ISSN 0006-2979, Biochemistry (Moscow), 2024, Vol. 89, No. 2, pp. 341-355 © Pleiades Publishing, Ltd., 2024.
341
REVIEW
Exploring Patterns of Human Mortality and Aging:
A Reliability Theory Viewpoint
Leonid A. Gavrilov
1,2,a
* and Natalia S. Gavrilova
1,2
1
NORC at the University of Chicago, 60637 Chicago, IL, USA
2
Institute for Demographic Research, Federal Center of Theoretical and Applied Sociology,
Russian Academy of Sciences, 109028 Moscow, Russia
a
e-mail: lagavril@yahoo.com
Received December 14, 2023
Revised January 27, 2024
Accepted January 28, 2024
AbstractThe most important manifestation of aging is an increased risk of death with advancing age, a mortal-
ity pattern characterized by empirical regularities known as mortality laws. We highlight three significant ones:
the Gompertz law, compensation effect of mortality (CEM), and late-life mortality deceleration and describe new
developments in this area. It is predicted that CEM should result in declining relative variability of mortality
at older ages. The quiescent phase hypothesis of negligible actuarial aging at younger adult ages is tested and
refuted by analyzing mortality of the most recent birth cohorts. To comprehend the aging mechanisms, it is cru-
cial to explain the observed empirical mortality patterns. As an illustrative example of data-directed modeling
and the insights it provides, we briefly describe two different reliability models applied to human mortality pat-
terns. The explanation of aging using a reliability theory approach aligns with evolutionary theories of aging,
including idea of chronic phenoptosis. This alignment stems from their focus on elucidating the process of or-
ganismal deterioration itself, rather than addressing the reasons why organisms are not designed for perpetual
existence. This article is a part of a special issue of the journal that commemorates the legacy of the eminent
Russian scientist Vladimir Petrovich Skulachev (1935-2023) and his bold ideas about evolution of biological aging
and phenoptosis.
DOI: 10.1134/S0006297924020123
Keywords: aging, mortality, Gompertz model, compensation effect of mortality, mortality deceleration, reliability
theory of aging, evolutionary models of aging
* To whom correspondence should be addressed.
INTRODUCTION
The increasing interest in unraveling the intri-
cacies of aging underscores the necessity for a com-
prehensive theoretical framework. There has been a
significant surge in the volume of empirical data on
aging, reflecting a substantial expansion in our un-
derstanding of the aging processes. The exploration of
individual genes, pathways, and molecules in under-
standing the mechanisms that modulate aging has in-
deed seen significant progress. Researchers have made
strides in identifying the components involved in the
aging process. However, the challenge lies in compre-
hending how these various factors interact on a larger
scale to influence the aging processes. While we can
pinpoint specific genes, pathways, and molecules, the
integration of these elements into a comprehensive un-
derstanding of aging, including the emergence of func-
tional phenotypes like mortality laws, remains a com-
plex puzzle [1].
Evolutionary theories can provide a broader un-
derstanding of the aging phenomenon. Until recently,
there was a consensus among gerontologists that there
is no specific program of aging, because such program
could not appear when overwhelming majority of an-
imals in the wild and humans in the past did not sur-
vive to advanced ages [2-4]. However, this view was re-
vised in the last decade and there are more arguments
now in favor of aging as a program. Vladimir P. Sku-
lachev was among the researchers who promoted an
GAVRILOV, GAVRILOVA342
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
idea of programmed aging. He attempted to resurrect
Weismann’s theory of programmed death by proposing
the existence of a distinct self-destruction program for
entire organisms, termed “phenoptosis” [5-7]. This pro-
gram is believed to serve “crucial functions, cleansing
communities of organisms from undesirable individ-
uals”  [7]. As such, Weismann’s evolutionary theory of
programmed death continues to be a subject of scien-
tific debate and inquiry. Moreover, the ideas of harm-
ful role that older individuals play in population by
accumulating infections gain popularity among geron-
tologists  [6,  8]. Nevertheless, theories and hypotheses
about programmed aging primarily elucidate the rea-
sons organisms are not inherently designed to be flaw-
less and fall short in explaining the actual process of
age-related deterioration itself.
The primary limitation of evolutionary theories of
aging is that they rely on the concept of natural selec-
tion and the diminishing impact of natural selection
with age. However, aging is observed in non-replicat-
ing technical systems like automobiles, which lack the
capacity for procreation and are therefore unaffected
by evolution driven by natural selection.
Recognizing this broader context prompts the ex-
ploration of more universal explanations for aging that
transcend the constraints of evolutionary theories. Be-
yond overarching concepts, there is a crucial need to
explore empirical regularities in aging and mortality,
often referred to as mortality laws. This pursuit aims
to provide a comprehensive understanding of aging
phenomena that goes beyond the scope of traditional
evolutionary frameworks.
In this context, we examine the established mor-
tality laws and regularities alongside recent advance-
ments in the field of biodemography of aging. The ex-
isting patterns of mortality and aging are scrutinized
through the lens of reliability theory. Additionally, this
discussion encompasses the presentation of further
developments in reliability models of aging, contrib-
uting to an enhanced understanding of the underlying
mechanisms shaping the aging process.
MORTALITY LAWS IN THE BIOLOGY
OF AGING AND LIFESPAN
To gain a more precise understanding of the mech-
anisms behind an organism’s decline, we can explore
mortality patterns frequently referred to as mortali-
ty laws. Here we assume definition of aging applied
in reliability theory, which considers aging as a pro-
cess leading to increasing risk of death with age [9].
In ecology this process is also called an actuarial se-
nescence [10].
The Gompertz–Makeham law. In 1825, the Brit-
ish actuary Benjamin Gompertz unveiled a principle
of mortality, now recognized as the Gompertz law [2,
11-13]. According to the Gompertz law, human mortal-
ity rates double approximately every 8 years of adult
age:
μ(x) = Re
αx
, (1)
where x is age, R and α are parameters (Gompertz in-
tercept and Gompertz slope, since Gompertz law rep-
resents a straight line in semi-log coordinates).
An exponential (Gompertzian) increase of mortal-
ity rates with advancing age is observed across various
biological species, encompassing fruit flies [2, 14], flour
beetles like Tribolium confusum [2], mice  [14,  15], ba-
boons [16,  17], and others. Most importantly, this law
describes mortality of humans [2,  11-14,  18,  19].
In reality, organisms’ hazard rates can encompass
both non-aging and aging components, as seen in the
Gompertz–Makeham law of mortality [2,  11,  13,  19,  20]:
μ(x) = A + Re
αx
. (2)
Within this equation, the initial, age-independent
term (known as the Makeham parameter, “A”) signifies
the constant, “non-aging” facet of the hazard rate, pre-
sumably stemming from external causes of death, such
as accidents and acute infections. In biodemography
this constant term is called background mortality [2].
The subsequent, age-dependent term (the Gompertz
function, Re
αx
) represents the “aging” component, pre-
sumably arising from age-related degenerative diseases
like cancer and heart disease. This aging-related term
received the name of senescent mortality [21]. Carnes
and Olshansky proposed another way to classify total
mortality using causes of death. Mortality from exter-
nal causes of death and infections they called an extrin-
sic mortality while mortality from aging-related causes
they called an intrinsic mortality[22]. Thus, extrinsic
and intrinsic types of mortality may both depend on
age. It is important to note that the slope coefficient
α characterizes an “apparent aging rate,” indicating
the speed of age-related deterioration in mortality. If α
equals zero, there is no apparent aging, which means
that death rates remain constant with age.
Some authors have proposed a generalized form
of the Gompertz–Makeham law (GGM) [23]. In this
alternative perspective, the exponent, indicative of
stress resistance, is considered to be a function be-
yond a linear form. The Makeham term, traditionally
treated as a constant, is redefined to be associated
with mortality resulting from inherently irresistible
stresses and depends on age. This departure from the
standard Gompertz–Makeham law introduces a more
flexible approach to modeling mortality [23].
For technical systems one of the most popular
models for failure rate of aging systems is a Weibull
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 343
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
Fig. 1. Mortality of Norwegian women in 2010 in semi-log scale and its linear fit according to the Gompertz law. Data source:
Human Mortality Database (www.mortality.org).
model, the power-function increase in failure rates
with age [2]:
μ(x) = αx
β
for x≥0, where α,β>0. (3)
The Weibull law, proposed by Swedish engineer
and mathematician W. Weibull in 1939 to describe the
strength of materials, is widely employed in charac-
terizing the aging and failure of technical devices [9].
Ithas also found occasional application in the study of
a limited number of biological species including nema-
todes C. elegans [24-26]. According to the Weibull law,
the logarithm of failure rates exhibits a linear increase
with the logarithm of age.
A comparative meta-analysis of 129 life tables for
fruit flies and 285 life tables for humans demonstrated
that the Gompertz law of mortality offers a more ac-
curate fit to the data for each of these two biological
species compared to the Weibull law [2].
While working with human data it was found that
mortality of women in semi-log scale demonstrated
slightly more concave trajectory than predicted by the
Gompertz model and mortality of men was somewhat
more convex [2]. Similar deviations from the Gompertz
law were observed by other researchers [27]. Further
analyses of local estimates of the Gompertz slope using
life table aging rate (LAR) found increase of LAR after
age 60 mostly among women [28,  29]. Recently accel-
eration of LAR after age 60 was found for both men
and women and for period and cohort mortality using
contemporary data [30]. Despite these small deviations
Gompertz law describes mortality remarkably well (see
Fig.1). There are indeed small fluctuations of mortal-
ity around the main Gompertz trajectory whereas the
general direction of mortality line remains stable. In
humans the Gompertz model fits both period and co-
hort data [31, 32].
In addition to the commonly used Gompertz and
standard two-parameter Weibull laws, another mor-
tality law known as the binomial law of mortality has
been proposed and theoretically justified through the
lens of system reliability theory [2, 9]. This particular
law is considered a special case of the three-parameter
Weibull function. Notably, it incorporates a negative
location parameter, offering a nuanced perspective on
the dynamics of mortality and aging within the frame-
work of reliability theory:
μ(x) = α (x
0
+ x)
β
. (4)
In this equation, the parameterx
0
is referred to as
the initial virtual age of the system [2, 9, 33]. This pa-
rameter, measured in units of time, represents the age
at which an initially ideal system would accumulate
damage equal to those of a real system at the starting
age (x=0).
When a system is in its undamaged state with an
initial virtual age of zero, the mortality rate adheres
to a power function of age, consistent with the charac-
teristics outlined in the Weibull law. However, as the
initial damage load accumulates, there is a deviation
from the Weibull law. This departure becomes more
GAVRILOV, GAVRILOVA344
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
pronounced with increasing levels of initial damage
load, eventually leading to a transition where the fail-
ure dynamics align with the quasi-Gompertz mortality
law [9].
Compensation law or compensation effect of
mortality. Another empirical finding, known as the
compensation law or compensation effect of mortality
(CEM), in its strong form pertains to mortality conver-
gence at older ages. This is when higher values for the
slope parameter α (in the Gompertz function) are offset
(compensated) by lower values of the intercept param-
eter R in different populations of a specific species:
ln(R) = ln(M) − Bα, (5)
where B andM represent species-specific constants.
This relationship is occasionally referred to as the
Strehler–Mildvan correlation [13,  34], although this
correlation was significantly affected by biases in pa-
rameter estimation arising from neglecting the age-in-
dependent mortality component, the Makeham term  A
[2,  20]. Parameter B is called the species-specific lifes-
pan, and parameter M is called the species-specific
mortality rate [2, 9,  35]. Recent estimates of parame-
ters B andM using data for contemporary human pop-
ulations showed that these estimates remain rather
stable over time [31,  32]. These parameters represent
the coordinates where all mortality trajectories con-
verge into a single point when extrapolated using the
Gompertz function [2]. This signifies that in disadvan-
taged populations within a specific species, high mor-
tality rates are compensated for by a slower apparent
“aging rate” (resulting in a longer mortality doubling
period). As a consequence of this compensation, rela-
tive differences in mortality rates have a tendency to
decrease with age within a given biological species.
The term “Compensation Effect of Mortality” was
introduced in 1978 when incorporating the Makeham
parameter yielded substantially different parameter
estimates for the Strehler–Mildvan correlation [35].
This effect is defined by the convergence of senescent
(age-dependent) mortality patterns at advanced ages
[2,  35]. CEM is observed not only for humans but for
some other species [2,  36,  37].
There were assertions that the Strehler–Mildvan
correlation arises as a statistical artifact of spurious cor-
relation between the estimates of the Gompertz param-
eters and does not exist in reality [38]. Nevertheless,
even when controlling for collinearity, the correla-
tion coefficients between the Gompertz parameters
experience only a modest reduction, yet the core cor-
relation remains intact [31]. Furthermore, it is worth
noting that the convergence of mortality trajectories
in a semi-logarithmic scale (CEM) at older ages can
be observed independently of the estimation of the
Gompertz parameters [31, 32]. Also, in human popula-
tions CEM is observed for both period and cohort data
on mortality [31, 32].
As mortality convergence is approached, a de-
crease in the relative variation of mortality is expected.
Consequently, we anticipate a decline in the relative
variation of mortality, assessed through metrics like
the coefficient of variation and the standard deviation
of the log of mortality, as the convergence point cor-
responding to the species-specific lifespan is reached.
This theoretical prediction will be subject to validation
in upcoming studies.
Late-life mortality deceleration. At more ad-
vanced ages (beyond age 70 in humans), sometimes a
phenomenon known as “old-age mortality decelera-
tion” occurs, wherein death rates increase with age at
a slower pace than expected based on the Gompertz
law [2,  21,  39-41]. Some biologists called this cessation
of mortality growth (non-aging state) “a revolution
for aging research” [42], although for humans it was
known by actuaries since the Gompertz times. This de-
celeration in mortality ultimately leads to “late-life
mortality leveling-off” and “late-life mortality plateaus”
at extremely old ages [2,  43,  44].
Actuaries, including Gompertz himself, were among
the first to observe this phenomenon. They proposed a
logistic formula to model the age-related increase in
mortality, aiming to address the decrease in mortality
growth at advanced ages.
Figure 2 illustrates mortality deceleration at older
ages using an example of mortality for cohort of U. S.
women born in 1886. Note that after age 90 years the
observed mortality deviates from the Gompertz law in
this particular case.
In humans mortality deceleration is almost always
observed for cross-sectional data. In the case of cohort
data, there is no consensus on the pattern of late-life
mortality trajectories. Earlier studies suggested that
mortality after age 80 displayed a slower increase
with age compared to the exponential Gompertz law.
This deceleration in mortality at advanced ages was ob-
served in both cohort and period data for various coun-
tries [21]. However, for U. S. cohort data, it was revealed
that mortality adheres to the Gompertz law within wide
age range of 80-106 years [45, 46]. Similar results were
recently obtained using French cohort data [47]. Other
scholars have found that the extent of mortality decel-
eration varies among different countries [48, 49]. The
primary issue with these studies was the attempt to
deduce the existence of only one conceivable mortal-
ity trajectory shape. Recently the existing controver-
sy about the shape of mortality at advanced ages was
resolved by studying long series of U. S. cohort data.
It turns out that mortality of earlier birth cohorts
(born before 1887) always demonstrates mortality de-
celeration. On the other hand, later birth cohorts fol-
low the Gompertz law up to ages 105-106 years [50].
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 345
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
Fig. 2. Mortality deceleration for cohort of U. S. women born in 1886. Data source: Human Mortality Database (www. mortality.org).
This trend called gompertzialization of mortality tra-
jectory is observed in other countries as well, although
with slower pace [51].
EXPLORING FURTHER ASPECTS OF AGING
As widely recognized, the aging phase constitutes
a substantial portion of the lifespan of many species,
necessitating that any mortality model offer an expla-
nation for this extended period of life. For modeling
the aging process, it is important to take into account
some additional age-related phenomena.
Similarity between mortality of biological and
technical systems. A notable parallel exists between
living organisms and technical devices in their age-re-
lated mortality patterns, often following what’s known
as the “bathtub curve”. [9]. This curve comprises three
distinct periods. Initially, mortality rates are high and
gradually decline with age, referred to as the “work-
ing-in” period or the phase of “burning-out” defective
components. A similar early period, known as “infant
mortality,” can be observed in most living organisms,
including humans. Following this, there is the “normal
working period,” characterized by relatively low and
stable failure rates. While this period exists in humans
as well, it tends to be relatively short, lasting approx-
imately 10-15 years before transitioning to the third
period known as “the aging period.” During this phase,
mortality rates increase inexorably with age, follow-
ing an explosive exponential trajectory, akin to the
Gompertz curve. For humans, this aging period typi-
cally spans from around 20 to 100 years. This similari-
ty in mortality patterns between technical and biolog-
ical systems is further emphasized by the presence of
a fourth common period at extremely advanced ages.
This phase is recognized in biology as “late-life mortal-
ity leveling-off” and is also observed in technical sys-
tems [52].
Keeping in mind this similarity in the phases of
age- specific mortality between biological and techni-
cal systems, Siler suggested an empirical equation that
described mortality of biological organisms during the
entire life period [53]. This equation considers three
mortality terms. The first term describes mortality de-
cline with age after birth and can be described by de-
clining exponential function. Two other terms are rep-
resented by already known background and senescent
mortality. The minimum of mortality around age 10
received lately an attention from researchers. It turns
out that the minimum of mortality measured in period
data looks very narrow while for cohort data mortality
may be flat for much longer period. This long period
of almost flat mortality was called a “quiescent phase”
[54] and was observed for cohorts born around the
1920s (Fig.3).
This phenomenon serves as a clear example of an
almost zero apparent (actuarial) aging rate, arising from
the counteracting influences of two opposing forces
that offset one another: the mitigating effect of mortal-
ity reduction (attributable to advancements in health-
care and improved living conditions) and the inexora-
ble force of aging [32]. Figure3 shows that mortality at
age 10 for 1920th cohort is almost the same as at ages
GAVRILOV, GAVRILOVA346
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
Fig. 3. Mortality for selected Swedish female birth cohorts. Data source: Human Mortality Database (www.mortality.org).
30-40 years, so that there is no clear minimum of mor-
tality around ages 10-40 years. This phase looks like
rather long “normal working period” in technical sys-
tems, but is rather a result of declining mortality due
to healthcare achievements and improvement of living
conditions [51].
The authors of the original study of quiescent
phase limited their analysis by extinct cohorts only[54].
However, mortality at younger ages can be studied
for more recent birth cohorts. Figure 3 shows that for
1950 and 1980 birth cohorts mortality is not flat and
slowly increases with age. It also looks like the posi-
tion of mortality minimum is shifting to younger ages
over time. When mortality from infectious diseases
was mostly eradicated and historical mortality decline
slowed down, the quiescent phase practically disap-
peared and mortality started to grow from very young
ages (8-12 years) without visible hump. It should be
noted that we use here data for women in order to
avoid substantial external mortality due to social fac-
tors, which is common for men at ages 18-25 years.
Figure 3 shows that mortality for more recent birth
cohorts can be described by two rather than three
phases with disappearing the “normal working peri-
od” or quiescent phase.
The concept of high initial damage load. In 1991
it was suggested that early developmental processes
in living organisms generate an exceptionally high
load of initial damage, comparable in magnitude to
the subsequent accumulation of age-related deterio-
ration throughout the entire adult lifespan [2]. While
this concept, known as the High Initial Damage Load
(HIDL) hypothesis [55], may appear counterintuitive,
it aligns with empirical observations of developmental
noise and substantial cell losses during early develop-
ment [56]. Recent advancements in molecular develop-
mental biology have recognized the stochastic nature
of development, often referred to as “developmental
noise.” This phenomenon has the potential to induce
phenotypic heterogeneity even in the absence of any
other alterations in genes or the environment [57].
Ina human body, approximately a hundred thousand
cells are generated every second through mitosis, and
a comparable number undergo a physiological suicide
process known as apoptosis. A significant portion of
cells produced during mammalian embryonic devel-
opment experiences physiological cell death before
the conclusion of the perinatal period [58]. Significant
cell losses during early development create conditions
for the uneven distribution of organisms based on the
number of remaining cells, which can be modeled us-
ing the binomial or even the Poisson distribution. [2].
The idea of high initial damage of biological sys-
tems recently received a further development. It was
found that the mortality and incidence of age-related
diseases exhibit a U-shaped curve, with the minimum
occurring before puberty. However, quantitative bio-
markers of aging, such as somatic mutations and DNA
methylation, do not follow this pattern and continue to
increase starting from birth or even before birth [59].
It was suggested that aging initiates early but is con-
cealed by early-life mortality decline. According to this
idea aging may be represented by the growth of the
sum of deleterious changes, the deleteriome [59, 60].
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 347
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
The concept of a high initial damage load also
suggests that events occurring early in life can impact
survival in later adulthood through the level of ini-
tial damage. This prediction has been substantiated
for early-life indicators like parental age at a person’s
conception and the month of a person’s birth [61-66].
There is an increasing body of evidence supporting the
notion of fetal origins of adult degenerative diseases
[67,  68] and the early-life programming of aging and
longevity [55, 69-71].
Loss of functional elements in an organism over
its lifespan. Aging is characterized by a progressive
loss of functional tissue influenced by a combination
of genetic and environmental factors, nutrition, and
lifestyle choices [72, 73].
The accumulation of damage in various cellular
structures, coupled with the loss of fully differentiated
and irreplaceable cells like neurons and cardiomyo-
cytes, should be considered as irreversible. If these ir-
replaceable components of an organism age and even-
tually perish, it follows that the organism as a whole
will experience aging as well. This emphasizes the crit-
ical role of cellular damage and the loss of vital com-
ponents in the overall aging of organisms [74].
Cell death mechanisms have traditionally been
categorized into two types: programmed cell death
(PCD) mechanisms, which require energy, and necrotic
cell death mechanisms, which do not [73,  75]. The in-
creased occurrence of PCD during aging is implicated
in the decline of the immune system, skeletal muscle
wasting (sarcopenia), loss of cells in the heart, and neu-
rodegenerative diseases. Throughout the aging pro-
cess, several tissues experience cell loss attributed to
either PCD or PCD-like processes. In mammals, there
is an aging-associated skeletal muscle atrophy known
as sarcopenia, characterized by reductions in muscle
fiber size and fiber loss [73]. Necroptosis, a regulated
form of cell death, plays a role in the genesis and pro-
gression of various life-threatening diseases, including
cancer, neurological disorders, cardiac myopathy, and
diabetes [73].
Research indicates a decline in the numbers of spe-
cialized cells with age, including a significant reduc-
tion in nephrons in healthy human kidneys with aging.
Comparing the youngest (18-29 years) and oldest (70-
75 years) age groups, there was a 48% decrease in the
number of nonsclerotic glomeruli, while cortical vol-
ume only decreased by 16% [76].
While there is substantial knowledge about spe-
cific molecular mechanisms of cell death [77], the age-
specific loss of cell numbers is less explored and is
presented in a limited number of publications [78-80].
Existing studies have uncovered the possibility that
some cells within the aging organism may exhibit
nonaging characteristics. Notably, in a diverse spectrum
of aging-related neurodegenerative conditions (18 di-
verse examples of inherited and acquired neurode-
generation including Parkinson’s disease), the rate of
neuronal death does not increase with age [79,  81,  82].
Neurons in different parts of brain of cognitively
healthy humans show constant rate of atrophy with
age [83]. These findings align with the observation that
“an impressive range of cell functions in most organs
remain unimpaired throughout the life span” [11], p.425.
Therefore, the current understanding of the kinetics of
cell loss with age indicates that an exponential distri-
bution (with constant mortality) is a plausible approx-
imation for the mechanism of loss in vital elements,
such as functional cells or telomeres.
EXPLAINING MORTALITY LAWS
THROUGH THE LENS OF RELIABILITY THEORY
There is a whole spectrum of models attempting
to explain the observed mortality phenomena. Hetero-
geneity models were among the first models attempt-
ing to explain human aging and mortality and became
popular after the pioneer publication by Beard in 1959
[84]. These models assume exponential increase for
baseline risk of death. Heterogeneity in these models
is introduced usually by postulating gamma distribu-
tion of individual risks [85-87]. Heterogeneity models
are famous for explaining late-life mortality decelera-
tion and mortality plateau as a result of mortality se-
lection. However, the exponential growth of mortality
risk with age is postulated in these models in advance.
Thus, it would be more interesting to consider a mod-
el, which can use organism’s structure and properties
to derive the observed empirical mortality regularities.
It is noteworthy that the phenomena of increasing
mortality with age, followed by a plateauing of mortal-
ity rates, are inherently anticipated features of reliabil-
ity models that conceptualize aging as the progressive
accumulation of damage or loss of vital elements with
age [2, 9, 33]. Reliability models of aging presented in
more detail earlier [2, 33] consider living organism
as a system with series-parallel reliability structure.
In a system of n independent components connected
in series, the entire system fails if any one of the com-
ponents fails. This means that the reliability of the sys-
tem is contingent on the reliability of each individual
component in series. It is clear that human organs like
heart, brain or liver can be considered as vital com-
ponents connected in series because failure of any of
these organs leads to organism’s death. On the other
hand, in a parallel system of n independent compo-
nents, the system fails only when all the components
simultaneously fail. In this configuration, the system
remains operational as long as at least one of the par-
allel components continues to function, providing a
redundancy that enhances overall system reliability.
GAVRILOV, GAVRILOVA348
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
It is also clear that functional cells in each vital organ
can be considered as components connected in paral-
lel, although in real organs the threshold for normal
functioning may be higher than one element (cell).
Thus, reliability models in the context of living organ-
isms consider the inherent functional structure and
dynamics of biological systems. These models acknowl-
edge and incorporate observed processes, such as the
loss of functional cells with age. This approach helps
in understanding the intricate relationships between
different elements within biological systems and their
impact on overall mortality.
Consider a parallel system comprising n non-ag-
ing elements, each characterized by a constant failure
rate denoted as µ and a reliability (survival) function
expressed as e
–µx
[9]. In such a scenario, the reliabil-
ity function for the entire parallel system can be de-
scribed as:
S(x) = 1 – (1 – p)
n
= 1 – (1 − e
−μx
)
n
. (6)
This equation applies to the most straightforward
scenario where the failures of individual elements are
statistically independent. As a result, the failure rate of
the entire system, denoted as µ
s
(x), can be expressed
in the following manner:
μ
s
(x) =
dS(x)
S(x)dx
=
nμe
−μx
(1 − e
−μx
)
n−1
1 − (1 − e
−μx
)
n
, (7)
≈ 
n
x
n−1
when x << 1/µ (early-life period approxima-
tion, when 1 − e
–µx
≈ µx);
≈  μ when x >> 1/µ (late-life period approximation, when
1 − e
–µx
 ≈ 1).
Hence, the system’s failure rate initially exhibits
growth in accordance with an age-dependent power
function (following the Weibull law). According to this
model, systems with varying initial redundancy lev-
els (parameter n) will display distinct failure rates in
early stages, but these differences will diminish over
time as the rates converge toward the upper limit de-
termined by the rate of elements’ loss (parameter µ).
Consequently, the expected outcome of this model
aligns with the compensation law of mortality (in its
weak form).
The failure rate of a simple parallel system com-
posed of non-aging elements exhibits an increase with
age, contrary to the Gompertz law, with the initial fail-
ure kinetics following the Weibull law. This deviation
from the Gompertz law arises from the model’s as-
sumption that the system is constructed with initial-
ly ideal structures, where all elements are functional
from the outset. This limitation highlights the impor-
tance of considering real-world scenarios where com-
ponents may not be flawless initially and may experi-
ence variations in functionality over time, influencing
the overall failure kinetics of the system. In order to
obtain the quasi-Gompertz mortality growth we need
to consider models with distributed redundancy.
It is important to highlight that reliability models
align seamlessly with evolutionary models, including
the concept of programmed death. The accrual of dam-
age might follow a stochastic process, whereas the pa-
rameters governing this damage, such as the initial
redundancy level and the rate of damage, could be
preprogrammed. Evolutionary models elucidate why
organisms are constructed with distinct properties,
while reliability models specifically elucidate the pro-
cess of deterioration itself, which may be considered
as slow or chronic phenoptosis [71, 88].
Reliability model of initially homogeneous pop-
ulation. The model, which was published earlier [33]
examined a scenario where blocks (e.g., specific or-
gans) within each organism exhibit varying degrees of
redundancy, while the organisms themselves are ini-
tially considered initially identical to each other and
share an equal risk of death. This assumption can be
justified in the following cases [2]:
1.  Population homogeneity may occur when a rig-
id genetic program determines the initial degree of re-
dundancy for each block (organ) in the organism. This
situation may occur in the course of programmed cell
death during early development. In this case, the vari-
ability in block redundancy is not entirely random,
and the homogeneous models are applicable because
the genetically programmed distribution of blocks ac-
cording to their redundancy can be approximated by
the Poisson or binomial distribution.
2.  Organisms may have nearly identical initial dis-
tributions of the number of blocks with different re-
dundancy levels, even when the redundancy formation
mechanism is random. This occurs when the number
of irreplaceable (vital) blocks is very large. Consequent-
ly, the population is practically homogeneous in terms
of the risk of each organism dying, despite potential
heterogeneity in the risk of failure of individual blocks.
The simplest model within this family of reliabil-
ity models is the series-parallel structure with distrib-
uted redundancy within the organism. This model as
outlined in [33] considers the distribution of subsys-
tems based on initially functional elements, described
by the Poisson law due to a high initial damage load.
In such systems, the failure rate can be initially ap-
proximated by the exponential (Gompertz) law, with
subsequent mortality leveling-off [33]. In systems with
lower damage levels, where initially functional ele-
ments follow a binomial distribution, the failure rate
experiences initial growth in line with the binomial
law [2].
It is worth noting that there have been allegations
of errors in this straightforward model [89]. However,
it is important to point out that the authors failed to
acknowledge that this model was designed for a homo-
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 349
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
geneous population and improperly recommended the
use of a formula tailored to a heterogeneous popula-
tion [89].
One interesting conclusion from the model of ini-
tially homogeneous population is related to the oppor-
tunity of estimating the rate of vital elements loss in
human organism or true aging rate. This rate is ap-
proximately equal to the inverse of the species-spe-
cific lifespan [1/B, Bcomes from equation(5)]. It was
found that the estimated species-specific lifespan is
stable over time and is equal to 95-97 years [2, 31].
Thus, the estimated rate of loss of vital elements is ap-
proximately equal to 1% per year. It is interesting that
this rate is consistent with the empirical estimates of
annual cell loss in several neural tissues– 0.6-1.6% [78].
Empirically estimated rate of telomere loss in periph-
eral blood mononuclear cells in humans is somewhat
lower – 0.5% base pairs per year [90], but still is of
similar magnitude. It was found that the rate of telo-
mere loss is a species-specific trait and is proportional
to the lifespan of animal species [90].
Reliability model of heterogeneous population.
Accounting for population heterogeneity leads us to an-
other model, which provides an explanation for all the
basic mortality laws, even in the simplest case where
the organism comprises a single vital block with n el-
ements.
The model considers the simplest case when the
organism consists of a single vital block with n ele-
ments connected in parallel with q being the proba-
bility that an element is initially functional. Then the
probability of encountering an organism with i initial-
ly functional elements out of a total number n of ele-
ments is given by the binomial distribution law.
The final formula for failure rate in heteroge-
neous population,µ
p
(x), is (see[2] for more detail):
µ
p
(x) = −
F′(x)
1 − F(x)
=
nqμe
−μx
(1 − qe
−μx
)
n−1
1 − (1 − qe
−μx
)
n
, (8)
≈ Cnqμ
(1 − q + qμx)
n−1
for x << 1/µ;
≈ μ for x>>1/µ,
where C is a normalizing factor.
Thus, the hazard rate of a heterogeneous popula-
tion at first grows with age according to the binomial
law of mortality, then asymptotically approaches an
upper limit µ:
µ
p
(x) ≈ Cn(qμ)
n
[
1 − q
qμ
+ x
]
n−1
 = Cn(qμ)
n
(x
0
+ x)
n−1
, (9)
for x << 1 / μ
μ
p
(x) ≈ μ for x >> 1 / μ
where x
0
= (1 − q) / qμ, a parameter which is called the
initial virtual age of the population. This parameter
has the dimension of time, and corresponds to the age
by which an initially homogeneous population would
have accumulated as much damage as a real popula-
tion actually possesses at the initial moment in time
(at x  =  0). In particular, when q = 1, i.e., when all the
elements in each organism are functional at the out-
set, the initial virtual age of the population is zero and
the hazard rate of population grows as a power func-
tion of age (the Weibull law). However, when the pop-
ulation is not initially homogeneous (q <  1), we arrive
at the already mentioned binomial law of mortality.
Thus, the heterogeneous population model described
here can also provide a theoretical justification for the
binomial law of mortality.
Heterogeneity model also addresses the compen-
sation effect of mortality. The compensation effect of
mortality is evident when variations in mortality re-
sult from differences in the number of elements in the
organism(n) between populations, while other param-
eters, including the true aging rate(µ), remain nearly
identical for all populations of the same species [2, 9].
Figure  4 illustrates Gompertz, binomial and Weibull
models fitting mortality of Norwegian women born
in 1920. In the case of binomial model, it is assumed
that every block (e.g., vital organ) has on average 50
vital elements and the initial virtual age (indicator
of initial damage load) estimated using nonlinear re-
gression method is equal to 370 years. Note that bi-
nomial model fits mortality data almost as well as
the Gompertz model with similar values of Akaike
Information Criterion of goodness of fit (-466 and
-462 respectively) while Weibull model significantly
underestimates mortality at younger ages (see Fig.  4).
It is important to note that extrinsic mortality below
the age of 60 years is significant. However, adjusting
for extrinsic mortality using the Gompertz–Makeham
model becomes challenging due to the variation of
Makeham term over time in cohort data. Thus, even
this simplified heterogeneity model is able to fit aging-
related mortality of humans.
In summary, the model of a heterogeneous pop-
ulation provides an explanation for the regularities
of organism mortality: the initial quasi-exponential
growth in the hazard rate, followed by mortality decel-
eration, along with the compensation effect of mortali-
ty as follows from the model formulas [2].
The heterogeneous population model shares basic
conclusions with the initially homogeneous model of
series- connected blocks with varying degrees of redun-
dancy [33]. However, these are two distinct models.
Inthe first model, individual mortality risk is uniform
across all organisms and grows exponentially with age.
In the second model, there initially exist n subpopu-
lations of living organisms with differing death risks
that increase as a power function rather than expo-
nentially with age.
The fact that these two different models yield near-
ly identical interpretations of certain mortality phe-
nomena offers some reason for optimism. For instance,
GAVRILOV, GAVRILOVA350
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
Fig. 4. Fitting mortality of Norwegian women born in 1920 by binomial law and competing Gompertz(a) and Weibull(b) laws.
Data source: Human Mortality Database (www.mortality.org).
the compensation effect of mortality is a common fea-
ture of all these models, occurring only when the rate
of irreversible age changes (true aging rate) remains
constant within a species. This treatment of the com-
pensation effect of mortality is not unique to the mod-
els discussed here, as it is consistent with other models
as well [34,  35,  91,  92]. The existence of various com-
peting models does not hinder the reliable and mean-
ingful interpretation of many mortality phenomena,
as multiple models can reach agreement on several
aspects.
Further developments of reliability models. We
would like to mention here several studies developing
reliability models further. One of the first reliability
models of aging was suggested in 1978 and was based
on the phenomenon of linear decline in the function or
cells over time [91]. This simple model could explain
the Gompertz law, CEM and late-life mortality plateau.
This approach was developed further by Milne who
created “nested binomial” model also explaining exist-
ing mortality regularities [93]. These models explained
mortality in terms of probability of dying and did not
consider organism’s structure.
Laird and Sherrat extended the approach of ap-
plying reliability theory to aging of biological systems
described above by considering three alternative types
of element/genetic architecture [94]. In addition to the
“Parallel” model, they also presented a “Series” model
and a third type of model– a “Cascade” model, analo-
gous to the multi-stage model of disease progression in
which irreparable damage occurs in a strict sequence.
They showed that redundancy leads to actuarial se-
nescence in the Parallel and Cascade models but not
in the Series model. Finally, the authors attempted to
add evolutionary dynamics to their reliability model
and found that a population’s equilibrium redundancy
is sensitive to the environmental conditions that pre-
vailed during its evolution, such as the rate of extrin-
sic mortality [94]. For some reason, the authors did not
consider a series-parallel reliability model while only
this model is close to the real organism’s structure
where vital organs are connected in series and spe-
cialized cells in each organ are connected in parallel.
Theauthors do not attempt to fit their models to real
data, so it is difficult to check the model value.
Another model is able to explain all three mor-
tality regularities mentioned above. This is a simple
mathematical model combining the heterogeneity of
populations with an assumption that the mortality in
each subpopulation grows exponentially with age [95].
It has been proven that this model is capable of repro-
ducing the entire mortality pattern in a human popula-
tion including the observed peculiarities at early- and
late-life intervals. The authors found that the evolution
of the model parameters validates the applicability of
the compensation law of mortality to each subpopula-
tion separately. This study has indicated that the pop-
ulation structure changes so that the population tends
to become more homogeneous over time[95].
In a separate study, the authors systematically
examined the practical utility of redundancy models
for investigating the mechanisms of aging in a quanti-
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 351
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
tative manner [96]. The authors analyzed predictions
of the reliability model of homogeneous population de-
scribed in the previous section [33]. They showed that
redundancy models fit the data well and argue that this
is a strength of redundancy models over non-mecha-
nistic models because (i) when contrasting aging pat-
terns can be understood within the framework of a
single mechanistic model this indicates that the mod-
el may capture the essence of the aging process, and
(ii) redundancy parameter inference may teach us
something about the underlying mechanisms and can
as such be used to develop new hypotheses[96].
In summarizing it can be concluded that reliabil-
ity models of aging continue to be developed further
and are able to explain existing mortality regularities
outlined earlier.
CONCLUDING REMARKS
In this article, we examined various empirical
phenomena associated with the aging process and in-
troduced new insights into describing patterns of mor-
tality using reliability theory concept.
A substantial body of research on aging has yield-
ed a multitude of important and varied discoveries,
prompting the need for a unified theoretical frame-
work to consolidate this wealth of knowledge. Evolu-
tion-based theories of aging, grounded in the concept
of diminishing natural selection with advancing age,
demonstrate the practical applicability of broad theo-
retical principles in the field of aging research [97-99].
V. P. Skulachev has played a significant role in advanc-
ing evolutionary concepts. His contributions have been
instrumental in shaping contemporary perspectives
on aging, suggesting the presence of a specific pro-
gram or mechanism that influences the aging process.
Skulachev’s work has added depth to the exploration
of evolutionary theories, challenging traditional views
and paving the way for further investigations into the
molecular and genetic aspects of aging. He posited that
if aging is indeed programmed, it could be delayed, pre-
vented, or potentially reversed through interventions
that disrupt the execution of this program, similar to
our ability to intervene in cell death programs [5, 100].
His perspectives have generated numerous hypotheses
characterizing aging as a programmed process.
In this article, our aim was to offer a more com-
prehensive explanation of aging as a process of de-
terioration, extending beyond reproductive species,
through the application of the general systems fail-
ure theory known as reliability theory. This approach
aligns seamlessly with evolutionary theories, including
those related to programmed aging, when consider-
ing aging as a chronic phenoptosis. It became evident
that redundancy plays a central role in understanding
aging, particularly within a systemic framework. Sys-
tems that incorporate redundancy in essential compo-
nents inevitably undergo degradation (i.e., aging) over
time, even if constructed from elements that do not
age. Redundancy has a dual impact, enhancing dam-
age tolerance to reduce mortality and extend lifespan,
while also allowing the accumulation of damage, thus
giving rise to the aging phenomenon.
Systems with higher redundancy levels exhibit a
higher apparent aging rate or expression of aging, all
else being equal. This insight offers valuable perspec-
tive on the observation of negligible senescence in cer-
tain environments, revealing that some cases of negli-
gible senescence may be attributed to a lack of system
redundancies. For example, birds exhibit long lifespans
compared to their weight and low actuarial aging rate
(negligible senescence). However, they have relatively
high mortality at younger ages suggesting low level of
redundancy (cell reserves) [9]. On the other hand, com-
plex, redundant systems designed for greater durabili-
ty may exhibit more pronounced expressions of aging.
Throughout their lifespans, organisms deplete their
cells and reserve capacity, providing a potential ex-
planation for phenomena such as the CEM, mortality
convergence in older ages, late-life mortality deceler-
ation, leveling-off, and mortality plateaus. Organisms
appear to start their lives with an initial load of dam-
age [55, 59], and their lifespan and aging patterns can
be sensitive to early-life conditions that determine this
initial damage load during development. This concept
of early-life programming has potential implications
for interventions aimed at promoting health and lon-
gevity.
Aging is an intricate phenomenon, and adopting a
holistic approach, incorporating reliability theory, may
assist in analyzing, comprehending, and potentially
managing it. Today, gerontologists realize that a top-
down systemic approach is necessary to fully under-
stand and explain the phenomenon of aging [1].
Acknowledgments. We express our deepest grati-
tude to the late Professor Vladimir Petrovich Skulachev
(1935-2023), a distinguished Russian scientist who
served as our invaluable scientific mentor and advisor
from the 1970s onward. This article is part of a spe-
cial journal Issue dedicated to honoring his memory.
Professor Skulachev played a pivotal role in inspiring
the creation of our book, “The Biology of Life Span,”
referenced in this article, for which he served as the
scientific editor [2]. Furthermore, his encouragement
led to a significant collaborative study with him on the
variability of human life history traits [101].
Contributions. Leonid A. Gavrilov designed the
study, analyzed and interpreted results, and edited the
manuscript. Natalia S. Gavrilova conducted statistical
analyses and prepared the manuscript.
GAVRILOV, GAVRILOVA352
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
Funding. This work was partially supported by
the National Institutes of Health (project no. NIH
R21AG054849).
Ethics declarations. This work does not contain
any studies involving human and animal subjects.
The authors of this work declare that they have no
conflicts of interest.
REFERENCES
1. Cohen, A.A., Ferrucci,L., Fulop,T., Gravel,D., Hao,N.,
Kriete, A., Levine, M. E., Lipsitz, L. A., Rikkert, M.,
Rutenberg,A., Stroustrup,N., and Varadhan,R. (2022)
A complex systems approach to aging biology, Nat. Ag-
ing, 2, 580-591, doi:10.1038/s43587-022-00252-6.
2. Gavrilov, L. A., and Gavrilova, N. S. (1991) The Biology
of Life Span: A Quantitative Approach, Harwood Aca-
demic Publisher, New York.
3. Kirkwood, T. B. L. (1999) Time of Our Lives: The Science
of Human Aging, Weidenfeld & Nicolson, London, p.40.
4. Kowald,A., and Kirkwood, T.B.L. (2016) Can aging be
programmed? A critical literature review, Aging Cell,
15, 986-998, doi:10.1111/acel.12510.
5. Skulachev, M. V., and Skulachev, V. P. (2017) Pro-
grammed aging of mammals: Proof of concept and
prospects of biochemical approaches for anti- aging
therapy, Biochemistry (Moscow), 82, 1403-1422,
doi:10.1134/s000629791712001x.
6. Skulachev, V.P. (1999) Phenoptosis: programmed death
of an organism, Biochemistry (Moscow), 64, 1418-1426.
7. Skulachev, V. P. (2001) The programmed death phe-
nomena, aging, and the Samurai law of biology,
Exp. Gerontol., 36, 995-1024, doi: 10.1016/s0531-5565
(01)00109-7.
8. Lidsky, P.V., and Andino,R. (2022) Could aging evolve
as a pathogen control strategy? Trends Ecol. Evol.,
37, 1046-1057, doi:10.1016/j.tree.2022.08.003.
9. Gavrilov, L. A., and Gavrilova, N. S. (2006) Reliability
Theory of Aging and Longevity, in Handbook of the Bi-
ology of Aging (Masoro, E. J., and Austad, S. N. eds.),
6Ed., Academic Press, San Diego, pp.3-42, doi:10.1016/
B978-012088387-5/50004-2.
10. Gaillard, J.M., and Lemaitre, J.F. (2017) The Williams’
legacy: A critical reappraisal of his nine predictions
about the evolution of senescence, Evolution, 71, 2768-
2785, doi:10.1111/evo.13379.
11. Finch, C. E. (1990) Longevity, Senescence, and the
Genome, The University of Chicago Press, Chicago.
12. Olshansky, S. J., and Carnes, B. A. (1997) Ever since
Gompertz, Demography, 34, 1-15, doi:10.2307/2061656.
13. Strehler, B. L. (1978) Time, Cells, and Aging, 2nd edi-
tion, Academic Press, New York and London.
14. Greenwood,M. (1928) “Laws” of mortality from the bi-
ological point of view, J.Hyg., 28, 267-294, doi:10.1017/
S002217240000961X.
15. Kunstyr, I., and Leuenberger, H.-G. W. (1975) Geron-
tological data of C57BL/6J mice. I. Sex differences in
survival curves, J.Gerontol., 30, 157-162, doi:10.1093/
geronj/30.2.157.
16. Bronikowski, A. M., Alberts, S. C., Altmann, J., Pack-
er, C., Carey, K.D., and Tatar,M. (2002) The aging ba-
boon: comparative demography in a non-human
primate, Proc. Natl. Acad. Sci. USA, 99, 9591-9595,
doi:10.1073/pnas.142675599.
17. Bronikowski, A. M., Altmann, J., Brockman, D. K.,
Cords, M., Fedigan, L. M., Pusey, A., Stoinski, T., Mor-
ris, W. F., Strier, K. B., and Alberts, S. C. (2011) Aging
in the natural world: comparative data reveal simi-
lar mortality patterns across primates, Science, 331,
1325-1328, doi:10.1126/science.1201571.
18. Gompertz, B. (1825) On the nature of the function
expressive of the law of human mortality and on a
new mode of determining life contingencies, Phi-
los. Trans. R. Soc. Lond.A, 115, 513-585, doi: 10.1098/
rstl.1825.0026.
19. Makeham, W. M. (1860) On the law of mortality and
the construction of annuity tables, J. Inst. Actuaries,
8, 301-310, doi:10.1017/S204616580000126X.
20. Golubev, A. (2004) Does Makeham make sense?
Biogerontology, 5, 159-167, doi: 10.1023/b:bgen.
0000031153.63563.58.
21. Horiuchi,S., and Wilmoth, J.R. (1998) Deceleration in
the age pattern of mortality at older ages, Demogra-
phy, 35, 391-412, doi:10.2307/3004009.
22. Carnes, B. A., and Olshansky, S. J. (1997) A biolog-
ically motivated partitioning of mortality, Exp.
Gerontol., 32, 615-631, doi: 10.1016/s0531-5565
(97)00056-9.
23. Golubev, A. (2009) How could the Gompertz-Make-
ham law evolve, J.Theor. Biol., 258, 1-17, doi:10.1016/
j.jtbi.2009.01.009.
24. Ricklefs, R. E., and Scheuerlein, A. (2002) Biological
implications of the Weibull and Gompertz models of
aging, J.Gerontol. Ser.A Biol. Sci. Med. Sci., 57, B69-B76,
doi:10.1093/gerona/57.2.B69.
25. Vanfleteren, J.R., De Vreese,A., and Braeckman, B.P.
(1998) Two-parameter logistic and Weibull equa-
tions provide better fits to survival data from isogen-
ic populations of Caenorhabditis elegans in axenic
culture than does the Gompertz model, J. Gerontol.
Ser.A Biol. Sci. Med. Sci., 53, B393-B403, doi:10.1093/
gerona/53A.6.B393.
26. Stroustrup,N., Anthony, W.E., Nash, Z.M., Gowda,V.,
Gomez, A., Lopez-Moyado, I. F., Apfeld, J., and Fon-
tana,W. (2016) The temporal scaling of Caenorhabdi-
tis elegans ageing, Nature, 530, 103-107, doi: 10.1038/
nature16550.
27. Pakin, Y. V., and Hrisanov, S. M. (1984) Critical anal-
ysis of the applicability of the gompertz-makeham
law in human-populations, Gerontology, 30, 8-12,
doi:10.1159/000212600.
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 353
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
28. Horiuchi,S., and Wilmoth, J.R. (1997) Age patterns of
the life table aging rate for major causes of death in
Japan, 1951-1990, J.Gerontol. Ser. A Biol. Sci. Med. Sci.,
52, B67-B77, doi:10.1093/gerona/52A.1.B67.
29. Li,T., Yang, Y.C., and Anderson, J.J. (2013) Mortality
increase in late-middle and early-old age: heterogene-
ity in death processes as a new explanation, Demogra-
phy, 50, 1563-1591, doi:10.1007/s13524-013-0222-4.
30. Golubev, A. (2023) An underappreciated peculiar-
ity of late-life human mortality kinetics assessed
through the lens of a generalization of the Gompertz-
Makeham law, Biogerontology, doi: 10.1007/s10522-
023-10079-2.
31. Gavrilov, L. A., and Gavrilova, N. S. (2022) Trends
in human species-specific lifespan and actuarial
aging rate, Biochemistry (Moscow), 87, 1622-1633,
doi:10.1134/S0006297922120173.
32. Gavrilov, L. A., and Gavrilova, N. S. (2023) Actuarial
aging rates in human cohorts, Biochemistry (Moscow),
88, 1778-1785, doi:10.1134/S0006297923110093.
33. Gavrilov, L.A., and Gavrilova, N.S. (2001) The reliabil-
ity theory of aging and longevity, J. Theor. Biol., 213,
527-545, doi:10.1006/jtbi.2001.2430.
34. Strehler, B. L., and Mildvan, A. S. (1960) General
theory of mortality and aging, Science, 132, 14-21,
doi:10.1126/science.132.3418.14.
35. Gavrilov, L.A., Gavrilova, N.S., and Yaguzhinsky, L.S.
(1978) Main regularities of animal aging and death
viewed in terms of reliability theory [In Russian],
Zhurn. Obshch. Biol., 39, 734-742.
36. Golubev, A. (2019) A 2D analysis of correlations be-
tween the parameters of the Gompertz-Makeham
model (or law?) of relationships between aging, mor-
tality, and longevity, Biogerontology, 20, 799-821,
doi:10.1007/s10522-019-09828-z.
37. Shen, J., Landis, G. N., and Tower, J. (2017) Multiple
metazoan life-span interventions exhibit a sex-spe-
cific Strehler-Mildvan inverse relationship between
initial mortality rate and age-dependent mortality
rate acceleration, J.Gerontol. Ser.A Biol. Sci. Med. Sci.,
72, 44-53, doi:10.1093/gerona/glw005.
38. Tarkhov, A. E., Menshikov, L. I., and Fedichev, P. O.
(2017) Strehler-Mildvan correlation is a degenerate
manifold of Gompertz fit, J.Theor. Biol., 416, 180-189,
doi:10.1016/j.jtbi.2017.01.017.
39. Economos, A. C. (1980) Kinetics of metazoan mor-
tality, J. Soc. Biol. Struct., 3, 317-329, doi: 10.1016/
0140-1750(80)90039-1.
40. Vaupel, J. W., Carey, J. R., Christensen, K., Johnson,
T. E., Yashin, A. I., Holm, N. V., Iachine, I. A., Kannis-
to,V., Khazaeli, A.A., Liedo, P., Longo, V. D., Zeng,Y.,
Manton, K. G., and Curtsinger, J. W. (1998) Biodemo-
graphic trajectories of longevity, Science, 280, 855-860,
doi:10.1126/science.280.5365.855.
41. Greenwood,M., and Irwin, J.O. (1939) The biostatistics
of senility,
Hum. Biol., 11, 1-23.
42. Rose, M.R., Rauser, C.L., Mueller, L.D., and Benford,G.
(2006) A revolution for aging research, Biogerontolo-
gy, 7, 269-277, doi:10.1007/s10522-006-9001-6.
43. Curtsinger, J. W., Fukui, H., Townsend, D., and Vau-
pel, J.W. (1992) Demography of genotypes: failure of
the limited life-span paradigm in Drosophila melan-
ogaster, Science, 258, 461-463, doi: 10.1126/science.
1411541.
44. Economos, A.C. (1983) Rate of aging, rate of dying and
the mechanism of mortality, Arch. Gerontol. Geriatr.,
1, 3-27, doi:10.1016/0167-4943(82)90003-6.
45. Gavrilov, L. A., and Gavrilova, N. S. (2011) Mortality
measurement at advanced ages: a study of the social
security administration death master file, North Am.
Actuar. J., 15, 432-447, doi: 10.1080/10920277.2011.
10597629.
46. Gavrilova, N. S., and Gavrilov, L. A. (2015) Biode-
mography of old-age mortality in humans and ro-
dents, J. Gerontol. Ser. A Biol. Sci. Med. Sci., 70, 1-9,
doi:10.1093/gerona/glu009.
47. Dang,L., Camarda,C., Ouellette,N., Mesle,F., Robine,
J.-M., and Vallin,J. (2023) The question of the human
mortality plateau: Contrasting insights by longevity
pioneers, Demograph. Res., 48, 321-338, doi: 10.4054/
DemRes.2023.48.11.
48. Bebbington, M., Green, R., Lai, C. D., and Zitikis, R.
(2014) Beyond the Gompertz law: exploring the late-
life mortality deceleration phenomenon, Scand. Actu-
ar.J., 3, 189-207, doi:10.1080/03461238.2012.676562.
49. Feehan, D. M. (2018) Separating the signal from the
noise: evidence for deceleration in old-age death
rates, Demography, 55, 2025-2044, doi:10.1007/s13524-
018-0728-x.
50. Gavrilov, L.A., and Gavrilova, N.S. (2019) New trend
in old-age mortality: Gompertzialization of mortali-
ty trajectory, Gerontology, 65, 451-457, doi: 10.1159/
000500141.
51. Gavrilov, L. A., and Gavrilova, N. S. (2023) Living to
100 in the time of COVID-19: A study of late-life mor-
tality trajectories, in 2023 Living to 100 Compendium,
Society of Actuaries, Shaumburg, IL, p.26.
52. Economos, A. C. (1979) A non-gompertzian paradigm
for mortality kinetics of metazoan animals and fail-
ure kinetics of manufactured products, Age, 2, 74-76,
doi:10.1007/BF02432250.
53. Siler,W. (1979) Competing-risk model for animal mor-
tality, Ecology, 60, 750-757, doi:10.2307/1936612.
54. Engelman,M., Seplaki, C. L., and Varadhan, R. (2017)
A quiescent phase in human mortality? Exploring the
ages of least vulnerability, Demography, 54, 1097-1118,
doi:10.1007/s13524-017-0569-z.
55. Gavrilov, L. A., and Gavrilova, N. S. (2004) Early-life
programming of aging and longevity – The idea of
high initial damage load (the HIDL hypothesis), in
Strategies for Engineered Negligible Senescence: Why
Genuine Control of Aging May Be Foreseeable (DeGrey,
GAVRILOV, GAVRILOVA354
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
A. D. N. ed.) New York Acad. Sci., New York, pp. 496-
501, doi:10.1196/annals.1297.091.
56. Finch, C. E., and Kirkwood, T. B. L. (2000) Chance,
Development, and Aging, Oxford University Press,
NewYork, Oxford.
57. Roy, S., and Majumdar, S. (2022) Developmental noise
and stability, in Noise and Randomness in Living Sys-
tem, Springer Singapore, Singapore, pp. 119-124,
doi:10.1007/978-981-16-9583-4_12.
58. Vaux, D.L., and Korsmeyer, S.J. (1999) Cell Death in De-
velopment, Cell, 96, 245-254, doi: 10.1016/S0092-8674
(00)80564-4.
59. Kinzina, E. D., Podolskiy, D. I., Dmitriev, S. E., and
Gladyshev, V. N. (2019) Patterns of Aging Biomark-
ers, Mortality, and Damaging Mutations Illuminate
the Beginning of Aging and Causes of Early-Life
Mortality, Cell Reports, 29, 4276-4284, doi: 10.1016/
j.celrep.2019.11.091.
60. Gladyshev, V. N. (2016) Aging: progressive decline in
fitness due to the rising deleteriome adjusted by ge-
netic, environmental, and stochastic processes, Aging
Cell, 15, 594-602, doi:10.1111/acel.12480.
61. Gavrilov, L.A., and Gavrilova, N.S. (1997) Parental age
at conception and offspring longevity, Rev. Clin. Geron-
tol., 7, 5-12, doi:10.1017/S0959259897000026.
62. Gavrilov, L. A., and Gavrilova, N. S. (2015) Predic-
tors of exceptional longevity: Effects of early-life
and midlife conditions, and familial longevity, North
Am. Actuar. J., 19, 174-186, doi: 10.1080/10920277.
2015.1018390.
63. Gavrilov, L. A., Gavrilova, N. S., Kroutko, V. N., Evdo-
kushkina, G. N., Semyonova, V. G., Gavrilova, A. L.,
Lapshin, E. V., Evdokushkina, N. N., and Kushnare-
va, Y. E. (1997) Mutation load and human longevity,
Mutat. Res., 377, 61-62.
64. Doblhammer, G., and Vaupel, J. (2001) Lifespan de-
pends on month of birth, Proc. Natl. Acad. Sci. USA,
98, 2934-2939, doi:10.1073/pnas.041431898.
65. Gavrilov, L. A., and Gavrilova, N. S. (2011) Season of
birth and exceptional longevity: comparative study
of american centenarians, their siblings, and spous-
es, J. Aging Res., 2011, 104616, doi: 10.4061/2011/
104616.
66. Jarry, V., Gagnon, A., and Bourbeau, R. (2013) Mater-
nal age, birth order and other early-life factors: a
family-level approach to exploring exceptional sur-
vival, Vienna Yearbook Populat. Res., 11, 263-284,
doi:10.1553/populationyearbook2013s267.
67. Barker, D. J. P. (1998) Mothers, Babies, and Disease in
Later Life (2nd Edn), Churchill Livingstone, London.
68. Kuh, D., and Ben-Shlomo, B. (1997) A Life Course Ap-
proach to Chronic Disease Epidemiology, Oxford Uni-
versity Press, Oxford.
69. Gavrilov, L. A., and Gavrilova, N. S. (2003) Early-life
factors modulating lifespan, in Modulating Aging
and Longevity (Rattan, S. I. S. ed.) Kluwer Academic
Publishers, Dordrecht, The Netherlands, pp. 27-50,
doi:10.1007/978-94-017-0283-6_3.
70. Vaiserman, A. M. (2019) Early Life Origins of Ageing
and Longevity, Springer Cham, doi: 10.1007/978-3-
030-24958-8.
71. Leake, D. W. (2022) Tracing slow phenoptosis to the
prenatal stage in social vertebrates, Biochemistry (Mos-
cow), 87, 1512-1527, doi:10.1134/S0006297922120094.
72. Richardson, R.B., Allan, D.S., and Le,Y. (2014) Greater
organ involution in highly proliferative tissues asso-
ciated with the early onset and acceleration of age-
ing in humans, Exp. Gerontol., 55, 80-91, doi:10.1016/
j.exger.2014.03.015.
73. Tower,J. (2015) Programmed cell death in aging, Age-
ing Res. Rev., 23, 90-100, doi:10.1016/j.arr.2015.04.002.
74. Galkin,F., Zhang, B.H., Dmitriev, S.E., and Gladyshev,
V.N. (2019) Reversibility of irreversible aging, Ageing
Res. Rev., 49, 104-114, doi:10.1016/j.arr.2018.11.008.
75. Wu, H. E. (2014) Cell Death. Mechanism and Disease,
Springer New York, doi:10.1007/978-1-4614-9302-0.
76. Denic,A., Lieske, J.C., Chakkera, H.A., Poggio, E.D., Al-
exander, M.P., Singh,P., Kremers, W.K., Lerman, L.O.,
and Rule, A. D. (2016) The substantial loss of neph-
rons in healthy human kidneys with aging, J.Am. Soc.
Nephrol., 28, 313-320, doi:10.1681/asn.2016020154.
77. Fricker,M., Tolkovsky, A.M., Borutaite,V., Coleman,M.,
and Brown, G.C. (2018) Neuronal cell death, Physiol.
Rev., 98, 813-880, doi:10.1152/physrev.00011.2017.
78. Buetow, D. E. (1971) Cellular content and cellular pro-
liferation changes in the tissues and organs of the
aging mammal, in Cellular and Molecular Renewal
in the Mammalian Body (Cameron, I. L., and Thrash-
er, J. D., eds.) Academic Press, New York, pp. 87-107,
doi:10.1016/B978-0-12-156940-2.50010-5.
79. Clarke, G., Collins, R. A., Leavitt, B. R., Andrews,
D. F., Hayden, M. R., Lumsden, C. J., and McInnes,
R. R. (2000) A one-hit model of cell death in inherit-
ed neuronal degenerations, Nature, 406, 195-199,
doi:10.1038/35018098.
80. Clarke,G., and Lumsden, C.J. (2005) Scale-free neuro-
degeneration: cellular heterogeneity and the stretched
exponential kinetics of cell death, J.Theor. Biol., 233,
515-525, doi:10.1016/j.jtbi.2004.10.028.
81. Heintz, N. (2000) One-hit neuronal death, Nature,
406, 137-138, doi:10.1038/35018196.
82. Clarke, G., Lumsden, C. J., and McInnes, R. R. (2001)
Inherited neurodegenerative diseases: the one-hit
model of neurodegeneration, Hum. Mol. Genet., 10,
2269-2275, doi:10.1093/hmg/10.20.2269.
83. Holland, D., Desikan, R. S., Dale, A. M., and McEvoy,
L.K. (2012) Rates of decline in Alzheimers disease de-
crease with age, PLoS One, 7, 12, doi:10.1371/journal.
pone.0042325.
84. Beard, R. E. (1959) Note on some mathematical mor-
tality models, in The lifespan of Animals (Wolsten-
holme, E. W., and O’Connor, M. O., eds) Little, Brown
EXPLORING PATTERNS OF HUMAN MORTALITY AND AGING 355
BIOCHEMISTRY (Moscow) Vol. 89 No. 2 2024
and Company, Boston, pp. 302-311, doi: 10.1002/
9780470715253.app1.
85. Beard, R. E. (1971) Some aspects of theories of mortal-
ity, cause of death analysis, forecasting and stochastic
processes, in Biological Aspects of Demography (Brass,
W., ed.), Taylor and Francis, London, pp.57-68.
86. Vaupel, J. W., Manton, K. G., and Stallard, E. (1979)
Impact of heterogeneity in individual frailty on the
dynamics of mortality, Demography, 16, 439-454,
doi:10.2307/2061224.
87. Yashin, A. I., Vaupel, J. W., and Iachine, I. A. (1994)
Aduality in aging– the equivalence of mortality mod-
els based on radically different concepts, Mech. Ageing
Dev., 74, 1-14, doi:10.1016/0047-6374(94)90094-9.
88. Skulachev, M. V., and Skulachev, V. P. (2014) New
data on programmed aging – slow phenoptosis,
Biochemistry (Moscow), 79, 977-993, doi: 10.1134/
s0006297914100010.
89. Steinsaltz,D., and Evans, S.N. (2004) Markov mortality
models: implications of quasistationarity and varying
initial distributions, Theor. Popul. Biol., 65, 319-337,
doi:10.1016/j.tpb.2003.10.007.
90. Whittemore, K., Vera, E., Martínez-Nevado, E., San-
pera,C., and Blasco, M.A. (2019) Telomere shortening
rate predicts species life span, Proc. Natl. Acad. Sci.
USA, 116, 15122-15127, doi:10.1073/pnas.1902452116.
91. Gavrilov, L.A. (1978) Mathematical-model of aging of
animals, Dokl. Akad. Nauk SSSR, 238, 490-492.
92. Avraam, D., de Magalhaes, J.P., and Vasiev, B. (2013)
A mathematical model of mortality dynamics across
the lifespan combining heterogeneity and stochas-
tic effects, Exp. Gerontol., 48, 801-811, doi: 10.1016/
j.exger.2013.05.054.
93. Milne, E. M. G. (2008) The natural distribution of
survival, J. Theor. Biol., 255, 223-236, doi: 10.1016/
j.jtbi.2008.07.021.
94. Laird, R.A., and Sherratt, T.N. (2010) The evolution of
senescence in multi-component systems, Biosystems,
99, 130-139, doi:10.1016/j.biosystems.2009.10.008.
95. Avraam,D., Arnold,S., Jones,D., and Vasiev,B. (2014)
Time-evolution of age-dependent mortality patterns
in mathematical model of heterogeneous human
population, Exp. Gerontol., 60, 18-30, doi: 10.1016/
j.exger.2014.09.006.
96. Boonekamp, J.J., Briga,M., and Verhulst,S. (2015) The
heuristic value of redundancy models of aging, Exp.
Gerontol., 71, 95-102, doi:10.1016/j.exger.2015.09.005.
97. Charlesworth, B. (1994) Evolution in Age-structured
Populations, 2nd Edn., Cambridge University Press,
Cambridge, doi:10.1017/CBO9780511525711.
98. Charlesworth, B. (2000) Fisher, Medawar, Hamilton
and the evolution of aging, Genetics, 156, 927-931,
doi:10.1093/genetics/156.3.927.
99. Flatt, T., and Partridge, L. (2018) Horizons in the
evolution of aging, BMC Biol., 16, 13, doi: 10.1186/
s12915-018-0562-z.
100. Skulachev, V. P. (2011) Aging as a particular case of
phenoptosis, the programmed death of an organism
(A response to Kirkwood and Melov “On the pro-
grammed/non-programmed nature of ageing within
the life history”), Aging, 3, 1120-1123, doi: 10.18632/
aging.100403.
101. Gavrilova, N. S., Gavrilov, L. A., Severin, F. F., and
Skulachev, V.P. (2012) Testing predictions of the pro-
grammed and stochastic theories of aging: compari-
son of variation in age at death, menopause, and sex-
ual maturation, Biochemistry (Moscow), 77, 754-760,
doi:10.1134/S0006297912070085.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in pub-
lished maps and institutional affiliations.