius T. Towards a methodology and statistical disclosure control. Statistik Tid-skrift.
15:213-225, 1977.
3. Adam NR, Wortmann JC. Security-control methods for statistical databases: a comparative study.
ACM Comput Surv, 21:515-556, 1989.
4. Code of Virginia, Title 19.2-310.6. Unauthorized uses of DNA data bank; forensic samples;
penalties (1990, c. 669).
5.Jones v. Murray,763 F. Supp. 842 (W.D. Va. 1990).
6. Cellmark Diagnostics. Proposal for DNA databasing, Division of Criminal Investigation Forensic
Laboratory, South Dakota. Dec. 21, 1990.
7. Beck A, Shipley B. Recidivism of prisoners released in 1983. Bureau of Justice Statistics Special
Report NCJ-16261. Washington, D.C., 1989.
8. Lunt TF, Tamaru A, Gilham F, Jagannathan R, Neuman PG, Jalali C. IDES: a progress report.
Proceedings of the Sixth Annual Computer Security Applications , Tucson, Arizona: ACM
Press, 1990.
9. Miller J. The outlook for forensic DNA testing in the United States. Crime Lab Digest. 17(Suppl.
1 ):1-14, 1990.
10. U.S. Congress, Office of Technology Assessment. Genetic witness: forensic uses of DNA test.
OTA-BA-438. Washington, D.C.: U.S. Government Printing Office, 1990.
6
Use of DNA Information in the Legal System
ADMISSIBILITY
In the United States, there are two main tests for admissibility of scientific
information from experts. One is the Frye test, enunciated in Frye v. United
States.1 The other is a "helpfulness" standard found in the Federal Rules of
Evidence and many of its state counterparts. In addition, several states have
recently enacted laws that essentially mandate the admission of DNA typing
evidence.
of reliability are raised. For example, forensic DNA analysis frequently involves
the use of small, possibly contaminated samples of unknown origin, such as a
dried blood stain on a piece of clothing. Some experts have questioned the
reliability of DNA analysis of samples subjected to "crime scene" conditions. In
addition (as noted in Chapters 2 and 3), the details of the particular techniques
used to perform DNA typing and to resolve ambiguities evoke a host of
methodological questions. It is usually appropriate to evaluate these matters case
by case in accordance with the standards and cautions contained in earlier
portions of this report, rather than generally excluding DNA evidence. Of
particular importance once such a system of quality assurance is established
would be a demonstration that the involved laboratory is appropriately accredited
and its personnel certified. Some aspects (such as the validity of the theory
underlying RFLP analysis) might be so well established that judicial notice is
warranted. Others (such as quantitative correction of band shifting with a single
monomorphic fragment) might not be sufficiently well established to justify
admissibility.
Assumption 3—related to the adequacy of statistical databanks used to
calculate match probabilities—rests on unproven foundations. Many experts
question the adequacy of current databanks for making probability estimates, and
the use of multiplicative modes of combining probabilities are also open to
serious question (see Chapter 3). The solution, however, is not to bar DNA
evidence, but to ensure that estimates of the probability that a match between a
person's DNA and evidence DNA could occur by chance are appropriately
conservative (as described in Chapter 3).
The validity of assumption 4—that the analytical work done for a particular
trial comports with proper procedure—can be resolved only case by case and is
always open to question, even if the general reliability of DNA typing is fully
accepted in the scientific community. The DNA evidence should not be
admissible if the proper procedures were not followed. Moreover, even if a court
finds DNA evidence admissible because proper procedures were followed, the
probative force of the evidence will depend on the quality of the laboratory work.
More control can be exercised by the court in deciding whether the general
practices in the laboratory or the theories that a laboratory uses accord with
acceptable scientific standards. Even if the general scientific principles and
techniques are accepted by experts in the field, the same experts could testify that
the work done in a particular case was so flawed that the court should decide
that, under Frye, the jury should not hear the evidence.
The Frye test sometimes prevents scientific evidence from being presented
to a jury unless it has sufficient history to be accepted by some subspecialty of
science. Under Frye, potentially helpful evidence may be excluded until
consensus has developed.3 By 1991, DNA evidence had been considered in
hundreds of Frye hearings involving felony prosecutions
in more than 40 states. The overwhelming majority of trial courts ruled that such
evidence was admissible; there have been some important exceptions, however.
The first scientifically thorough Frye hearing concerning DNA typing was
conducted in People v. Castro,4 in which a New York trial court concluded that
the theory underlying DNA typing is generally accepted by scientists in genetics
and related fields, that forensic DNA typing has also been accepted and is
reliable, but that the technique as applied in the particular case was so flawed that
evidence of a match was inadmissible (although evidence of an exclusion was
admissible). The Castro court stated that the focus of the Frye test as applied to
DNA typing (or any other novel scientific evidence of similar complexity) must
include its application to the particular case. It held that flaws in the application
are not simply questions as to the weight to be given the evidence by the jury, but
go to admissibility as determined by the judge.5Castro determined that there were
serious flaws in the laboratory's declaration of a match between two samples, for a
number of reasons, including the presence of several anomalous bands. The court
did not credit the laboratory's explanation of the reasons for the anomalies and
criticized its failure to perform adequate follow-up testing. In addition, the court
concluded that the laboratory's population-frequency databank could not provide
an accurate estimate of the likelihood that the defendant was the source of the
DNA. The court's analysis and findings were careful, and they have generally
been approved by experts in the field.
In November 1989, the Supreme Court of Minnesota, deciding Statev.
Schwartz,6 became the first appellate court to reject the use of DNA evidence
analyzed by a forensic laboratory. In answering a certified question, the court
noted that "DNA typing has gained general acceptance in the scientific
community." Nevertheless, the court went on to hold that admissibility of specific
test results in a particular case hinges on the laboratory's compliance with
appropriate standards and controls and on the availability of its testing data and
results. It held that, "because the laboratory in this case did not comport with
these guidelines, the test results lack foundational adequacy and, without more,
are thus inadmissible." One matter that troubled the court was the failure of the
testing laboratory to reveal underlying population data and testing methods. The
court noted that the reliability of a test implies that it could be subjected to an
independent scientific assessment of the methods, including replication of the
test. Because such independent assessment had not occurred and could not take
place, owing to the laboratory's secrecy, the court held that the results were
inadmissible. In addition, the court was concerned that the testing laboratory (1)
had admitted having falsely identified two of 44 samples as coming from the
sample subject during a proficiency test performed by the California Asso
ciation of Crime Laboratory Directors and (2) had not satisfied relevant validation
protocols used by the FBI. In that regard, Schwartz makes a good case for
requiring laboratories to meet particular standards before they may provide
analysis of evidence to juries. Schwartzalso held that the use of population-
frequency statistics must be limited, because "there is a real danger that the jury
will use the evidence as a measure of the probability of the defendant's guilt or
innocence, and the evidence will thereby undermine the presumption of
innocence, erode the values served by the reasonable double standard, and
dehumanize our system of justice."7 The decision in Schwartzwas influenced by
Minnesota's unique position in limiting the use of probability estimates in trials.8
A new Minnesota statute not considered in Schwartz specifically requires
judges to admit population-frequency data generated by DNA testing. Thus, it is
not clear how influential Schwartz will be in its home state. Nevertheless, the
Minnesota judges' skepticism about statistical analysis is shared by other judges.
Particularly in regard to DNA typing, the manner in which probabilities should be
calculated requires great care.
In Cobey v. State,9 the Maryland Court of Special Appeals reached a
conclusion opposite to Schwartz, holding that evidence of DNA analysis from the
same laboratory that figured in Schwartz was admissible and finding that the
laboratory's databank was sound. The Cobey court was impressed by the absence
of expert testimony contradicting that in favor of admissibility. It did caution,
however, that "we are not, at this juncture, holding that DNA fingerprinting is now
admissible willy-nilly in all criminal trials." In 1989, Maryland became one of a
growing number of states to enact a law recognizing the admissibility of DNA
evidence.
thorough hearings on the admissibility of DNA evidence, with two courts finding
it admissible and one ruling it inadmissible.
The U.S. District Court for the District of Vermont conducted a detailed
analysis in United States v. Jakobetz.19 It reviewed the literature and FBI
practices. Despite a strong attack from the defense and its experts, the court found
that the DNA evidence was "highly reliable" and that its probative value
outweighed the potential for prejudice.20Strict application of the Frye test was
rejected in accordance with Second Circuit standards.21
After a thorough hearing that focused on FBI protocols, the U.S. magistrate
for the Southern District of Ohio in United States v.Yee22 also wrote a detailed
analysis with conclusions essentially tracking those in the Vermont case.
(Interestingly, an Arizona trial court considering the admissibility of DNA typing
in State v. Despain23carefully studied the transcript of Yee, but reached a
conclusion opposite to it. That might have been because it also reviewed the
transcript of another hearing in which four additional defense experts challenged
FBI protocols. Finding that there was a legitimate scientific controversy as to the
validity of DNA testing and that it had not gained general acceptance, the court in
Despain refused to admit evidence analyzed by the FBI laboratory.)
Most recently, the Superior Court for the District of Columbia reached the
opposite conclusion and held DNA typing inadmissible. In U.S.v. Porter24, the
court ruled that the technical reliability of DNA typing was generally accepted,
but that the FBI's method for calculating the probability of a coincidental match
was not. The court ruled that the scientific foundation of these probability
calculations bears on the admissibility (and not simply the weight) of the
evidence. Applying the Frye standard, the court found that "there is a controversy
within the scientific community [on this issue] which has generated further study,
the results of which will soon be available for scrutiny. It is after these studies and
others … when the court should be called upon to admit DNA evidence."
In addition, a number of state courts that apply analogues of the federal rules
have considered the admissibility of DNA evidence. In Andrews v. State,25 a
Florida court of appeals (the first higher-level state court to consider DNA
evidence) determined that the relevance approach was applicable under the
Florida evidence code that tracks the federal rules. The court admitted the
evidence presented by the plaintiff's three scientific experts, two of whom worked
for a private testing laboratory; the defense called no experts. The court
concluded that the DNA typing evidence offered by the plaintiff was clearly
helpful to the jury. With respect to the possibility of prejudice, the court found
that DNA typing is not particularly "novel," in that it had been used in
nonforensic applications for 10 years. The issue of differences between scientific
applications and forensic applications were not raised by the defense. The court
also noted the existence of specialized literature about the technique. As for the
possibility of erroneous test re
sults, the court credited testimony that an error in the testing process would mean
that there would be no result, rather than a false-positive or false-negative result.
The court also credited the efficacy of the laboratory's control runs and approved
the use of statistical data to determine the probability of a match.
In Spencer v. Commonwealth,26 the Supreme Court of Virginia affirmed a
trial court's finding that evidence derived from RFLP analysis was sufficiently
reliable to be admitted. The trial court heard testimony from three experts for the
prosecution in molecular biology and genetics. The defense called no expert
witnesses. The trial court credited testimony that there is no risk of false
positives, that the testing techniques are reliable and generally accepted in the
scientific community, and that the particular test was conducted in a reliable
manner.
At a later proceeding involving the same defendant, the Supreme Court of
Virginia held that evidence based on a sample analysis that used a PCR
technology was admissible. In discussing the standard for admitting novel
scientific evidence, it rejected the Frye test, asserting instead that the court should
make a ''threshold finding of fact with respect to the reliability of the scientific
method offered." Without discussing the details of the experts' testimony, the
court concluded that the evidence supporting admissibility was credible.27
A Delaware trial court held in State v. Pennell28 that DNA evidence was
admissible under a state statute similar to the federal rules, but refused to admit
probability statistics. There was no dispute about the underlying theory of DNA
typing or its general application in the particular case. The defendant challenged
the laboratory's claims that the population databank it used was in Hardy-
Weinberg equilibrium and that its "binning process" was valid. The defense held
that the state's experts' assessment of the probability of declaring a match was
overstated. The court accepted some of the defense contentions and faulted the
laboratory for its procedure. The state later introduced new evidence based on the
laboratory's revised procedure and a new databank. The court agreed to allow the
new evidence if the state would provide the raw data to the defendant, but the
state did not do so. The court expressed concern over testimony that the
measurements of allele size can depend on who is doing the measuring, and it
concluded that the state's evidence did not sufficiently support the probability
calculation.
sufficient controversy about the methods for assigning statistical weight so that
they could not be considered generally accepted. In the sole federal appellate
ruling, the Eighth Circuit Court of Appeals reversed a federal trial court's decision
to admit DNA typing evidence and directed the lower court to hold a full hearing
on admissibility.29In the spring and summer of 1990, an intermediate-level
appellate court in Texas30 and the supreme courts of South Carolina,31
Georgia,32North Carolina,33 and Massachusetts34 were among the courts that
considered the admissibility of DNA evidence. These opinions are of particular
interest, because they were issued after sustained debate in the legal and scientific
communities about possible flaws in DNA typing technology and possible
inadequacies in the population databanks. The courts in Texas, South Carolina,
Georgia, and North Carolina upheld the admissibility of DNA evidence;
Massachusetts rejected it because of concerns about the adequacy of population
genetic interpretation.
In Kelly v. Texas, the defendant appealed from a murder conviction,
challenging as error the trial court's admission of evidence that compared a semen
sample from the crime scene to a blood sample of the suspect. The defendant did
not challenge the principles of DNA typing or the general qualifications of the
state's five experts. He did attack the methods of the testing laboratory and the
statistical expertise of the witnesses. The appellate court was informed that
outside experts had twice verified the laboratory's procedures and results. In
upholding the trial court's decision to admit the evidence, the appellate court
specifically acknowledged the "validity" of the laboratory's techniques.
In July 1990, the Supreme Court of Georgia decided Caldwell v. State, a
death-penalty case. The appeal grew out of a trial court's decision after a Frye
hearing (that involved testimony by 10 experts) to admit DNA evidence. Both at
the Frye hearing and on the appeal, no challenge was made to the scientific
principles or general techniques used by the forensic laboratory. The focus was on
how the laboratory declared a match between samples, the validity of its
probability calculations, and its procedures to ensure quality control. In deciding
the appeal, the court first considered whether it was appropriate for the trial court
to use a Frye hearing to determine whether the laboratory had performed its test
with reliable techniques and in an acceptable manner. It concluded that, because
of the complexity of the issues and a lack of national standards, the inquiry was
appropriate. Although noting that errors, including false positives, could occur,
the court ruled that the laboratory's protocol was "adequate to meet these
concerns."
The court addressed how the laboratory had conducted a band shift analysis
and calculated the power of identity. Despite band shifting, the laboratory had
originally decided a match by visual examination. During the course of the trial,
as a result of criticism of that technique, it reana
lyzed the samples with a monomorphic probe. Such a probe provides an arguably
invariant reference point to analyze band shifts across samples. After review of
the testimony concerning the reanalysis, the appellate court concluded that this
approach to the problem of band shifting was acceptable.
The appellant in Caldwell also attacked the calculations that led the testing
laboratory to conclude that the chance that a randomly selected person would
have the same DNA pattern as that of the sample source and the suspect was 1 in
24,000,000. Only one of the 10 experts had actually examined the laboratory's
population databank, and he, a defense witness, insisted that it was not in Hardy-
Weinberg equilibrium. The court ruled that, in the absence of supporting
testimony, the probability statement generated by the laboratory assumptions
could not be accepted. But the court did accept the concept of appropriate
statistical calculations, which it erroneously thought did not depend on population
theory. (See the discussion of the population genetics question in Caldwell in
Chapter 3.)
In January 1991, the Supreme Judicial Court of Massachusetts, in
Commonwealth v. Curnin,34 became the second state supreme court to refuse to
admit DNA typing evidence. After being convicted of rape, in part on the basis of
DNA typing evidence, the defendant appealed, arguing that there was no general
agreement concerning test methods, use of control samples, or the need for a
testing laboratory to meet external performance standards. The high court did not
address those arguments, focusing instead on the "lack of inherent rationality" of
the process by which the testing laboratory concluded that 1 Caucasian in
59,000,000 would have the DNA pattern represented by the semen stain and the
defendant's blood. The court was particularly impressed by the testimony of an
expert for the defense who criticized the product rule as unsupported by the
laboratory's reference databank, raised the possibility of calculation errors due to
ignorance of population substructure, and explained why no assumption would be
made as to whether the relevant population was in Hardy-Weinberg equilibrium.
Despite its decision to reverse the trial court, the high court made clear that it
would not be surprised if the prosecution could correct the weaknesses of its
testimony. In the court's words, "it may even be that, by the time of the retrial of
this case the prosecution can support the admissibility of evidence of the
probability of the alleles disclosed by the DNA test being found elsewhere in the
human population. …"
Admissibility Statutes
Since 1987, the admissibility of DNA typing evidence was raised repeatedly
in the courts, largely in the context of Frye hearings. Challenges to admissibility
have become more sophisticated over the last 2 years. State legislatures have
recently begun to address the matter. Several states have
enacted laws that declare that appropriately performed DNA tests are admissible.
Although they do not specify what an appropriate test is, the statutes must have
been passed with single-locus RFLP analyses by Southern blotting in mind.
Arguably, some of them should not be interpreted as applying to technologies
that were not in general use and therefore could not have been evaluated by the
legislatures that passed the statutes. Such technologies could be validated by
amended statutes or by courts in Frye or Rule 702 hearings. For most purposes,
states with such laws have statutorily resolved disagreements over the scientific
reliability of DNA testing, although the questions of whether tests were
performed properly in a given case and of the adequacy of statistical calculations
based on test results probably remain subject to challenge.
The state laws are of two types. A number of states—including Arkansas,
Connecticut, Michigan, Montana, and New Mexico—now specifically admit
DNA evidence to assist in the resolution of paternity—noncriminal—cases (and,
by inference, probably other disputes concerning biological relationships).35
Louisiana, Maryland, Minnesota, Virginia, and Washington have enacted laws
that recognize the admissibility of DNA evidence in criminal cases.36
Maryland requires that the DNA report be delivered to the defendant 2
weeks before the criminal proceeding and specifies that the defendant may
require a witness who analyzed the sample to testify as to the chain of custody.
The Minnesota statute states that in any civil or criminal trial or hearing DNA
evidence is admissible without "antecedent expert testimony that DNA analysis
provides a trustworthy and reliable method of identifying characteristics in an
individual's genetic material upon a showing that the offered testimony meets the
standards for admissibility set forth in the Rules of Evidence"; a companion
provision specifically permits the admission of "statistical population frequency
evidence … to demonstrate the fraction of the population that would have the
same combination of genetic markers as was found in a specific human
biological specimen." Louisiana provides that "evidence of deoxyribonucleic acid
profiles, genetic markers of the blood, and secretor status of the saliva offered to
establish the identity of the offender of any crime is relevant as proof in
conformity with the Louisiana Code of Evidence.''
Legislative interest in DNA evidence remains active, and it is likely that
other states will enact laws generally favorable to its admissibility.
persons. For some DNA typing methods, the technical basis is well accepted. For
others, important scientific questions must be resolved before they are appropriate
to use in court.
New developments in DNA technology probably will, and at first should, be
the subject of in limine hearings (those conducted by a court in deciding on
admissibility), as has been the case in recent instances when present technology
has been tested. As a general rule, generation of evidence with such new
technology should be encouraged if it is adequately supported in court hearings.
It is highly desirable that experts in molecular genetics and statistical analysis
review new developments and pass on them at a variety of conferences and
through published papers. Until there is some consensus in this field, results of
using new techniques may not be admitted; a testing period for the new
techniques will be needed to determine whether there are unforeseen errors or
difficulties, and it will take time to compile the necessary databanks. Otherwise,
the normal rules with respect to new developments can be relied on. In fact, new
developments should present less difficulty than has been posed by present DNA
typing technology, because much of the theory will have already been tested and
accepted by the courts.
The issue for courts will be to discern when a technology is so different as to
require a full admissibility hearing. Admissibility hearings might be required to
evaluate the underlying principle of a scientific method of identification, the
particular method for applying the principle, and the performance of a test in a
particular case. Regarding the underlying principles, there is, as we have noted,
no longer any question concerning the principle that DNA can be used to obtain
identification information; admissibility hearings need no longer address the
question. Regarding the particular method for applying the principle, the inquiry
will depend on the new method or technology. For example, use of a previously
unused DNA probe in the context of the basic RFLP technique might require an
admissibility hearing on whether the properties of the particular probe (e.g.,
pattern, sensitivity, or population genetics) are scientifically accepted. Methods
of correcting for shifted DNA patterns (that would otherwise fall outside the
usual matching rule) might require an admissibility hearing concerning whether
the correction procedure has gained scientific acceptance, inasmuch as this
substantially changes the method of declaring a match. The use of PCR
amplification for sample preparation might require a pretrial hearing on the
properties of the technique, because it introduces a novel issue considered by only a
few courts thus far—the synthesis of evidence by amplification. And the use of
various detection technologies for PCR products might require a pretrial hearing
about the characteristics of the detection method and its sensitivity to artifacts. In
each case, the court can properly limit inquiry to the substantially novel aspects
of the technology, focusing on whether
the method is accepted for scientific applications and whether it has been
validated for forensic identification. Minor changes in protocols will typically not
require pretrial hearings, unless they are likely to affect key issues (such as the
matching rule).
who testify should remain open to inquiry. Ultimately, DNA typing evidence
should be used without any greater inconvenience than traditional fingerprint
evidence.
The Jury
Because a jury might overvalue or undervalue scientific evidence, it is
appropriate where permitted for the judge to question DNA experts with an eye to
aiding the jury. The judge can explain to the jury the role of experts and the role
of the jury in evaluating the experts' opinions.
When probability statements are admissible, the judge should not be
expected to instruct the jury in detail on how probabilities are computed or how
probabilities available from an analysis of DNA material should be combined
with probability estimates based on more traditional testimony and other
evidence. Those matters are better left to the experts and to the lawyers on
summation. The court should encourage the use of charts, written reports, and
duplicates of materials that are relied on by the experts, so that the jury can be as
well educated as possible in the evaluation of DNA typing evidence. To that end,
the court should insist that technical terms be reduced to understandable lay
language and that scientific information be presented to the jury in the least
confusing form possible.
Special forms of charges are not required. DNA typing may be assessed
within the framework of normal forensic laboratory work and can be readily
handled with the present rules and forms of charges.
The Prosecutor
The prosecutor will work closely with the investigators and will normally
have access to adequately staffed and organized forensic laboratories. The
prosecutor should carefully supervise the investigation activities to ensure that
DNA typing evidence will be admissible, if it proves relevant.
The prosecutor has a strong responsibility to reveal fully to defense counsel
and experts retained by the defendant all material that might be necessary in
evaluating the evidence. That includes information on tests that proved
inconclusive, on retesting, and on the testing of other persons. Adoption of rules
or statutes that require the prosecutor to involve the defense in analysis of DNA
samples at the earliest possible moment is highly recommended.
The committee recommends going beyond what is required by the fed
The Defense
Defense counsel must have access to adequate expert assistance, even when
the admissibility of the results of analytical techniques is not in question, because
there is still a need to review the quality of the laboratory work and the
interpretation of the results. When the prosecutor proposes to use DNA typing
evidence or when it has been used in the investigation of the case, an expert
should be routinely available to the defendant. If necessary, he or she should be
able to apply for funds early in the discovery stages to retain experts without a
showing of relevance that might reveal trial strategy. Whenever possible, a
portion of the DNA sample should be preserved for independent analysis by the
defense.
The prosecutor should promptly reveal to defense counsel that DNA was
involved in the investigation and might be available for analysis at the trial.
Normally, the criminal-justice system will not provide for the appointment of
counsel for the defendant or for payment for experts until the defendant has been
arrested or charged. Where a sample of the defendant's tissue is sought for DNA
typing, application to the court for DNA experts should be possible even before
an arrest has been made.
In our judicial system, jurors are relatively independent. Nevertheless,
through limitations on the admissibility of evidence and on the form of its
presentation and through the use of a variety of instructions, the court exercises
considerable influence. DNA evidence, like other scientific and statistical
evidence, can pose special problems of jury comprehension. Courts and attorneys
should cooperate to facilitate jury understanding. Innovative techniques, such as
allowing jurors to take notes or ask questions, might be considered. Jargon should
be avoided, and information should be presented simply, clearly, and fairly.
Unless limited by law or court rules, judges should be free to pose questions to
witnesses when they feel that the answers might clarify the testimony. Reports
and relevant materials should be admitted into evidence so that they can be
studied by courts at their leisure. Finally, a judge would not be amiss in pointing
out to attorneys the wisdom of including jurors who are found to have a
background that enhances their ability to understand the expert testimony.
TESTING LABORATORIES
Other chapters have indicated appropriate standards for the operation of
testing laboratories and the collection and analysis of DNA samples. Uniformity
in reporting, completeness of reporting (including laboratory protocols and
written criteria for interpretation), and stringent quality assurance of laboratories
are essential. The court and the jury should have no reason to doubt the accuracy
of the processing of information. Laboratories and experts have a particular
responsibility to ensure that they are open and candid with the courts. Any
reservations about inadequacies or errors should be promptly revealed, and failure
to do that should be dealt with seriously. The court should not hesitate to exercise
contempt powers and exclude experts who have misled deliberately in the past.
Private trade associations and other appropriate groups should also apply pressure
to ensure accuracy and candor.
PROTECTIVE ORDERS
Protective orders should not be used to prevent experts on either side from
obtaining all relevant information, which can include original materials, data
sheets, software protocols, and information about unpublished databanks. A
protective order might be appropriate to limit disclosures by attorneys and experts
to third parties about proprietary information acquired in the course of a
particular case; but as a general rule, any scientific information used in a case
should be open to widespread scientific scrutiny. One exception might be when
the expert is involved in a current or recently completed study on which he or she
does not directly rely to develop an opinion. That will ensure that the expert does
not lose his or her opportunity to publish as a consequence of testifying.
Protective orders to prevent unnecessary intrusion into the privacy of such
persons as those who have been cleared after investigation or who are juveniles
are appropriate.
SUMMARY OF RECOMMENDATIONS
Having carefully reviewed the issues, the committee offers the following
recommendations:
28. 1989 WL 167430 at ll (Del. Super. Ct. Nov. 6, 1989) (Gebelein, J.).
29. U.S.v. Matthew Sylvester Two Bulls,918 F.2d 56 (8th Cir. 1990), 925 F. 2d 1127 (8th Cir. 1991)
(ca bane), vacated after death of defendant. See Weinstein, Rule 702 of the Federal Rules of
Evidence is sound; it should be amended (138 F.R.D. 1991) (discussing 7'wo Bulls).
30. Kelly v. Texas,No. 2089-026-CR (Ct. of Appeals, 2d District, Fort Worth, June 27, 1990).
31. State v. Ford,S.C., 392 S, E, 2d, 78.3 (1990).
32. Caldwell v. Stale,260 Ga. 278, 393 S.E. 2d. 436 (1990).
33. State v. Pennington,327 N.C. 89, 393 S.E. 2d 847 (1990).
34. Commonwealth v. Curnin,409 Mass. 218, 565 NE 2d 443 (1991).
35. Arkansas Act 723 (1989); Connecticut P.A. No. 89-360 (1989); Michigan Public Act 258 11989);
Montana Code Ann. Sect. 40-5-201 (1989); New Mexico Slat. Ann. Sect. 40-11-5 (1989).
36. Virginia Code Sect. 19.2-270.5; Louisiana Act 340 11989); Maryland Chap. 430 11989);
Minnesota Stat Sect. 634.23 (1989); Washington Chap. 350 (1989).
37. Cal. Penal Code 290.2; Colo. Code Ann. 17-2-201(g)(I); 111. Code Ann. 38 1005-4-3; S. Dak.
Code Ann. 23-544 et seq; Ariz. Code Ann. 3128. 1 Fla. Code Ann. 943.325; Iowa Code
Ann. 13.10; Minn. Code Ann. 609.3461; Nev. Code Ann. 176.111; Wash. Code Ann.
43.43.754; Virg. Code Ann. 19.2-310.2.
38. Report of the Joint Subcommittee Studying Creation of a DNA Test Data Exchange to the
Governor and the General Assembly of Virginia. Senate Document No. 29, Commonwealth
of Virginia, Richmond, 1990, p.I 1.
39. Lawrence R. Jones, el al. v.Edward W. Murray,Director of The Department of Corrections, et al.
W.D. Virginia, Civil No. 90-0572-R. Order for Summary Judgment, March 4, 1991.
40. See, in general, Ballantyne J, Sensabaugh G, Witkowski J., eds. DNA Technology arid Forensic
Science. Banbury Report 32. Cold Spring Harbor, New York: Cold Spring Harbor
Laboratory Press, 1989.
7
DNA Typing and Society
The introduction of any new technology is likely to raise concerns about its
impact on society. Financial costs, potential harm to the interests of individuals,
and threats to liberty or privacy are only a few of the worries typically voiced
when a new technology is on the horizon. DNA typing technology has the
potential for uncovering and revealing a great deal of information that most
people consider to be intensely private.
The federally established human genome program will yield an
unprecedented amount of genetic information and generate new databanks.1,2
Even apart from the human genome program, DNA technology is moving
forward; but this large-scale program, projected to take 10-15 years, is bound to
accelerate the acquisition of genetic information. At the same time, it contains a
mandate for examining the ethical, social, and legal implications of mapping the
human genome, with specific allocation of funds for examining these aspects.2 A
central concern raised by these developments is the safeguarding of the
confidentiality of personal genetic information. With greater understanding of the
human genome, the potential of misuse of DNA samples collected or preserved
for purposes of criminal justice will increase. The more databanks are
established, the greater the risk of breaches of confidentiality and misuse of the
information.3
The social, economic, and ethical concerns in this chapter overlap with the
legal aspects addressed in Chapter 6 and the issues in development and use of
databanks discussed in Chapter 5.
ECONOMIC ASPECTS
The forensic use of DNA technology will have various economic impacts.
The proliferation of DNA evidence in investigations and trials requires a fairly
rapid expansion in the number of reliable experts and laboratories. The cost of the
equipment, training and proficiency programs, supplies, and personnel will be
very large. For example, the three proposed regional laboratories in New York
state are estimated to cost $1.4 million per year.4 The Commonwealth of Virginia
has committed several million dollars over the last 3 years to its forensic DNA
activities (Paul Ferrara, personal communication, 1990). Material will have to be
stored for databanks and for checking suspects. Costs will be associated with the
upgrading and changing of databanks when new procedures are adopted. Those
costs will affect budgets for police, prosecutors, and courts. Indigent criminal
defendants might have a constitutional due-process right to have an expert
witness paid for by the government.
The courts themselves must be supplied with reliable assistance in
evaluating DNA material. In the federal system, the court can request an expert
or panel of experts to assist it, pursuant to Rule 706 of the Federal Rules of
Evidence. A special register of scientific experts can be maintained for ready
access. The government will generally have to bear this cost. However, if a
defendant can afford the cost and asks for expert assistance, the court can assess
some costs against the defendant and some against the state.
New costs will also be related to training and certification. The
implementation of any new technology requires training and certifying of
personnel. Additional costs will be incurred to develop mechanisms to ensure
quality control of laboratories that conduct forensic DNA testing.
New technology can grow and make ever larger fiscal demands on society.
It is difficult to predict the total cost of DNA testing when it becomes generally
available nationwide, but it is reasonable to expect it to amount to tens of millions
of dollars a year. That cost is unavoidable, but, given the present fiscal problems
at all levels of government, cannot be ignored. Setting up regional and
cooperative services is one way of controlling costs. It might not be feasible or
appropriate for some small forensic science laboratories to create their own DNA
testing capabilities. A major DNA testing center run by the FBI might reduce
costs to smaller localities. That potential reduction in monetary cost needs to be
balanced against the risks to privacy and confidentiality of having a powerful
federal law-enforcement agency in charge of DNA testing and storage of DNA
information. If laboratories come to share information, everything could
eventually become linked. At the same time, the risks that privacy and
confidentiality will be breached might be as great or greater with local
control, in that state laws governing the use of criminal records vary widely.
It is likely that the cost of criminal justice will be increased. In some cases,
however, early exclusion of suspects who have been cleared by forensic DNA
evidence will reduce cost in the judicial system. On balance, the increased costs
are small relative to the cost of operating the entire system. The committee
believes that the expenditures are warranted by the advantages to be expected.
ETHICAL ASPECTS
Ethical considerations regarding the use of DNA technology in forensic
science overlap with various issues addressed in social and legal analyses,5
including substantive and procedural rights of people and overall nonmonetary
costs and benefits likely to result from establishing the use of the new technology
in courtroom proceedings.
A threshold question for any ethical inquiry is whether the action or practice
under discussion is intrinsically wrong. An action or practice is intrinsically
wrong if it violates fundamental ethical principles. These have traditionally been
held to include prohibitions against enslavement, torture, gratuitous infliction of
harm on human beings, and modes of exploitation that use humans as merely a
means (usually without their knowledge or consent) to serve the ends of others.6
To hold that such actions or practices are intrinsically wrong is not to claim that
they can never be justified. For example, if torturing a terrorist who knows the
location of a bomb planted to kill a million people is the only way to avert the
tragedy, then torture might be justified. That would not yield the conclusion that
torture is ethically right, but rather would show that evil acts can sometimes—
albeit rarely—be justified as a means of preventing much greater harm.6
DNA technology in forensic science is unlikely to violate any fundamental
ethical principle of the type described above. Although DNA technology involves
new scientific techniques for identifying or excluding people, the techniques are
extensions and analogues of techniques long used in forensic science, such as
serological and fingerprint examinations, handwriting analyses, photography, and
examination of teeth. Ethical questions can be raised about other aspects of this
new technology, but it cannot be seen as violating a fundamental ethical
principle.
A new practice or technology can be subjected to further ethical analysis by
using two leading ethical perspectives. The first examines the action or practice in
terms of the rights of people who are affected; the second explores the potential
positive and negative consequences (nonmonetary costs and benefits) of the
action or practice, in an attempt to determine whether the potential good
consequences outweigh the bad.6
Moral Rights
Two main questions can be asked about rights: Does the use of DNA
technology give rise to any new rights not already recognized? Does the use of
DNA technology enhance, endanger, or diminish the rights of anyone who
becomes involved in legal proceedings? In answer to the first question, it is hard
to think of any new moral rights not already recognized that come into play with
the introduction of DNA technology into forensic science. The answer to the
second question requires a specification of the classes of people whose rights
might be affected and what those rights might be.
The people whose rights might be endangered or diminished seem to be
chiefly those who are suspected or accused of or indicted for a crime or involved
in other legal proceedings, such as paternity suits, denaturalization, or
immigration matters. Does use of DNA technology interfere with or diminish
their rights in any way? Might it enhance their rights? Which rights might be
endangered?
The current use of DNA technology appears to pose no greater threat to the
right to privacy than does normal fingerprinting, placement of photographs in
evidence, collection of blood or saliva samples, or other established forensic
techniques. DNA technology is not different in principle from those other
techniques, although it holds the promise of providing a more definitive
identification than most others (fingerprinting is likely to remain the best for a
while). If the use of DNA information can be strictly limited to defendant
identification, it involves no greater intrusion into the privacy of an accused
person than do traditional methods in forensic science, whose aim is to make as
definitive an identification as possible. Without strict limits, however, DNA
information can be more intrusive into privacy, in that it provides more
information about a person.
In some ways, the use of DNA information about suspects can be less
intrusive than traditional methods. ''Rounding up the usual suspects" by checking
a DNA sample against a computerized databank is both much easier and less
intrusive than rounding up the suspects themselves. But people who are rounded
up are made aware that they are under suspicion and can take protective steps.
Where databanks already exist, a fresh blood sample would have to be taken from
suspects for confirmation. Thus, it is a complex matter to determine whether the
rights of suspects are enhanced or endangered by the use of DNA evidence in the
forensic setting, which requires empirical evidence to be subjected to careful
analysis.
Concerns about intrusions into privacy and breaches of confidentiality
regarding the use of DNA technology in such enterprises as gene mapping are
frequently voiced, and they are legitimate ethical worries.1,6,7,8The concerns are
pertinent to the role of DNA technology in forensic science, as
well as to its widespread use for other purposes and in other social contexts. A
potential problem related to the confidentiality of any information obtained is the
safeguarding of the information and the prevention of its unauthorized release or
dissemination;5,7 that can also be classified under the heading of abuse and
misuse (discussed below), as well as seen as a violation of individual rights in the
forensic context.
People have a right not to be wrongly convicted of a crime. To protect that
right, a high standard of proof is imposed before a person may be found guilty. In
addition, techniques used in gathering and analyzing evidence must have proven
reliability (comprising accuracy, precision, specificity, and sensitivity) and should
be accepted by a consensus of the scientific community. If DNA technology is as
good as or better than other methods used to identify criminals and if the
implications and limitations of DNA evidence are recognized by judges and
jurors, its use should pose no greater danger to the rights of accused people than
the use of currently approved techniques of forensic identification. Moreover, the
reliability of DNA evidence will permit it to exonerate some people who would
have been wrongfully accused or convicted without it. Therefore, DNA
identification is not only a way of securing convictions; it is also a way of
excluding suspects who might otherwise be falsely charged with and convicted of
serious crimes.
will show the same markers as a person in question can range from 1 in 2 (such
as in type O) to 1 in several thousand (such as when many systems are typed and a
relatively rare type is found). The literature and case law on paternity disputes
suggest that a likelihood of 1 in 20 is reasonably corroborative and that a
likelihood of only 1 in 100 can strongly influence the triers of fact.
Thus, although conventional serology can exclude a person, it can also
include many members of a population group as the possible origin of a blood,
saliva, or seminal fluid stain. Conventional serology is further limited, in that
analysis of mixed-fluid stains in which two or more contributors are involved can
mask an individual donor. Similarly, only 75-80% of the population are secretors
(exhibit their ABO blood type in their other physiological fluids). Thus, the
combination of those factors severely limits the power of conventional forensic
serological examinations as an individual identifier. Results of serological
analysis also are more subjective and can give rise to differing conclusions when
interpreted by equally qualified scientists.
Hair evidence is often encountered in sexual assault and other violent
crimes. It is valuable as exculpatory evidence and can be informative as to
identity, but it lacks specificity. Although hair examiners can associate a hair with
racial characteristics and body source (trunk, head, or pubic area) the variations
among hairs on a given person make definitive association of a single hair with an
individual problematic. The microscopic comparison of hairs is also subjective
and can lead to differences of opinion among equally qualified experts. With the
advent of DNA technology, especially PCR amplification techniques, the use of
hair as an individual identifier will become more common.
Some other forms of individual identification are available to forensic
scientists, but have very limited application. For example, examination of teeth is
useful in identifying deceased persons or bite marks.
Providing scientific evidence that DNA technology is at least as reliable as
other forensic methods and is therefore more likely to result in definitive
identification or exclusion of persons suspected or accused of a crime satisfies
both the ethical concerns about individual rights and the conditions of an ethical
analysis based on weighing good and bad consequences. However, additional
assurances are required for the risk-benefit ratio to be favorable in each case in
which the technology is used. Therefore, a critical step in accepting the use of
DNA technology in criminal trials is establishing safeguards and seeking to
prevent abuses.
uses of DNA technology that would constitute serious intrusions into the privacy
of ordinary citizens requires the setting of guidelines that separate proper use from
misuse of the technology.
The release of DNA information on a criminal population for purposes other
than law enforcement also constitutes misuse. Employers and insurance
companies will certainly have an interest in DNA information on potential
employees or customers.1,8,9 Biomedical and behavioral scientists are likely to
want to screen felon databanks and develop new databanks to study various
characteristics of convicted offenders. Legal sanctions should be established to
deter the unauthorized dissemination or procurement of DNA information that
has been obtained for forensic purposes.
EXPECTATIONS
The introduction of a powerful new technology is likely to set up
unwarranted or unrealistic expectations. Various expectations regarding DNA
typing technology are likely to be raised in the minds of jurors and others in the
forensic setting10 (see Chapter 6).
For example, public perception of the accuracy and efficacy of DNA typing
might well put pressure on prosecutors to obtain DNA evidence whenever
appropriate samples are available. As the use of the technology becomes widely
publicized, juries will come to expect it, just as they now expect fingerprint
evidence, surveillance photographs, and audio and visual eavesdropping.
Moreover, prosecutors will not want to give defense attorneys the opportunity to
ask on summation, "If my client was the perpetrator, where is the DNA
evidence?"
Once a prosecutor produces DNA evidence, the defense will be under great
pressure to undermine it through the use of reports and experts, because of an
assumption that the jury would interpret a failure to call a defense expert as an
admission that the DNA evidence is persuasive. Mere cross examination by a
defense attorney inexperienced in the science of DNA testing will not be
sufficient.
Two aspects of DNA typing technology contribute to the likelihood of its
raising inappropriate expectations in the minds of jurors. The first is the jury's
perception of an extraordinarily high probability of enabling a definitive
identification of a criminal suspect; the second is the scientific complexity of the
technology, which results in laypersons' inadequate understanding of its
capabilities and failings. Taken together, those two aspects can lead to the jury's
ignoring other evidence that it should be considering.
Expectations regarding the power of DNA typing can lead to overlooking or
ignoring sources of error or mistakes in applying the technology. For example,
jurors' focusing on the probability of correctly identifying a per
petrator might lead them to discount the possibility of laboratory error, whether it
stems from incompetence or carelessness of personnel, malfunctioning
equipment, or unavoidable mistakes.
The efficacy and accuracy of a new technology typically are initially
demonstrated by the most highly competent and knowledgeable practitioners. As
DNA typing becomes routine, the quality of laboratories and personnel using it
might decrease while still meeting the standards required for accreditation or
licensing. However, the expectations of judges and juries might remain high,
because of the superior knowledge and competence of the initiators of the
technology. Later gains in experience and improved typing could lead to an
increase in quality.
As large felon databanks are created, the forensic community could well
place more reliance on DNA evidence, and a possible consequence is the
underplaying of other forensic evidence. Unwarranted expectations about the
power of DNA technology might result in the exclusion of relevant evidence.
Both prosecutors and defense counsel are entitled to benefit from the power
of DNA evidence, but they should not oversell it. DNA evidence is not infallible;
all laboratory work is subject to error; and, given current population databanks
and laboratory protocols, a witness or prosecutor will seldom (if ever) be justified
in stating that the probability that a reported DNA match involves someone other
than the suspect is so low as to make that possibility entirely implausible. Claims
that treat DNA identifications as though they are as reliable as fingerprint
identifications in the typical rape or murder case are unjustified; until technology
and databanks improve, they are likely to remain so.
Presentations suggesting to a judge or jury that DNA typing is infallible can
rarely be justified and should generally be avoided. However, there might be
instances where a prosecutor could legitimately argue that the DNA evidence
conclusively proves that the defendant committed the offense. Two examples are
illustrative:
INTERNATIONAL EXCHANGE
The need for international cooperation in law enforcement calls for
appropriate scientific and technical exchange among nations. As in other areas of
science and technology, dissemination of information about DNA
typing and training programs for personnel likely to use the technology should be
encouraged. It is desirable that all nations that will collaborate in law-
enforcement activities have similar standards and practices, so efforts should be
furthered to exchange scientific knowledge and expertise regarding DNA
technology in forensic science.
SUMMARY OF RECOMMENDATIONS
REFERENCES
1. U.S. Congress, Office of Technology Assessment. Mapping our genes—the genome projects: how
big, how fast? Washington, D.C.: U.S. Government Printing Office, 1988.
2. U.S. Department or Health and Human Services and U.S. Department of Energy. Understanding
our genetic inheritance: the U.S. Human Genome Project, the first five years FY 1991-1995.
Springfield, Virginia: National Technical Information Service, 1990.
3. President's Commission for the Study of Ethical Problems in Medicine and Biomedical
and Behavioral Research. Screening and counseling for genetic conditions. Washington,
D.C.: U.S. Government Printing Office, 1983.
4. DNA Report of New York State Forensic Analysis Panel. Albany, New York, 1989.
5. U.S. Congress, Office of Technology Assessment. Genetic witness: forensic uses of DNA tests.
Chapters 3-5. OTA-BA-438. Washington, D.C.: U.S. Government Printing Office, 1990.
6. Beauchamp TL, Childress JF. Principles of biomedical ethics. Chapter 2. 3rd ed.New York: Oxford
University Press, 1989.
7. de Gorgey A. The advent of DNA databanks: implications for information privacy. Am J Law
Med. 16:381-398, 1990.
8. Macklin R. Mapping the human genome: problems of privacy and free choice. Pp. 107-114in:
Milunsky A, Annas G J, eds. Genetics and the law. Ill. New York: Plenum Press, 1984.
9. President's Commission for the Study of Ethical Problems in Medicine and Biomedical and
Behavioral Research. Splicing life. Washington, D.C.: U.S. Government Printing Office,
1982.
10. Annas GJ. DNA fingerprinting in the twilight zone. Hastings Center Rep. 20:35-37, March/April
1990.
11. National Association of Attorneys General. Resolution, adopted at winter meeting. December
10-13, 1989, Phoenix, Arizona.
Organizational Abbreviations
GLOSSARY 167
Glossary
GLOSSARY 168
GLOSSARY 169
Gametic the state at loci on different chromosomes when the allele at one locus in the
(phase) gamete varies independently of that at the other loci; in gametic (phase)
equilibrium disequilibrium, a specific allele at one locus is associated with an allele at
another locus on a different chromosome with a frequency greater than
expected by chance (see linkage disequilibrium)
Gel semisolid matrix (usually agarose or acrylamide) used in electrophoresis to
separate molecules
Gene the basic unit of heredity; a sequence of DNA nucleotides on a chromosome
Gene fre- the relative occurrence of a particular allele in a population
quency
Genetic drift random fluctuation in allele frequencies
Genome the total genetic makeup of an organism
Genotype the genetic makeup of an organism, as distinguished from its physical
appearance or phenotype
Guanine a purine base; one of the four nitrogen-containing molecules present in
nucleic acids DNA and RNA; designated by the letter G
Haploid having one set of chromosomes (compare diploid)
Hardy- equilibrium the condition, for a particular genetic locus and a particular
Weinberg population, with the following properties: allele frequencies at the locus are
constant in the population over time and there is no statistical correlation
between the two alleles possessed by individuals in the population; such a
condition is approached in large randomly mating populations in the absence
of selection, migration, and mutation
Heredity the transmission of characteristics from parent to offspring
Heterozygotea diploid organism that carries different alleles at one or more genetic loci on
its homologous chromosomes
Heterozy- having different alleles at a particular locus; for most forensic DNA probes,
gous the autoradiogram displays two bands if the person is heterozygous at the
locus
HLA see human leukocyte antigen
Homology similarity between two structures or functions indicative of a common
evolutionary origin
Homozygote a diploid organism that carries identical alleles at one or more genetic loci on
its homologous chromosomes
Homozygous having the same allele at a particular locus; for most forensic DNA probes,
the autoradiogram displays a single band if the person is homozygous at the
locus
Human protein-sugar structures on the surface of most cells, except blood cells, that
leukocyte differ among individuals and are
antigen
(HLA)
GLOSSARY 170
GLOSSARY 171
GLOSSARY 172
Serology the discipline concerned with the immunologic study of body fluids
Serum the liquid that separates from blood after coagulation
Sex chromo- chromosomes that are different in the two sexes and that are involved in sex
somes (x and determination
y chromo-
somes)
Sex-linked a genetic characteristic, such as color blindness, that is determined by a gene
characteris- on a sex chromosome and shows a different pattern of inheritance in males
tic and females; X-linked is a more specific term
Single-locus a DNA probe that detects genetic variation at only one site in the genome; an
probe autoradiogram that uses one single-locus probe usually displays one band in
homozygotes and two bands in heterozygotes
Somatic cells the differentiated cells that make up the body tissues of multicellular plants
and animals
Southern blot the nylon membrane to which DNA adheres after the process of Southern
blotting
Southern the technique for transferring DNA fragments that have been separated by
blotting electrophoresis from the gel to a nylon membrane
Standards criteria established for quality control and quality assurance; established or
known test reagents, such as molecular-weight standards
T single-letter designation of the pyrimidine base thymine; also used in
diagrams to represent a nucleotide containing thymine
Tandem multiple copies of an identical DNA sequence arranged in direct succession
repeats in a particular region of a chromosome
Taq poly- a DNA polymerase used to form double-stranded DNA from nucleotides and
merase a single-stranded DNA template in the PCR technique
Thymine a pyrimidine base; one of the four nitrogen-containing molecules present in
nucleic acids DNA and RNA; designated by the letter T
Uracil a pyrimidine in RNA
Variable repeating units of a DNA sequence for which the number varies between
number of individuals
tandem re-
peats (VN-
TR)
VNTR variable number of tandem repeats
Zygote diploid cell that results from the fusion of male and female gametes
PARTICIPANTS 177
Participants
PARTICIPANTS 178
INDEX 179
Index
INDEX 180
D
C
Databanks, 15, 104
Caldwell v. State, 140-141 access to, 18-19, 93-94, 123
California Association of Crime Labora- allele frequency, 10, 74-75, 85-86, 91,
tory Directors, 88, 89 122, 124, 125
Carryover contamination, 58-59, 66-67 criminal profile, 17-20, 86-87, 111-129,
Ceiling frequencies, 13, 82-83, 84, 90-93 142-143, 155, 159, 160
Ceiling principle, 13, 14, 25, 82-85, 92, and legal proceedings, 135, 139, 140, 141
93, 158 privacy issues, 14, 17-20, 24-25, 32,
Cell repositories, 14, 90-91 86-87, 113-116, 121-122, 155 -156,
Census categories, 12 158-160
Centre d'Etude du Polymorphisme samples, 112-113, 116-123, 125-126, 143
Humaine, 91 sequence variability, 44
Certification of personnel, 16, 98, 100, see also Automated fingerprint identifi-
145, 153 cation systems;
Chain of custody, 131, 142 Cell repositories;
Civil cases, 20, 131-132 Computer technology;
Cleavage sites, 38, 54 Networks
Clinical Laboratory Improvement Act, 17, Defense counsels, 146, 147, 160, 161
102, 107 Defense laboratories, 123
Cobey v. State, 136 Demography,
College of American Pathologists, see Ethnic groups; Subpopulations
101-102, 107 Dental records, 125, 158
Commonwealth v. Curnin, 141 Department of Defense, 125-126
Computer technology, 44, 48-49, 60, 61; Department of Health and Human Ser-
see also Networks vices, 17, 107, 108
for databanks, 18, 113-114, 117, 123, 125 see also National Institutes of Health
Confidence limits, 9, 14-15, 75-76, 92 Department of Justice, 17, 107, 108
Confidentiality, see also Bureau of Justice Assistance;
see Privacy issues Federal Bureau of
Consequentialist ethical analysis, 24-25,
154, 156-157, 158
Constitutional issues, 20, 131, 143, 153
Contamination, 20, 52, 55, 131, 134
in PCR procedures, 65-67, 68, 70
in RFLP procedures, 58-59
Control primers, 70
Controls, in testing, 55, 57-58
for laboratory proficiency, 62-63
Convicted-felon DNA databanks, 17-20,
86-87, 111-129, 142-143, 155, 159,
160, 161
Cooperative services, 153
Copyright National Academy of Sciences. All rights reserved.
DNA Technology in Forensic Science
INDEX 181
INDEX 182
G Information networks,
see Networks
Genetic disorders, In limine hearings, 144, 145
see Disease testing Inspections of laboratories, 106
Genetic drift, 82, 83-84
Insurance companies, 114, 160
Genetic mapping, 91, 152, 155 Internal molecular-weight standards, 60-61
Genetics Society of America, 72
International technology exchange, 25-26,
Genomes, 2-3, 9, 34 162-163
Genotypes, 32
Interstate commerce, and laboratory
Governor's Select Committee on DNA
accreditation, 108
Typing, New York, 103
Isotopic labeling, 37-38
Grandparentage86
Guanine, 2, 33
Guidelines for a Quality Assurance Pro- J
gram for DNA RFLP Analysis, 98-99
Jeffreys, Alex, 4, 40, 43-44
Judicial notice, 133-134, 145
H Jury presentations, 14, 20-23, 25, 89, 132,
136-137, 146, 147, 160 -161
Hair, 28, 120, 158 Juveniles, 148
Haploid genome, 3
Hardy-Weinberg ratios, 4-5, 11, 45, 78,
81-81, 91, 139, 141 K
Helpfulness standard of evidence, 21-23,
132, 136-137 Kelly v. Texas, 140
Heterogeneity of population, 82, 92 Kidnapping, 118, 121
Heterozygosity, 4, 36 Kits, for PCR analysis, 68, 69
frequency calculation of, 11, 45, 48, 78
and PCR analysis, 68, 69
L
and RFLP analysis, 38-40, 58
HLA (human leukocyte antigens), 10, 27, Laboratories, 8, 15, 18-19, 20, 28, 53-55,
32, 44, 68, 70, 114 97, 101, 105, 112, 116, 161, 162
Homicide, 19, 112, 118-120, 143 error rates, 14, 15, 88-89, 94
Homogeneity of populations, 12, 14, 84 and legal proceedings, 133, 135,
Homozygosity, 4, 36, 87 140-141, 145, 148
frequency calculation of, 11, 45, 48, 78, licensing, 100-101
79 protocols, 8, 53-55, 105, 112, 133, 135,
and PCR analysis, 68 140-141, 145, 148
and RFLP analysis, 38-40, 58 regulation of, 15, 16-17, 97-108, 162
Human genome project, see also Match criteria;
see Genetic mapping Proficiency testing; Technicians
Human leukocyte antigens, Laboratory Accreditation Board, Ameri-
see HLA can Society of Crime Laboratory
Hybridization, 37, 38, 40, 42, 48 Directors, 17, 102-103, 105-106, 107
and band analysis, 57-59, 60 Latent fingerprints, 17-18, 111-113
reverse dot, 42, 67, 68 Leakage between samples, 59
Hypervariable systems, see VNTR Licensing of laboratories, 100-101
I
Immigrants, 84, 155
Immortalized cell lines, 14, 90-91
Individual identification, 9, 28, 43-44,
74-75, 89-90, 113
from fingerprints, 29-31, 113
INDEX 183
Likelihood ratios, 85 O
Linkage equilibrium, 36, 78, 81, 91
Oligonucleotides, 40, 42
Organizations,
M see Professional organizations
INDEX 184
INDEX 185
T
Technical Working Group on DNA Analy-
sis and Methods, 16, 98-99, 104,
106, 116-117
Technicians, 15, 16, 17, 25-26, 66, 97, 100
Technological advancement, 15, 55,
70-71, 143-145, 153
Technology transfer, 71
international, 25-26, 162-163
Thermocyclers, 64
Thymine, 2, 33
Trade secrets, 93-94, 148, 162
Twins, 29, 44
Training, 15, 16, 17, 25-26, 96, 104, 108,
125, 153
U
UNC (uracil N-glycolase enzyme), 67
Unidentified bodies, 121, 125, 126, 158
United States v. Jakobetz, 138
United States v. Porter, 138
United States v. Yee, 138