You are on page 1of 10

Received: 10 December 2019 Revised: 28 February 2020 Accepted: 10 March 2020

DOI: 10.1002/wfs2.1373

PRIMER

An introduction to postmortem interval estimation in


medicolegal death investigations

Lerah Sutton | Jason Byrd

Department of Pathology, Immunology,


and Laboratory Medicine, University of
Abstract
Florida, College of Medicine, Gainesville, Throughout the history of forensic medicine, the postmortem interval has been
Florida
one of the most commonly and thoroughly investigated problems. The impor-
Correspondence tance of an accurate PMI in the context of a medicolegal death investigation can-
Jason Byrd, Department of Pathology, not be understated due to its utility and application toward investigative
Immunology, and Laboratory Medicine,
determinations including inclusion or exclusion of suspects, determination of
University of Florida, College of Medicine,
Gainesville, FL. time of assault versus time of death, and preliminary victim identification. How-
Email: jhbyrd@ufl.edu ever, despite its importance, the question of postmortem interval estimation is
often answered with a low degree of accuracy as compared to the rates of cer-
tainty within other forensic disciplines. While there are various methods that
may be utilized for answering the time of death question including both scientific
and investigative techniques, the variables that affect the application of these
methods to death investigations are abundant. As a result, there are numerous
limitations associated with time of death determination in a medicolegal death
investigation and the estimation of the postmortem interval must be done with
caution and deference to the many variables that affect its accuracy.

This article is categorized under:


Forensic Anthropology > Time Since Death Estimation
Forensic Medicine > Death Scene Investigation
Crime Scene Investigation > Education and Formation

KEYWORDS
death investigation, decomposition, forensic medicine, postmortem interval, time since death

1 | INTRODUCTION

Within the forensic sciences, medicolegal death investigations account for a large portion of investigative efforts.
Medical examiners, coroners, crime scene investigators, death investigators, autopsy technicians, research scientists,
and subject matter experts all coalesce under the umbrella of forensic medicine—the application of medical science to
the investigation of crimes and deaths. These joint efforts contribute toward successful death investigations involving
elements including positive identification of the victim, identification of a suspect, verification of suspect alibis, and the
analysis and processing of physical evidence. In many cases, the evidence within these investigations involves a
deceased and decomposing human body and the question of time since death is raised.
The interest in establishing time since death is not a new concept. Rather, it has been investigated and studied for
hundreds of years. The earliest documented study of the postmortem interval comes from 13th century China in a text
entitled “The Washing Away of Wrongs.” Often associated with the first documented murder case solved by forensic

WIREs Forensic Sci. 2020;e1373. wires.wiley.com/forensicsci © 2020 Wiley Periodicals LLC 1 of 10


https://doi.org/10.1002/wfs2.1373
2 of 10 SUTTON AND BYRD

entomology, this text—authored by an investigator named Sung Tz'u of the Southern Song Dynasty in 1237 A.D.—
describes the relationship between postmortem changes in the body condition and assessment of the time since death
(Tz'u & McKnight, 1981) (Figure 1). Although rudimentary attempts were made between this earliest documented asso-
ciation and the modern method used to estimate time since death, it was not until the mid to late 1800s that controlled
scientific studies were conducted in an attempt to better determine the postmortem interval, many of which focused
solely on body temperature (Knight & Madea, 2016). However, even into the 21st century despite decades of diligent
research assessing the innumerable variables that affect the rates of decomposition as associated with time since death,
the accuracy with which scientists and investigators can measure the postmortem interval remains regrettably low
(Henssge & Knight, 2002).

2 | THE ROLE O F P OSTMORTEM INTERVAL ESTIMATIONS IN


MEDICOLEGAL DEATH INVESTIGATIONS

The longer the interval of time between death and examination of the body, the more difficult postmortem interval esti-
mation becomes. Likewise, the longer the interval between death and examination of the body, the more important

FIGURE 1 An excerpt from “The Washing Away of Wrongs” wherein investigator Sung Tz'u describes and labels the human skeleton
for use in death investigations. Photo courtesy of Wikimedia Commons
SUTTON AND BYRD 3 of 10

postmortem interval estimation becomes as it is a crucial precursor to a thorough case investigation. In some cases,
there may be a period of survivability between the initial injury and the actual death itself. Understanding the relation-
ship between such events and being able to clearly discern time since death is crucial for a medicolegal death investiga-
tion. This necessitates a twofold approach to the investigative practices surrounding postmortem interval estimation.
First, antemortem and perimortem changes and injuries such as wound pattern analysis, infection, or other injury must
be assessed to determine what, if any, effect this may have on the period of survivability as it relates to the time since
death. And second, the postmortem changes must be assessed to provide a preliminary estimation of the time since
death (Madea & Henssge, 2016). This information is vital to law enforcement as they cannot effectively begin to investi-
gate these cases without a reliable idea of when the victim died. This death determination—and its accuracy—plays a
role in implicating or excluding suspects, corroborating or refuting suspect alibis, and uncovering the specific circum-
stances and chain of events that led to the death of the individual in question. The importance of the role that an accu-
rate and reliable postmortem interval plays in a medicolegal death investigation cannot be understated. However, it is
crucial to note and understand the significance of the word interval in this case. While the goal would, of course, be to
be able to pinpoint the exact time since death down to the minute, that is virtually impossible. Even if another individ-
ual is present at the moment of death to witness the occurrence, there remains a small margin of error. In forensic
cases, however, the margin of error can be vast and it falls to the investigators to use all available evidence and methods
to provide their best educated guess at the postmortem interval estimation, keeping in mind that the key distinction
that the interval provided is, indeed, an estimation (Camps, 1959).

3 | CLASSIFICATIONS OF EVIDENCE TYPES A ND METHODS

When assessing the body of a deceased individual for postmortem interval analysis, the body itself is both the evidence
and the method used for assessment. Thus, a combination of both numerous types of evidence and numerous methods
of estimation should be jointly evaluated in order to provide a more accurate time since death determination. Within
classifications of evidence and methodological assessments, both scientific and investigative techniques should be uti-
lized. That is, techniques that can be quantified and objectively evaluated should be supplemented with more circum-
stantial evidence that gains meaning from context and association with other evidence and analysis. Because every item
of evidence and every current method for postmortem interval estimation hold limitations to their accuracy and appli-
cation, it is prudent to take a broader approach to the classification of evidence and methods and then utilize each clas-
sification type together for a more reliable time of death determination.

3.1 | Corporal, environmental/associated, and anamnestic evidence

When assessing the various elements of a medicolegal death investigation that must come together in order to effec-
tively and reasonably assess the postmortem interval, no singular piece of evidence used can answer that question fully
and independently. Rather, it is the summation of numerous classifications of evidence working together that provides
a more comprehensive assessment of postmortem interval. Generally, there are three primary classifications of types of
evidence in a death investigation: corporal evidence, environmental/associated evidence, and anamnestic evidence.
Corporal evidence is defined as evidence that presents itself anatomically within the body and is generally associated
with the scientific methods of postmortem interval estimation. Examples of this include postmortem changes such as
livor mortis or rigor mortis. Environmental or associated evidence is defined as evidence that is present within the
vicinity of the body and is also generally associated with scientific methods of time since death determination. Exam-
ples of this include the utility of forensic entomology—or more specifically, the insects that colonize decomposing
remains—for postmortem interval estimation or other physical evidence associated with the scene and body that may
be processed and analyzed. Anamnestic evidence is defined as evidence cleaned from deviations or abnormalities
assessed within the deceased's ordinary habits, movements, and day-to-day activities. This type of evidence is generally
more investigative, rather than scientific, as it is difficult to quantify statistically or numerically in death investigations.
Examples of anamnestic evidence include the failure of a normally routine-oriented and reliable individual to show up
for work or the sudden cessation of credit card or phone use. Corporal and environmental/associated evidence may pro-
vide more statistically significant and accurate information for estimation of the postmortem interval as they rely on
4 of 10 SUTTON AND BYRD

scientific methods, but when those methods are further supplemented with the anamnestic evidence gleaned through
investigative methods, the reliability of a postmortem interval estimation is generally much greater.

3.2 | Rate and concurrence methods

In the same way that no singular piece of evidence should be used in isolation to solve a case, but instead multiple evi-
dentiary items should be used together, no singular method should be employed in isolation to answer questions related
to postmortem interval estimation. Rather, as many methods as possible—provided they maintain relevance to the
case—should be utilized in tandem to help bolster the accuracy and reliability of the time since death determination.
There are two primary classification groups under which all methods of postmortem interval estimation will fall: rate
methods and concurrence methods. Rate methods are defined as those which measure the change produced by a pro-
cess which takes place at a known rate. The process being measured must have either been initiated or stopped by the
death event in question. Most rate methods are considered scientific in nature as they can be quantitatively measured
and evaluated. Examples of rate methods include postmortem decompositional changes such as the onset and dissipa-
tion of rigor mortis and the visibility and fixation of livor mortis. Concurrence methods are defined as those which com-
pare the occurrence of events which took place at known times with the occurrence of the death event in question.
Concurrence methods are generally less scientific and more investigative as they relate circumstances and events to the
death even itself. An example of a concurrence method is the shattering of a watch in the context of a fight where the
victim later succumbed to multiple blunt traumatic injuries. The time at which the watch was stopped could be corre-
lated to the time of injury which eventually led to the death of that individual.

4 | COMMON METHODS O F P OSTMORTEM INTERVAL ESTIMATION

Nearly everything associated with the body of a deceased individual plays a role in the process of decomposition, thus
these same items affect the estimation of the postmortem interval. Factors relating to the individual themselves includ-
ing age, sex, body weight, or antemortem health conditions can affect the rate at which their body decomposes. Like-
wise, environmental factors and factors relating to the body deposition and disposition including ambient temperature,
precipitation, clothing, wrapping, concealment, burial, soil composition, and insect activity can also affect the rate of
decomposition. Other factors potentially relating to the cause of death such as infection, disease, drug use, or trauma
play a role as well. Given that each of these factors individually act as variables that can dramatically alter the rate of
decomposition, they also affect the ability to accurately estimate the postmortem interval. Much research has been con-
ducted, both empirically and anecdotally, throughout recent history in an attempt to control and quantity the effects
each of these variables have on the rate of decomposition and to use them to be able to estimate the time since death.
To fully summarize and describe the primary methods associated with each variable that affects postmortem interval
estimation would be an exhaustive task not to be undertaken within the confines of this article as the purpose of this
article is to provide the reader with a basic education on the most historically significant and widely used methods, but
a resource that does an excellent job of summarizing many of the recent methods of postmortem interval estimation is
a text entitled Estimation of the Time Since Death edited by Burkhard Madea (Madea, 2016). While there is a great deal
of literature currently being published relating to new and novel methods of postmortem interval estimation including
degradation of RNA/DNA, continuous indices such as total body score being developed for application to the decompo-
sition process, and exploration into the microbiome, they fall outside the scope of this article. However, it would be irre-
sponsible to discuss the importance of the postmortem interval estimation without engaging in at least a foundational
discussion on the most common methods actively utilized in forensic casework to help determine time since death.
These common methods include assessing the postmortem changes associated with artifacts of human decomposition,
temperature-based methods, and forensic entomology.

4.1 | Artifacts of decomposition

Estimation of the postmortem interval is especially important in decomposition cases where the circumstances sur-
round the death event are not immediately clear. Because a longer interval between time of death and time of discovery
SUTTON AND BYRD 5 of 10

makes the postmortem interval estimation process far more difficult, it is important to understand both the available
options and limitations of assessing postmortem interval when there are visible postmortem changes to the body. The
body decomposes through simultaneous processes of autolysis and putrefaction. Autolysis is defined as the self-
dissolution of the body by the endogenous digestive enzymes which break down the carbohydrates and proteins that
are released from disintegrating cells. More simply put, it is the destruction of cells or tissues by their own enzymes
(Madea, Kernbach-Wighton, Henssge, Doberentz, & Musshoff, 2016). Putrefaction is the breakdown of tissues into
gases, liquids, and salts due to the bacteria and micro-organisms already present within the body. This process causes
the offensive odor that is generally associated with decomposition due to the release of mercaptans in the decomposi-
tion process. Although both autolysis and putrefaction occur simultaneously to perpetuate decomposition, it is gener-
ally the putrefactive changes that are recognized and assessed in discrete stages when estimating postmortem interval
in decomposition cases (Lew & Matshes, 2005). There are five stages that are most commonly associated with the
decomposition process: fresh, bloat/early decay, active decay, advanced decomposition, and skeletonization. It is only
within the first two and perhaps the very early stages of the third that postmortem interval can even remotely reliably
be estimated from the artifacts of decomposition and postmortem changes observed in the body (Hayman &
Oxenham, 2016b).
Two of the most common artifacts of decomposition associated with postmortem interval estimation are livor mortis
and rigor mortis (Figure 2). Livor mortis, also known as lividity or hypostasis, is defined as the postmortem discolor-
ation of the body due to settling of the blood in dependent portions of the body, except in areas exposed to direct pres-
sure. It can onset as quickly as 30 min postmortem but is generally readily visible approximately 2 hr after death. It also
shifts positionally until it becomes fixed at approximately 4–6 hr postmortem (Madea, 2016b). Although it remains visi-
ble into the active stage of decomposition when the body systematically begins to discolor, its utility for postmortem
interval estimation is only limited to the very early stages of decomposition. It is, however, excellent for assessing the
position of the body at the time of death and determining whether the body may have been moved after death by evalu-
ating the lividity versus the position of the body at the time of discovery. Rigor mortis is defined as the postmortem stiff-
ening of the body due to the actin–myosin bond in the absence of ATP that is depleted after death. It is important to
note that it is not due to muscle contraction, although the visible results may appear similar to the untrained eye. The
onset of rigor mortis can be variable, generally coming discernible at approximately 4–6 hr after death in the smaller
muscle groups, reaching its maximum at approximately 12–24 hr after death wherein all muscle groups are stiff, and
then dissipating 24–46 hr after death in the same fashion as it occurred. However, this is not always a reliable indicator
of time since death as the onset of rigor mortis can—in rare cases—be immediate. It can also be affected by numerous
outside factors including temperature (heat will accelerate the process and cold will decelerate), perimortem illness,
activity before death, physical conditions of body placement and discovery, and the age of the individual as rigor mortis
is generally poorly formed in the very young and very old (Krompecher, 2016). While it does offer a longer potential
time window for postmortem interval estimation as compared to lividity, due to its extreme variability it is not generally
considered a reliable indicator of postmortem interval.
Another commonly explored method for postmortem interval estimation within the scope of decompositional
changes is that of changes in concentration of electrolytes in the vitreous humor, specifically potassium. It is well

F I G U R E 2 A pig carcass exemplifying two of the common


artifacts of decomposition that can be used to estimate time since
death in the earlier stages of decomposition—lividity and rigor
mortis. Photo courtesy of Wikimedia Commons
6 of 10 SUTTON AND BYRD

established in the scientific literature that there is a relationship between time since death and the increase in potas-
sium content in vitreous humor (Coe, 1989; Harper, 1989; Khorasgani, Arzi, Salahcheh, & Latifi, 2009; Lange,
Swearer, & Sturner, 1994; Stephens & Richards, 1987; Sturner, 1963). However, there are myriad factors which may
affect the utility and interpretation of PMI based on rise in potassium content in vitreous humor. These include dura-
tion of the death event itself, the composition of the vitreous humor sample, the sampling method (i.e., placement of
needle, depth of insertion, and force used for extraction), ambient temperatures during the decompositional phase, age,
and drug or alcohol use at the time of death (Harper, 1989; Madea et al., 2016). While there is a generally well-
documented linear relationship in the increase in potassium content of the vitreous humor after death as best calcu-
lated by Sturner's equation of PMI = 7.14 × K+ concentration − 39.1, it is also generally agreed upon that it holds lim-
ited value for PMI estimation within the first 24 hr after death (Coe, 1989; Sturner, 1963).

4.2 | Temperature-based methods

One of the most significant variables that affects postmortem interval estimation in decomposition cases is temperature.
It is a well-known fact that higher temperatures increase the rate of decomposition and lower temperatures slow down
the rate. But, given the intervariability of daily temperature fluctuations over the course of the decomposition process
of an individual, it is often very difficult to accurately determine the extent to which temperature altered the rate of
decomposition. Numerous studies have been undertaken in attempts to understand, control, and quantify the effects of
temperature on the process of decomposition as it relates to postmortem interval estimation. The premise of these stud-
ies is a phenomenon called algor mortis. Although it is generally and colloquially defined as the postmortem cooling of
the body, that is not entirely accurate. More correctly, it is the propensity of a body to reach equilibrium with the ambi-
ent temperature after death. Some of the earlier studies on algor mortis as a tool for estimating postmortem interval
established the “rule of thumb” principle that a body will drop an average of 1.5 F per hour after death, after the initial
plateau phase (Hayman & Oxenham, 2016a). The difficulty with this method—and others that seek to create a single
predictive factor for the rate at which a body changes temperature after death—is that it does not take into account the
numerous variables that affect temperature fluctuation after death. Rather, it assumes a standard and controlled tem-
perature and airflow environment which is not realistic when applied to death investigations (Knight, 1988). In an
effort to better account for temperature variation and other factors that affect postmortem temperature change, exten-
sive mathematic modeling was undertaken and produced a series of nomograms—that is, diagrams that represent the
relationship between multiple variables that can be quantified mathematically (Figure 3). These nomograms did
improve the accuracy of postmortem interval estimation from temperature readings, but they were also rife with limita-
tions, the primary ones being the ways in which data must be obtained. The nomogram data was created and tested
using rectal and liver temperatures which are acceptable in a laboratory environment, but greatly discouraged in foren-
sic casework. Inserting a temperature probe into either the liver or the rectum before a complete forensic autopsy has
been conducted may alter or destroy evidence that may have provided critical to the case. But, the nomograms would
not be effective unless the temperature readings were taken immediately upon body discovery. Some efforts have been
made to use other sources of temperature data including infrared thermometers, but the additional limitations of cor-
rective factors, airflow, clothing, body size, perimortem infection, and others that affect temperature remain difficult to
control (Henssge, Rutty, Hubig, Muggenthaler, & Mall, 2016). Thus, using postmortem temperature readings alone as a
determination for postmortem interval is not considered a reliable method due to the extreme difficulty in identifying,
determining, predicting, and controlling for all the variables that affect it. Furthermore, even the most promising math-
ematical modeling in terms of accuracy loses its effectiveness beyond the 24-hr range postmortem, which leaves a major
deficit in the types of cases to which it can be applied.

4.3 | Forensic entomology

One of the most statistically accurate methods for estimating the postmortem interval in decomposition cases
(i.e., cases with an approximate time since death of 3 days or greater) is the use of forensic entomology. By applying
knowledge of insects and their arthropod relatives to legal cases, entomologists can aid investigators in establishing a
time since death (Byrd & Castner, 2010; Smith, 1986) (Figure 4). This is generally done in one of two ways. The first is
through the use of insect succession wherein postmortem interval is assessed by evaluating the presence or absence of
SUTTON AND BYRD 7 of 10

F I G U R E 3 An example of one
of Henssge's nomograms used to
predict postmortem interval based
on the relationship between
temperature variables. Photo
courtesy of Wikimedia Commons

certain insect species that are known to be associated with specific decompositional stages, which can be recognized
when describing the continuum of gradual changes that occurs with decomposition (Boulton & Lake, 1988; Schoenly &
Reid, 1987; Schoenly & Reid, 1989). This is the more straightforward method of the two, but presents limitations in that
there is significant geographic variation among insect species and there is limited species variation in the earlier stages
of the postmortem interval, thus limiting the preferred application of this method to the later stages of decomposition
(Goff, 1993; Payne, 1965). The second method is more statistically accurate and is what most associate with forensic
entomology in application to medicolegal death investigations. This method uses the temperature-dependent develop-
ment of insects to help determine a portion of the postmortem interval. When it is warmer, insects will develop faster;
likewise, when it is colder, insects will develop slower. Taking this development information—which must be species
specific—and standardizing the temperatures against time, entomologists can help determine the time since death
(Catts, 1992; Greenberg, 1991; Tomberlin, Mohr, Benbow, Tarone, & Vanlaerhoven, 2011).
With many death scene investigations, the PMI is a common and essential question (Amendt et al., 2007). The infor-
mation an entomologist can supply is that insect development on a corpse can be used to determine how long the
deceased has been dead. This period of time is known as the PMI (Catts, 1990). However, it can be argued that such
8 of 10 SUTTON AND BYRD

F I G U R E 4 Blowflies are some of the most common insects


associated with decomposition cases, the development of which can
be used to aid in postmortem interval estimation. Photo courtesy of
Dr J.H. Byrd

calculations should be referenced as the period of insect activity (PIA) or the postcolonization interval (PCI). These
terms have been further refined to the period of time between initial detection of the corpse by an insect (PIA), and the
PCI that is exclusive of periods prior to colonization (Tomberlin et al., 2011). Therefore, entomological estimations
related to a PMI can envelope a significant portion of the time since death, or only be partial estimation. It is important
to note that many practicing forensic entomologists discourage the direct application of the term “time since death” to
entomological evidence analysis as there are many factors which can slow down or entirely prevent insect colonization
of decomposing remains including freezing temperatures, heavy rains, tight wrapping/cover, concealment, and burial.
Rather, the concept is better explained as a minimum time since colonization, or a minimum postmortem interval. The
distinction of the word minimum with relation to this concept is integral to addressing the limitations of this method.
Because the time period during which the individual in question was deceased but not yet colonized by insects can vary
greatly and generally cannot be effectively estimated by forensic entomologists, a minimum time interval is given.
There is certainly a need for standardized terminology for specific situations in forensic entomology analysis (Michaud,
Moreau, & Schoenly, 2014).

5 | L I M I T A T I O N S O F P O S T M O R T E M I N TERVAL ESTIMATION METHODS


AND A PPLICATION CONS IDER ATIONS

While some of the discipline-specific limitations have already been discussed in previous sections, there are general
considerations one must understand with regard to the limitations of postmortem interval estimation and the ability to
realistically apply certain methods to medicolegal death investigations. Even with the advances in scientific methods
and techniques that aim to control and quantify the variables that affect postmortem interval estimation, there will
always be an inherent error rate associated with our inability to fully understand, comprehend, and control every vari-
able in a death investigation case. One of the most common and pervasive limitations to postmortem interval estima-
tion methods comes from the subjective nature of the methods often used to make this estimation. Much of the data
collection used outside of temperature data is subjective; that is, it can be altered by the perceptions, experiences, and
confidences of the observer. Very few studies have been conducted on the reliability of observer data as it directly
relates to postmortem interval estimation, but one of the few published studies demonstrated that out of 110 cases with
a known time of death, experienced forensic pathologists were only able to correctly estimate the time since death in
11 of the cases—only approximately a 10% success rate. In 57 of the cases the time since death was underestimated and
in 32 cases the time since death was overestimated (James & Knight, 1965). Investigatively, it is safer to underestimate
the postmortem interval than to overestimate given the consideration of the minimum postmortem interval. However,
the underlying commonality between nearly all studies assessing postmortem interval is that even in the most con-
trolled laboratory settings, the postmortem interval is very difficult to accurately predict. Given the limitations of accu-
rate time since death determination that are present in the controlled environment of laboratory a setting, removing
the control barriers and introducing perhaps inestimable additional variables that prove consistently difficult to quan-
tify, the task of determining of an accurate postmortem interval in a forensic case becomes that much more challenging
(Madea, 2016). At best, the goal is to minimize the effects of estimation errors, account for as many variables as possible
SUTTON AND BYRD 9 of 10

by using a combination of multiple methods into the estimation process and make a cautious estimation of the post-
mortem interval with all of these factors in mind (Camps, 1959).

6 | C ON C L U S I ON

Estimation of the postmortem interval within the context of a medicolegal death investigation is simultaneously one of
the most common and most difficult to answer questions within forensic medicine. It plays a role in nearly every aspect
of a death investigation including implicating or excluding suspects, determining the time of assault versus the time of
death, whether there was a period of survivability or neglect throughout the chain of events leading to the death of an
individual, and may even be useful in determining the identity of the deceased. Because of its varied applications, a
great many methods have been established and implemented in an effort to better determine postmortem interval and
how evidence associated with the death event can be used to supplement a time of death estimation. However, even
with scientific and investigative methods that individually address nearly every measurable variable that can be experi-
mentally controlled, these variables do not exist in isolation in real-world investigations. Because of the way that vari-
ables such as environmental temperature, antemortem health, body size, clothing, concealment, insect activity, and
many others each independently affect the postmortem interval, combining them in the ways that are unique to each
death investigation makes an accurate prediction or estimation of the time since death nearly impossible to precisely
establish. Rather, taking a cue from forensic entomology and establishing a minimum interval of time since death—
while giving consideration to un-assessible factors that could elongate that timeframe—provides a far more prudent
and reasonable approach to answering this essential question.

CONFLICT OF INTEREST
The authors have declared no conflicts of interest for this article.

A U T H O R C ON T R I B U T I O NS
Lerah Sutton: Conceptualization; writing-original draft; writing-review and editing. Jason Byrd: Conceptualization;
writing-original draft; writing-review and editing.

ORCID
Lerah Sutton https://orcid.org/0000-0002-2082-9436
Jason Byrd https://orcid.org/0000-0001-7196-5270

F u r t h e r Re a d i n g
Byrd, J. H., & Tomberlin, J. K. (2020). Forensic entomology: The utility of arthropods in legal investigations (3rd ed.). Boca Raton: CRC Pres-
s/Taylor & Francis.

R EF E RE N C E S
Amendt, J., Campobasso, C.P., Gaudry, E. Reiter, C., LeBlanc H.N., Hall, M.J. 2007. Best Practice in forensic entomology: standards and
guidelines. Int J Legal Med. Mar; 121(2): 90–104.
Boulton, A. J., & Lake, P. S. (1988). Dynamics of heterotropic succession in carrion arthropod assemblages: A comment on Schoenly and
Redi (1987). Oecologia, 76, 477–480.
Byrd, J.H. & Castner, J.L. (Eds) 2010. Forensic Entomology: The Utility of Arthropods in Legal Investigations. CRC Press. Boca Raton. 681 pp.
Camps, F. (1959). Establishment of the time since death: A critical assessment. Journal of Forensic Sciences, 4, 73–76.
Catts, E.P. and Haskell, N.H. 1990. Entomology and Death: A Procedural Guide. Joyce's Print Shop Inc. Clemson. 1–182.
Catts, E. (1992). Forensic entomology in criminal investigations. Annual Review of Entomology, 37(1), 253–272.
Coe, J. I. (1989). Vitreous potassium as a measure of the postmortem interval: An historical review and critical evaluation. Forensic Science
International, 42(3), 201–213.
Goff, M. L. (1993). Estimation of postmortem interval using arthropod development and successional patterns. Forensic Science Review, 5(2),
81–94.
Greenberg, B. (1991). Flies as Forensic Indicators. Journal of Medical Entomology, 28(5), 565–577.
Harper, D. (1989). A comparative study of the microbiological contamination of postmortem blood and vitreous humour samples taken for
ethanol determination. Forensic Science International, 43(1), 37–44.
10 of 10 SUTTON AND BYRD

Hayman, J., & Oxenham, M. (2016a). Algor mortis and temperature-based methods of estimating the time since death. In Human body
decomposition (pp. 13–52).
Hayman, J., & Oxenham, M. (2016b). Human body decomposition. London, England: Elsevier Academic Press.
Henssge, C., & Knight, B. (2002). The estimation of the time since death in the early postmortem period. London, England: Arnold.
Henssge, C., Rutty, G., Hubig, M., Muggenthaler, H., & Mall, G. (2016). Postmortem body cooling and temperature-based methods. In Esti-
mation of the time since death (3rd ed., pp. 63–152). Boca Raton, FL: CRC Press, Taylor & Francis Group.
James, W. R. L., & Knight, B. H. (1965). Errors in estimating time since death. Medicine, Science and the Law, 5(2), 111–116.
Khorasgani, Z. N., Arzi, A., Salahcheh, M., & Latifi, M. (2009). A study of correlation between vitreous potassium level and postmortem
interval. Toxicology Letters, 189. https://doi.org/10.1016/j.toxlet.2009.06.798
Knight, B. (1988). The evolution of methods for estimating the time of death from body temperature. Forensic Science International, 36(1-2),
47–55.
Knight, B., & Madea, B. (2016). Historical review on early work on estimating the time since death. In Estimation of the time since death (3rd
ed., pp. 7–16). Boca Raton, FL: CRC Press, Taylor & Francis Group.
Krompecher, T. (2016). Rigor mortis: Estimation of the time since death by evaluation of cadaveric rigidity. In Estimation of the time since
death (3rd ed., pp. 41–58). Boca Raton, FL: CRC Press, Taylor & Francis Group.
Lange, N., Swearer, S., & Sturner, W. Q. (1994). Human postmortem interval estimation from vitreous potassium: An analysis of original data
from six different studies. Forensic Science International, 66(3), 159–174.
Lew, E., & Matshes, E. (2005). Postmortem changes. In Forensic pathology principles and practices (pp. 527–554). Elsevier Academic Press.
Madea, B. (2016). Estimation of the time since death. Boca Raton, FL: CRC Press, Taylor & Francis Group.
Madea, B. (2016b). Postmortem lividity: Hypostasis and timing of death. In B. Knight (Ed.), Estimation of the time since death (3rd ed.,
pp. 59–62). Boca Raton, FL: CRC Press, Taylor & Francis Group.
Madea, B., & Henssge, C. (2016). General remarks on estimating the time since death. In Estimation of the time since death (3rd ed., pp. 1–6).
Boca Raton, FL: CRC Press, Taylor & Francis Group.
Madea, B., Kernbach-Wighton, G., Henssge, C., Doberentz, E., & Musshoff, F. (2016). Autolysis, putrefactive changes, and postmortem chem-
istry. In Estimation of the time since death (3rd ed., pp. 153–212). Boca Raton, FL: CRC Press, Taylor & Francis Group.
Michaud, J.-P., Moreau, G., & Schoenly, K. G. (2014). On throwing out the baby with the bathwater: A reply to Wells. Journal of Medical
Entomology, 51(5), 494–495.
Payne, J. A. (1965). A summer carrion study of the baby pig Sus scrofa Linnaeus. Ecology, 46, 592–602.
Schoenly, K., & Reid, W. (1987). Dynamics of heterotropic succession in carrion arthropod assemblages: Discrete seres of a continuum of
change? Oecologia, 73, 192–202.
Schoenly, K., & Reid, W. (1989). Dynamics of heterotrophic succession in carrion revisited: A reply to Boulton and Lake. Oecologia, 79,
140–142.
Smith, K. G. V. (1986). A manual of forensic entomology. Ithaca, NY: Cornell University Press.
Stephens, R. J., & Richards, R. G. (1987). Vitreous humor chemistry: The use of potassium concentration for the prediction of the postmor-
tem interval. Journal of Forensic Sciences, 32(2).
Sturner, W. (1963). The vitreous humour: Postmortem potassium changes. The Lancet, 281(7285), 807–808.
Tomberlin, J., Mohr, R., Benbow, M., Tarone, A., & Vanlaerhoven, S. (2011). A roadmap for bridging basic and applied research in forensic
entomology. Annual Review of Entomology, 56(1), 401–421.
Tz'u, S., & McKnight, B. E. (1981). The washing away of wrongs: Forensic medicine in 13th-century China. Ann Arbor: University of Mich.,
Center for Chinese studies.

How to cite this article: Sutton L, Byrd J. An introduction to postmortem interval estimation in medicolegal
death investigations. WIREs Forensic Sci. 2020;e1373. https://doi.org/10.1002/wfs2.1373

You might also like