Sie sind auf Seite 1von 8

Journals of Gerontology: Biological Sciences

cite as: J Gerontol A Biol Sci Med Sci, 2017, Vol. 72, No. 4, 473–480
doi:10.1093/gerona/glw248
Advance Access publication February 2, 2017

Original Article

Assessing Health Span in Caenorhabditis elegans: Lessons


From Short-Lived Mutants
Jarod A.  Rollins,1 Amber C.  Howard,2 Sarah K.  Dobbins,3 Elsie H.  Washburn,4 and
Aric N. Rogers1
1
Davis Center for Regenerative Biology and Medicine, Mount Desert Island Biological Laboratory, Bar Harbor, Maine. 2College of Arts
and Sciences, University of Maine at Augusta. 3Department of Engineering, Stanford University, California. 4College of Math and Science,
California Polytechnic University, San Luis Obispo.

Address correspondence to Aric Rogers, PhD, Davis Center for Regenerative Biology and Medicine, Mount Desert Island Biological Laboratory,
159 Old Bar Harbor Road, Bar Harbor, ME 04609. E-mail: arogers@mdibl.org

Received August 16, 2016; Editorial Decision Date November 16, 2016

Decision Editor: Rafael de Cabo, PhD

Abstract
Genetic changes resulting in increased life span are often positively associated with enhanced stress resistance and somatic maintenance.
A  recent study found that certain long-lived Caenorhabditis elegans mutants spent a decreased proportion of total life in a healthy state
compared with controls, raising concerns about how the relationship between health and longevity is assessed. We evaluated seven markers of
health and two health-span models for their suitability in assessing age-associated health in invertebrates using C elegans strains not expected
to outperform wild-type animals. Additionally, we used an empirical method to determine the transition point into failing health based on the
greatest rate of change with age for each marker. As expected, animals with mutations causing sickness or accelerated aging had reduced health
span when compared chronologically to wild-type animals. Physiological health span, the proportion of total life spent healthy, was reduced
for locomotion markers in chronically ill mutants, but, surprisingly, was extended for thermotolerance. In contrast, all short-lived mutants
had reduced “quality-of-life” in another model recently employed for assessing invertebrate health. Results suggest that the interpretation of
physiological health span is not straightforward, possibly because it factors out time and thus does not account for the added cost of extrinsic
forces on longer-lived strains.

Keywords: Health—Life-span measurement—Phenotype—Sarcopenia—Invertebrate

An assumption underlying the desire for increased life span is that Although many of the studies linking enhanced longevity with
it will also improve quality of life by helping ameliorate the impact increased stress resistance have come from investigations utilizing
of diseases and diminution of health associated with aging. This Caenorhabditis elegans model system, a recent study provided evi-
possibility is generally supported by evidence from model organ- dence suggesting that the actual percent of life spent in a frail state,
isms showing that enhanced somatic maintenance, frequently meas- referred to as physiological gerospan, was increased among long-
ured by increased survival under various forms of stress, coincides lived C elegans mutants compared with wild-type animals (9). In their
with increased longevity (1–3). Other measurements of health have investigation, they examined several markers of health that changed
shown that prolongevity interventions maintain physiology in a with age in well-characterized long-lived mutants and compared
youthful state, including preservation of locomotion, feeding behav- them with wild-type animals. While generally confirming enhanced
ior, and accumulation of autofluorescent pigments in cells (4–6). chronological health span, their results suggested that the proportion
Furthermore, when genetic models of longevity are combined with of total life spent healthy in long-lived mutants, termed physiological
age-associated models of disease, results often indicate amelioration health span, was shorter than that of wild-type animals. A later study
or delay in the onset of pathology (7,8). Together, the evidence indi- demonstrated that the reduced physiological health span of long-
cates that health span can be extended through evolutionarily con- lived daf-2 mutants was a result of altered foraging behavior caused
served genes and pathways. by tracking unstimulated locomotion in the presence of food (10).

© The Author 2017. Published by Oxford University Press on behalf of The Gerontological Society of America. All rights reserved.
473
For permissions, please e-mail: journals.permissions@oup.com.
474 Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4

However, the reduced physiological health span of other long-lived of extrinsic forces on aging, which is not taken into consideration
mutants remains unexplained. Additionally, assessment of health when time is factored out.
span in mouse using markers for body composition, energetics, and
movement showed little association with mortality suggesting an Materials and Methods
uncoupling of health span and life span in more complex systems
Nematode Culture and Strains
(11). These provocative studies have raised concerns about the way
in which we examine health in animal systems and the applicabil- C elegans strains were cultured at 20°C and maintained on nema-
ity of such results to human health. The importance of considering tode growth media (NGM) plates seeded with the Escherichia coli
health span in the context of life span was made clear in the recent strain OP50. N2 wild-type, CF1038 daf-16(mu86) I, XA8203 sir-
convening of the Geoscience Network, which deemed extension of 2.1(ok434) IV, PS3551 hsf-1(sy441) I, and TK22 mev-1(kn1) III
life span alone as insufficient evidence of delayed aging (12). were obtained from the Caenorhabditis Genetics Center (CGC). All
Obstacles to the characterization of invertebrate health span strains were permitted to grow at least three generations under nor-
include the choice of which strain and markers to use. For the former, mal, unstressed conditions prior to use in experiments. Except for
we used short-lived C elegans mutants displaying signs of accelerated thermotolerance assays, all experiments were carried out at 20°C.
aging and/or chronic illness, allowing assessment of candidate health
markers and recently reported models that assess age-associated health. Life Spans
For the latter, we wanted to make a distinction between health markers Survival was scored starting from the first day of adulthood using 100
strongly dependent on biological aging and those which only have a or more worms per strain divided among three seeded NGM plates
temporal relationship or are weakly dependent on the aging process. containing 200 µM fluorodeoxyuridine (FUDR) to prevent contami-
Given that our chosen short-lived strains displayed signs of chronic nation by progeny. Survival status was determined by touch provo-
illness, we adopted definitions from Brody and Schneider (13) used for cation. Worms with vulval protrusion were censored from the assay.
chronic human diseases to make this distinction for health markers. In Differences in survival curves were calculated using the log-rank test
their manuscript, only disorders with mortality rates that increase with in the Graphpad Prism 6.0 software. Maximum life span was calcu-
age were defined as age dependent. Applying a similar logic to the pre- lated by taking the average life span of the top 10 longest-lived worms.
sent study, markers which failed sooner chronologically in short-lived
mutants were defined to be aging dependent, as the earlier failing of
Development and Fecundity
both health and longevity is presumably regulated by similar processes.
Several dozen gravid adults were allowed to lay eggs on spotted
Conversely, markers were considered time dependent if they changed
NGM plates for 1 hour and removed. Time to egg-laying adulthood
with age but not in a manner consistent with life span.
was started from this point. Eggs were allowed to hatch and 20–22
For short-lived strains, we used animals with mutations in daf-
animals were randomly picked and separated onto individual plates.
16 (abnormal dauer formation), mev-1 (abnormal methyl viologen
Plates bearing single young adults were checked every hour for the
sensitivity), hsf-1 (heat shock factor), and sir-2.1 (silent informa-
presence of eggs. Animals from the development assay were trans-
tion regulator). daf-16 encodes a FOXO transcription factor that
ferred to fresh plates every 24 hours following initiation of egg lay-
acts in the insulin/insulin-like growth factor signaling (IIS) pathway
ing. Hatched progeny and unfertilized oocytes were counted for each
and is required for increased life span and enhanced resistance to
plate 48 hours after removal of the egg-laying adult.
stress when the IIS pathway is attenuated (14). mev-1 encodes the C
elegans ortholog for human mitochondrial succinate dehydrogenase
cytochrome b560 subunit, the mutation of which is associated with Health-Span Assays
significantly decreased life span, increased reactive oxygen species Speed, bending angle, and worm size were assayed for each time
(15), and severely impaired reproduction (16). hsf-1 encodes the point by sampling from a population of age-synchronized animals
heat shock transcription factor important for mediating the response on seeded NGM plates. Worms were tracked in real time after gently
to unfolded proteins in the cytoplasm and signals a series of pro- tapping plates to stimulate movement and distinguish the ability to
longevity genes (7,17). Its expression is required for life span and move from food-seeking behavior (10). Forward and backward speed
stress-resistance phenotypes associated with attenuated IIS (17) and were recorded with a Stingray F504B ASG digital (Allied Technologies
loss-of-function leads to defects in development and reproduction GmbH) camera equipped with an AF Micro-Nikkor 60mm f/2.8D lens
(18,19). sir-2.1 encodes a NAD-dependent deacetylase important for (Nikon). 30-second intervals of continuous (unbroken) tracking were
metabolic homeostasis, with lost function resulting in shortened life analyzed using Wormlab ver 3.0 (MBF Bioscience). For pharyngeal
span and increased sensitivity to stress (20). pumping, worms were observed individually with a Leica M165FC
Using mutants for these genes, we determined which previ- stereo microscope and digitally recorded (VLC, ver. 2.1.3) on seeded
ously used markers of health span were aging dependent and thus NGM plates. Pumps were counted manually by playing 30 seconds of
informative candidates for health-span determination linked to life video at a reduced speed. For autofluorescence measurements, worms
span. Additionally, we found maximum bending amplitude, which is were transferred to 1% agar pads containing 20-mM sodium azide
an indicator of muscle integrity (21), to be a new aging-dependent spotted on glass slides. Autofluorescence was measured on a Leica
health-span marker with possible associations with longevity. To M165FC stereo microscope in both the green fluorescent protein
determine when the healthy period of life ends, we posited an empiri- (GFP) (GFP narrow band filter set, excitation ET470/40 nm, emission
cal method based on the rate of change for a given marker. Although ET510/10  nm) and the 4′,6-diamidino-2-phenylindole (DAPI) (ET
chronological health span was reduced in the chosen short-lived DAPI filter set, excitation AT350/50x nm, emission ET460/50m) chan-
mutants, physiological (time normalized) health span was actually nels using a 3-second exposure. Quantification of autofluorescence
greater for thermotolerance in short-lived strains in this model. was performed using ImageJ for mean pixel intensity and corrected
Although the apparent simplicity of the physiological health-span for background fluorescence. Thermotolerance assays were conducted
model suggests a straightforward interpretation, these unexpected with 100 worms (aged as indicated) and distributed among two plates
results indicate additional consideration is required, such as the cost at 35°C. Survival was measured every hour by touch provocation.
Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4 475

Resistance to oxidative stress was tested using worms distributed to


individual wells (12 worms/well, 4 wells/strain) of a 24-well culture
plate containing 7.5-mM hydrogen peroxide in S basal buffer. Survival
was measured every other hour by touch provocation. All health-span
assays were independently repeated three times.

Statistical Analysis
All statistics other than log-rank tests were performed in the soft-
ware package R (ver 3.1.3). For each health marker, analysis of
variance was used to estimate effects of genotype, time, and the
interaction of the two. Replicate experiment was used as a block-
ing factor. Post hoc Tukey’s honest significant difference was used
to compare means between genotypes and time points. The rate of
change in each aging-dependent health marker was modeled using
linear regression. The linear regression estimated the effect of time
with replicate experiment as a blocking factor. As maximum speed
and maximum amplitude did not decrease until Day 5 for most gen- Figure 1.  Short-lived Caenorhabditis elegans mutants demonstrate deficiencies
in health during development and adulthood. (A) Representative survival
otypes, Day 1 was omitted from the linear regression analysis for
curves of N2 wild-type and short-lived mutants from four biological repeats
these parameters. For similar reasons, heat stress survival on Day 2 (Supplementary Table  1). Survival curves were compared using Mantel-Cox
was omitted from the regression analysis. To determine the onset of log-rank tests and all short-lived mutants had significantly (p < .05) shorter
rapid health decline in wild type, linear regression was used as above median life spans than wild-type N2 (mean survival difference for daf-16(mu86)
but with a sliding window of 3 consecutive time points. The time by was −33%, mev-1(kn1) was −29%, hsf-1(sy441) was −37%, and sir-2.1(ok434)
which each genotype entered gerospan was imputed using the regres- was −26%). (B) Mean survival at 35°C on Day 6 of adulthood was reduced
in daf-16, mev-1 and hsf-1, but not in sir-2.1 (p = 8.0E-8, 2.4E-11, 2.4E-11, and
sion function from the linear models. Because measurements were
0.79, respectively, calculated using Tukey’s test). Data were pooled from three
not taken on Day 19 for thermotolerance, the missing values were
independent experiments, n  =  300/strain. (C) Mean survival was reduced in
extrapolated for each genotype using the linear regression from the all short-lived strains in response to 7.5-µM hydrogen peroxide at Day 5 of
other time points. Cumulative quality-adjusted survival was deter- adulthood (p < .0001, Tukey’s test). Data were pooled from three independent
mined for each genotype as described by Hahm and colleagues (10) experiments, n = 138/strain. (D) The size of selected strains based on imaged
by first calculating the normalized health metric score and then the area on Day 5 of adulthood showed that only mev-1 were smaller compared
with wild type (p = 2.8E-13, Tukey’s test). Data were pooled from 3 independent
quality-adjusted survival for each aging-dependent trait.
experiments, n = 45/strain. Data were pooled from 3 independent experiments.
E) Time to development for each strain as measured from time to egg laying
to first day of adulthood. Compared with wild type, mev-1 and hsf-1 had a
Results
longer development time while daf-16 and sir-2.1 were similar (p  =  1E-16,
Reduced Whole-Life Health and Accelerated Aging p = 1.0E-16, p = .088 and p =.049, respectively Tukey’s test). Data were pooled
from 4 independent experiments, n = 70/strain. F) Fecundity of each selected
in Short-Lived Mutants
strain as measured by the number of eggs hatched. The mutants mev-1 and
Testing was carried out to confirm the short-lived and impaired sir-2.1 had reduced hatched progeny compared with wild type (p  =  1.0E-
stress response status of daf-16(mu86), sir-2.1(ok434), hsf-1(sy441), 12 and 1.4E-12, respectively in Tukey’s test). Data were pooled from four
and mev-1(kn1) mutants. Life-span survival curves are shown in independent experiments, n = 64/strain. Error bars indicate SEM. ***p < .001.
Figure  1A and median life span ranged from 26% to 38% below
that of N2 wild type (17.0  days for sir-2.1(ok434), 16.5  days for
mev-1(kn1), 15.5  days for daf-16(mu86), and 14.3  days for hsf- were produced in mev-1 (129 compared with 259 in wild type,
1(sy441)). Information pertaining to statistics and replicates are p = 1.02E-12, Tukey’s test) and sir-2.1 (199, p = 1.44E-12) mutants
reported in Supplementary Table 1. Survival under heat or oxidative (Figure 1F). Interestingly, a significant number of unfertilized oocytes
stress was also reduced in these strains (Figure 1B and C), indicating were laid in the daf-16 (78.5, p = 1.03E-12, Tukey’s test) and hsf-1
defects in somatic maintenance. (41.7, p = 1.07E-12) backgrounds compared with wild-type N2 (0.7;
In long-lived strains, increased life span and resistance to Supplementary Figure  1) near the end of the reproductive period,
stress often coincide with tradeoffs in growth and reproduction. indicative of dysregulated ovulation. However, there was not a sig-
Such defects in short-lived strains would indicate illness at all life nificant impediment to the production of viable progeny in daf-16
stages. Growth was assessed by analysis of nematode size quanti- (254, p  =  .95) and hsf-1 (246, p  =  .34) compared with wild type
fied on Day 5 of adulthood, by which time all strains were com- (259) in the conditions tested (Figure  1F). Finally, the bulk distri-
pletely grown. Only mev-1 mutants were significantly smaller bution of reproduction in sir-2.1 animals was shifted to after the
than wild-type mutants (106.2  mm2 vs 154.7  mm2, p  =  2.87E-13, second day of egg laying, unlike in the wild-type background where
Tukey’s test; Figure  1D). Development was assessed by measur- more than 50% of viable progeny were laid within the first 48 hours
ing the time required for each strain to develop from an egg to an of the onset of egg laying (Supplementary Figure 2).
egg-laying adult (Figure  1E). Development was delayed in hsf-1 Together, the data for longevity, stress resistance, development,
(92.8 h compared with 89.2 h in wild type, p = 6.85E-06, Tukey’s and reproduction suggested that the hsf-1, mev-1, and sir-2.1 mutants
test) and severely delayed in mev-1 (112.5 h, p = 1.0E-16, Tukey’s used here model sickness due to their deficits in early- and late-life
test). Reproductive health was assessed by counting hatched progeny metrics of health. While the high number of unfertilized oocytes in
(Figure  1F), and unfertilized oocytes (Supplementary Figure  1), as daf-16 animals suggests anomalous regulation of ovulation, total
well as by measuring the distribution and duration of reproduction fecundity was not significantly reduced. Conversely, daf-2 mutants
(Supplementary Figure 2). Significantly reduced number of progeny produce fewer unfertilized oocytes compared with wild type (22),
476 Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4

presumably due to increased daf-16 activity. Thus, with the absence reduced compared with wild type, in accordance with an aging-
of major developmental or reproductive deficiencies, daf-16 mutants dependent marker. To our knowledge, this is the first characterization
may be considered a model of accelerated aging, although lack of of maximum bending amplitude with age in C elegans on solid media.
functional DAF-16 means these animals have chronically reduced Thus, locomotion markers in Figure  2A–C are all aging dependent,
stress resistance. With these characterizations in mind, we analyzed making them strong candidates for health-related diagnostics of aging.
the lifelong performance of health markers concerning motility, feed- Heat stress can disrupt cellular homeostasis though protein mis-
ing, autofluorescence, and stress resistance. folding. Thermotolerance decreased with age in all strains tested,
and short-lived mutants had significantly reduced thermotolerance
Locomotion and Thermotolerance Are compared with wild-type worms. The strain-specific reduction
Aging-Dependent Markers of Health was greatest for hsf-1, which was 34% (p  =  3.0E-09), followed
by 29% for mev-1 (p = 1.3E-06), 28% for sir-2.1 (p = 4.1E-06),
Locomotion is dependent on the integrity of muscle mass, connective
and 24% for daf-16 (p = 2.6E-04) on Day 13 (Figure 2D; Tukey’s
tissue, and neuronal signaling (21,23–26). We investigated changes
test). Results indicate that thermotolerance is an aging-dependent
in movement with age using digital video recording and nematode
marker.
tracking software (Wormlab 3.1, see Methods). Mean and maxi-
mum speed declined over time in all strains tested (Figure 2A and B).
Short-lived mutants were typically slower than wild type for both Autofluorescence and Pharyngeal Pumping Are
mean and maximum speed. By Day 5 of adulthood, the maximum Time-Dependent Markers of Health
speeds of daf-16, sir-2.1, mev-1, and hsf-1 were slower than wild Both DAPI and GFP channel autofluorescence have been used as
type by 30%, 32%, 48%, and 51%, respectively. Differences in aging metrics in C elegans (27,28). To better understand the relation-
maximum speed increased by Day 15 (slower by 60%, 70%, 59%, ship between autofluorescence and age, we measured DAPI chan-
and 58%, respectively). The decline in mean speed began after Day nel (350/50ex,450/50em) and GFP channel (480/20ex, 510/20em)
1 for the short-lived mutants and after Day 5 for wild type, with autofluorescence throughout adulthood. DAPI channel fluorescence
similar results for maximum speed except that the decline began was seen to increase with age in N2 wild-type, mev-1 and hsf-1
after Day 5 for daf-16 and hsf-1. As mean and maximum speed both strains (Figure  3A). Surprisingly, daf-16 DAPI autofluorescence
declined with age sooner in all short-lived mutants compared with stopped increasing after Day 12, whereas sir-2.1 remained steady
wild-type animals, we categorized these traits as aging dependent. throughout life. Although GFP channel autofluorescence increased
Of the markers tested in the present study, maximum speed had the with time for all strains, values in daf-16 and sir-2.1 mutants were
highest correlation with survival rate (Table 1). never greater than wild-type and were significantly lower by Day 19
Locomotion is also characterized by maximum bending ampli- (p = .039 and .002 respectively, Tukey’s test, Figure 3B). Conversely,
tude, which depends on thick filament organizing centers that trans- mev-1 mutants displayed significantly higher GFP autofluorescence
mit the force of muscle contraction into locomotion (21). Maximum by Day 19 (p = .0002). hsf-1 mutants had significantly higher GFP
bending amplitude decreased after Day 5 in all strains (Figure 2C). By autofluorescence on Day 12 of adulthood (p = .006) but not on Day
Day 12 and beyond, values for all short-lived strains were significantly 15 (p = .99). Based on results for the strains used, DAPI and GFP

Figure 3.  Autofluorescence and pharyngeal pumping were not altered sooner


in short-lived mutants. (A) DAPI channel autofluorescence, (B) GFP channel
autofluorescence, and (C) pharyngeal pumping of wild-type and short-lived
Figure  2.  Locomotion and thermotolerance were reduced sooner mutants throughout life. Insufficient numbers of hsf-1 mutants survived to
chronologically in short-lived mutants. Mean and maximum speed, Day 19 to take complete measurements for that time point. Pumping was
maximum bending amplitude, and thermotolerance throughout life are not significantly different at any time point for any strain compared with wild
shown in A–D. Error bars indicate SEM. Data were pooled from three type. Error bars indicate SEM. Data were pooled from three independent
independent experiments. n = 45/strain for locomotion parameters. n = 300/ experiments. n = 45/strain for autofluorescence measurements, n = 200/strain
strain for thermotolerance. *p < .05, ***p < .001, as compared with N2 wild for pharyngeal pumping. *p < .05, **p < .01, ***p < .001, as compared with
type on the same day by analysis of variance with post hoc Tukey’s test. wild type on the same day by analysis of variance with post hoc Tukey’s test.
Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4 477

channel autofluorescence are supported as time-dependent markers were obtained for mean speed (Figure  4A). The onset of gerospan
of health. for maximum bending amplitude was 9.6  ±  0.5  days for daf-16,
The ability of the worms to feed is based on the pumping rate of 4.5 ± 0.9 days for mev-1, 7.6 ± 0.5 days for hsf-1, and 9.4 ± 0.4 days
their pharynx, which they modulate according to availability of food for sir-2.1, earlier than wild type (18.6 ± 2.0 days) (Figure 4C). The
(29) and which has been used as a marker for health in C elegans onset of gero span for thermotolerance (Figure  4D) was also ear-
(4,5,27). Pumping rates were measured on Days 1, 5, 8, 12, and 15, lier than wild type (15.3  ±  0.2  days) in daf-16 (13.3  ±  0.1  days),
after which no appreciable pumping was detected. Although pump- mev-1 (13.5  ±  0.1  days), hsf-1 (12.1  ±  0.3  days), and sir-2.1
ing declined over time (Figure 3C), rates were similar or higher in (14.0 ± 0.2 days). Results were in line with expectations for intrinsi-
short-lived mutants compared with wild type, indicating that pump- cally health-deficient animals.
ing is time dependent. This time dependency was also observed
for long-lived mutants (9) where pumping was not detected after Physiological Comparisons Reveal That Health Span
15 days. Is Nonproportional With Life Span
We calculated the physiological health span for each aging-
Chronological Health Span Is Reduced in dependent marker in Figure  4E–H. Health span for maximum
Short-Lived Strains speed was reduced in daf-16 (40.9 ± 1.6%), sir-2.1 (25.8. ± 2.2%),
To further investigate the relationship between health span and life mev-1 (23.6. ± 2.5%), and hsf-1 (41.1  ±  1.2%) compared with
span, we calculated the amount of time spent healthy in wild-type wild type (49.7 ± 1.3%; Figure 4F), with similar results for mean
animals for each marker. Because candidate aging markers do not speed (Figure  4E). Physiological health span for maximum bend-
always decrease (eg, autofluorescence; Figure 3) nor change by more ing amplitude was also reduced in daf-16 (48.5  ±  2.4%), hsf-1
than 50% (maximum bending angle; Figure 2), we implemented a (20.8 ± 4.0%), and sir-2.1 (44.8 ± 1.8%) as compared with wild
method that uses the most dynamic period of change taking place type (35.2  ±  1.4%; Figure  4G). Surprisingly, physiological health
in a wild-type animal to determine when health falters for a given span for thermotolerance was extended in all of the short-lived
marker. Thus, this method is customized to the natural transition for mutants (daf-16: 60.8 ± 0.5%, mev-1: 55.0 ± 0.4%, hsf-1: 55.6% ±
the health marker being used. This was determined by measuring 0.5%, and sir-2.1: 58.3  ±  0.4%) compared with wild type
rates of change over a sliding window of three time points, with the (47.0 ± 0.4%; Figure 4H). As the observed maximum life span of a
last time point in the window that demonstrates the steepest slope strain may be dependent on the size of the population, we addition-
indicating the first time point of the geriatric state (see Methods). ally calculated physiological health span for each strain normalized
Based on this method, which we refer to as Sliding Window Analysis by their median life span (Supplementary Figure 5). However, no
of Vigor, the onset of gerospan for aging-dependent markers in qualitative differences in the physiological health span were evi-
N2 wild-type animals was Day 12 for speed and maximum speed, dent between the two methods. These results make interpreting
Day 13 for thermotolerance, and Day 19 for maximum bending previously reported decreases in physiological health span markers
amplitude (Table 2; Supplementary Figure 3). With these transition in long-lived mutants (9) challenging and suggest that the apparent
points, the percent changes from peak values at the time of gero- simplicity of health-span metrics normalized to remove time may
span onset were near middle- to late-middle age for average speed, be misleading.
maximum speed, and thermotolerance (36.3%, 48.3%, and 43.6%, For comparison, we also evaluated strains using a cumula-
respectively; Table 2). However, the onset of gerospan for maximum tive quality-adjusted survival metric (10,30) (Figure 4I–L). In this
bending amplitude was 72.8%, indicating that it is the only aging- method, survival rate and health performance at each time point
dependent marker with a transition point that occurs in late life, are first normalized to the maximum value from wild type and
closer to the time humans are typically characterized as entering then multiplied together to yield a survival curve adjusted by the
a geriatric state. In order to determine analogous transitions into quality of life experienced for each strain. The area under the qual-
the geriatric state in the short-lived strains, interpolation using lin- ity-adjusted survival curve between each consecutive time point is
ear regression models (Supplementary Figure 4 and Supplementary then calculated and cumulatively summed to provide a diagnostic
Table 2) was carried out based on wild-type transition values. of both total quality and duration of life. Based on this analysis,
The transition to chronological gerospan using aging-depend- wild-type worms have the highest quality of life for all aging-
ent markers, based on interpolation of linear regression data, dependent markers. A similar reduced total quality of life for daf-
occurred earlier in short-lived mutants and is presented graphically 16 was reported for maximum speed, as compared with wild type
in Figure  4A–D. The transition for daf-16, mev-1, hsf-1, and sir- (10). Unlike physiological health span, the quality-adjusted metric
2.1 for maximum speed was 8.1  ±  0.3, 5.6  ±  0.5, 4.6  ±  0.5, and does not factor out time, which may be important if extrinsic fac-
8.7 ± 0.2 days, respectively, which were all reduced compared with tors that influence aging play a significant role in health among
wild type (13.4 ± 0.3 days; Figure 4B). Qualitatively similar results differently lived strains.

Table 1.  Survival Rate Is Highly Correlated With Maximum Speed and Well Correlated With Other Health Markers. Spearman Correlations
Between Survival Rate and Each Health-Span Parameter

Speed Max Speed Max Amp Pumping DAPI GFP Thermotolerance

Coefficient 0.88 0.95 0.69 0.89 −0.74 −0.37 0.89


p Value 9.7E-11 1.3E-15 2.5E-05 7.3E-11 3.7E-06 0.045 7.9E-11

Note: DAPI = DAPI channel autofluorescence; GFP = GFP channel autofluorescence; Max Amp = maximum bending amplitude; Max Speed = maximum move-
ment speed; Pumping = pharyngeal pumping; Speed = mean movement speed.
478 Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4

Table  2. Empirical Determination of the Geriatric State for from movement speed, as C elegans mutants deficient in components
Individual Health Markers of force-transmitting sarcomeres exhibited differences in bending
amplitude but not locomotion (21). Out of the aging-dependent
Marker Day Mean Max % Max
markers, bending amplitude was found to fail latest in life. A previ-
Speed (µm/s) 12 75.2 207.3 36.3 ous characterization of the locomotion of C elegans while swimming
Max speed (µm/s) 12 188.7 390.2 48.3 reported an increase in the maximum difference in curvature with
Max amplitude (µm) 19 150.6 206.9 72.8 age. These two opposing results suggest that bending amplitude on
Thermotolerance (hours) 13 2.3 5.3 43.6 solid media and curvature while swimming measure different aspects
of health and behavior. Supporting this notion, the kinematics of
Note: % Max  =  the percentage of the maximal value when geriatric; crawling has been shown to be distinct from that of swimming and
Day = day of adulthood geriatric onset for wild-type N2 worms; Max = max- is characterized by a greater amplitude and smaller frequency (32).
imum value each marker observed in wild-type over life span; Max
Finally, although the pharyngeal muscle may show variable deterio-
Amp = maximum bending amplitude (µm); Max Speed = maximum movement
ration among strains (not tested), cessation of pumping was largely
speed (µm/s); Mean = mean value for each marker when N2 becomes geriatric;
invariant among wild-type and short-lived strains, similar to findings
Speed = mean movement speed (µm/s).
in long-lived mutants (9), suggesting that this marker is related to
longevity mainly by virtue of being dependent on the passage of time.
Cellular autofluorescence, a second category with the poten-
tial to be a diagnostic of aging, can occur due to accumulation of
endogenous fluorescent particles such as lipofuscin, advanced glyca-
tion end products, flavins, dihydronicotinamide-adenine dinucleo-
tide phosphate, elastin, and collagen (6,33), each with distinct but
sometimes overlapping spectral excitation and emission properties.
Of these particles, the accumulation of lipofuscin, the oxidized bi-
products of lysosomal degradation, has been a focus in mammals
as a marker of aging (34). In worms, lipofuscin accumulation has
been proposed as a marker of aging by measuring changes in auto-
fluorescence using a DAPI channel filter (27). However, Coburn and
colleagues (28) demonstrated that the presence of DAPI channel
fluorescence did not increase with induced oxidative stress nor with
age in C elegans. Instead, they showed that this autofluorescence
occurs as a burst upon death and concluded that the signal was due
to anthranilate acid esters originating from gut granules. Although
that study and the present study were able to detect significant
increases in GFP channel autofluorescence with age, they were not
Figure  4.  Physiological health span is decreased for locomotion markers and
consistently greater in short-lived strains and so failed the criteria to
increased for thermotolerance in short-lived mutants. Chronological health be considered an aging-dependent marker. However, it is important
span and gerospan for movement speed, maximum speed, maximum bending to note that there are some differences in the way in which channel-
amplitude, and thermotolerance markers were calculated in A–D, respectively. specific fluorescence is measured (eg, see DAPI filter settings in (6,35)
(E,H) Same as in A–D, but for physiological health span and gerospan. Health compared with (9,28), and ones used in the present study). Thus,
span and gerospan are indicated by lighter and darker hues, respectively. Error
autofluorescence cannot be ruled out as a possible aging-dependent
bars indicate SEM for the prediction of the onset of gero span. The cumulative
marker, but will require standardization in approach and further
quality-adjusted survival, at metric sensitive to both the total quality and duration
of life was calculated for speed, maximum speed, maximum amplitude, and characterization of what is actually measured at specific wavelengths
thermotolerance in I–L. in C elegans.
The third category of health used to investigate aspects of aging
in C elegans involves proteostasis and the ability to maintain somatic
Discussion
function, measured by survival, in the presence of acute stress.
Evaluation of Health-Span Markers Mechanisms contributing to proteostasis are necessary for survival
Three major categories of health are frequently used to investigate under heat stress (36) and decline with age (37). On the first day of
aspects of aging in C elegans. They include muscle function/integ- adulthood, proteostatic maintenance undergoes a dramatic transi-
rity, cellular accumulation of autofluorescent pigments, and resist- tion in C elegans that can be measured in hours (38,39). As one of
ance to acute perturbation of homeostasis, particularly proteostasis. the most robust aging-dependent markers, we find a second transi-
In the clinic, muscle function and integrity with age are judged by tion that occurs in midlife for wild type that is delayed in short-lived
muscle wasting, or sarcopenia, which is diagnosed according to gait strains relative to their life span. Although the result is surprising, the
speed, grip strength, and muscle mass (31). In C elegans, movement ability to maintain proteostasis may be more influenced by extrinsic
speed is akin to human gait speed and maximum movement speed aging factors than other markers.
has been suggested to be the one of the most informative metrics of
health in C elegans due to its high correlation with longevity and Considering a Role for Extrinsic Factors in
ability to predict remaining life span (10). In agreement with this, Health Span
maximum speed was the highest correlated with survival out of the Mortality is determined by a combination of intrinsic and extrin-
markers tested in the present study. We also considered maximum sic factors. Intrinsic aging is thought to be dependent on somatic
bending amplitude as a marker akin to grip strength that is distinct maintenance and senescence, whereas extrinsic aging is driven by
Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4 479

environmental conditions (40). Although intrinsic and extrinsic fac- is funded by the NIH Office of Research Infrastructure Programs (P40
tors can affect longevity and health, the relative contribution on each OD010440).
is still largely unknown. Our assessment of aging-dependent health
parameters included the binary separation of health span and gero- References
span based on the physiological age of each strain. When comparing
1. Hansen M, Taubert S, Crawford D, Libina N, Lee SJ, Kenyon C. Lifespan
health span physiologically, time is factored out, but this discounts extension by conditions that inhibit translation in Caenorhabditis elegans.
the impact of extrinsic factors in longer-lived animals, which in the Aging Cell. 2007;6:95–110. doi:10.1111/j.1474-9726.2006.00267.x
current study is the N2 wild-type strain. For example, analysis of 2. Lithgow GJ, White TM, Melov S, Johnson TE. Thermotolerance and
thermotolerance indicated physiologically extended health span in extended life-span conferred by single-gene mutations and induced by ther-
all short-lived mutants compared with wild type (Figure 4H). This mal stress. Proc Natl Acad Sci USA. 1995;92:7540–7544. doi:10.1073/
disproportionality between health span and life span could be due to pnas.92.16.7540
a requirement of the passage of time to allow extrinsic factors, like 3. Epel ES, Lithgow GJ. Stress biology and aging mechanisms: toward under-
standing the deep connection between adaptation to stress and longevity.
background radiation (41), to damage proteins to the point where
J Gerontol A Biol Sci Med Sci. 2014;69(suppl 1):S10–S16. doi:10.1093/
thermotolerance is impacted. Thus, although physiological compari-
gerona/glu055
sons have limited use in accessing the overall quality of life a particu-
4. Huang C, Xiong C, Kornfeld K. Measurements of age-related changes
lar intervention may provide, they may be beneficial in estimating of physiological processes that predict lifespan of Caenorhabditis ele-
the susceptibility of a particular health marker to extrinsic forces. gans. Proc Natl Acad Sci USA. 2004;101:8084–8089. doi:10.1073/
pnas.0400848101
The Relative Importance of Life Span and 5. Lee GD, Wilson MA, Zhu M, et  al. Dietary deprivation extends
lifespan in Caenorhabditis elegans. Aging Cell. 2006;5:515–524.
Health Span
doi:10.1111/j.1474-9726.2006.00241.x
The question remains, is the disproportionality between life span
6. Gerstbrein B, Stamatas G, Kollias N, Driscoll M. In vivo spectrofluorim-
and health span cause for concern in the translation of aging etry reveals endogenous biomarkers that report healthspan and dietary
research? On one hand, the extended period of frailty observed restriction in Caenorhabditis elegans. Aging Cell. 2005;4:127–137.
in C elegans long-lived mutants could have “disastrous economic doi:10.1111/j.1474-9726.2005.00153.x
and social consequences if applied to humans” (9). On the other 7. Cohen E, Bieschke J, Perciavalle RM, Kelly JW, Dillin A. Opposing activi-
hand, we could apply reduction ad absurdum to this notion with ties protect against age-onset proteotoxicity. Science. 2006;313:1604–
our results from short-lived mutants. If longer life span without a 1610. doi:10.1126/science.1124646
proportional increase in health span is disastrous, then conversely, 8. Steinkraus KA, Smith ED, Davis C, et  al. Dietary restriction sup-
presses proteotoxicity and enhances longevity by an hsf-1-dependent
shorter life span with a disproportionately longer health span, like
mechanism in Caenorhabditis elegans. Aging Cell. 2008;7:394–404.
we observed for thermotolerance, would be advantageous. Although
doi:10.1111/j.1474-9726.2008.00385.x
it may be true that there are economic benefits of having a shorter-
9. Bansal A, Zhu LJ, Yen K, Tissenbaum HA. Uncoupling lifespan and health-
lived, healthier population, such a proposition brings to mind images span in Caenorhabditis elegans longevity mutants. Proc Natl Acad Sci
of a dystopian world akin to that portrayed in Logan’s Run, where USA. 2015:112;E277–E286. doi:10.1073/pnas.1412192112
the aged are culled on their 30th birthday before they can burden 10. Hahm JH, Kim S, DiLoreto R, et al. C. elegans maximum velocity corre-
society. Additionally, this view of longevity research is missing one lates with healthspan and is maintained in worms with an insulin receptor
important consideration, that time itself is a precious commod- mutation. Nat Commun. 2015;6:8919. doi:10.1038/ncomms9919
ity which scientists and economists alike have strived to appraise 11. Fischer KE, Hoffman JM, Sloane LB, et al. A cross-sectional study of male
(42–44). Therefore, while chronological versus physiological com- and female C57BL/6Nia mice suggests lifespan and healthspan are not
necessarily correlated. Aging. 2016. doi:10.18632/aging.101059
parisons are useful to evaluate the effect environment plays in the
12. Huffman DM, Justice JN, Stout MB, Kirkland JL, Barzilai N, Austad SN.
deterioration of health, evaluating the utility of pathways that affect
Evaluating health span in preclinical models of aging and disease: guide-
aging for therapeutic potential is best considered using metrics that
lines, challenges, and opportunities for geroscience. J Gerontol A Biol Sci
account for both health and time. Med Sci. 2016;71:1395–1406. doi:10.1093/gerona/glw106
13. Brody JA, Schneider EL. Diseases and disorders of aging: an hypothesis. J
Chronic Dis. 1986;39:871–876.
Supplementary Material 14. Henderson ST, Johnson TE. daf-16 integrates developmental and environ-
Supplementary data are available at The Journals of Gerontology, mental inputs to mediate aging in the nematode Caenorhabditis elegans.
Series A: Biological Sciences and Medical Sciences online. Curr Biol. 2001;11:1975–1980.
15. Senoo-Matsuda N. A defect in the cytochrome B large subunit in complex
II causes both superoxide anion overproduction and abnormal energy
Funding metabolism in Caenorhabditis elegans. J Biol Chem. 2001;276:41553–
41558. doi:10.1074/jbc.M104718200
This work was supported by grants from the National Institute on Aging of 16. Ishii N, Takahashi K, Tomita S, et al. A methyl viologen-sensitive mutant
the National Institutes of Health (R00AG037621) and by the Ellison Medical of the nematode Caenorhabditis elegans. Mutat Res. 1990;237:165–171.
Foundation (AG-NS-1087-13), both to A.R. Additionally, this project was 17. Hsu A-L. Regulation of aging and age-related disease by DAF-16 and heat-
supported by Institutional Development Awards (IDeA) from the National shock factor. Science. 2003;300:1142–1145. doi:10.1126/science.1083701
Institute of General Medical Sciences of the National Institutes of Health 18. Hajdu-Cronin YM. The L-type cyclin CYL-1 and the heat-shock-fac-

(grant numbers P20GM0103423 and P20GM104318, respectively). tor HSF-1 are required for heat-shock-induced protein expression in
Caenorhabditis elegans. Genetics. 2004;168:1937–1949. doi:10.1534/
genetics.104.028423
Acknowledgments 19. Walker GA. Heat shock factor functions at the convergence of the stress
The authors thank George Sutphin and Santina Snow for discussions and response and developmental pathways in Caenorhabditis elegans. FASEB J.
critical comments. Some strains were gratefully provided by the CGC, which 2003;17:1960–1962. doi:10.1096/fj.03-0164fje
480 Journals of Gerontology: BIOLOGICAL SCIENCES, 2017, Vol. 72, No. 4

20. Berdichevsky A, Viswanathan M, Horvitz HR, Guarente L. C. elegans SIR- 32. Vidal-Gadea A, Topper S, Young L, et al. Caenorhabditis elegans selects dis-
2.1 interacts with 14-3-3 proteins to activate DAF-16 and extend life span. tinct crawling and swimming gaits via dopamine and serotonin. Proc Natl
Cell. 2006;125:1165–1177. doi:10.1016/j.cell.2006.04.036 Acad Sci USA. 2011;108:17504–17509. doi:10.1073/pnas.1108673108
21. Nahabedian JF, Qadota H, Stirman JN, Lu H, Benian GM. Bending
33. Knight AW, Billinton N. Distinguishing GFP from cellular autofluores-
amplitude—a new quantitative assay of C. elegans locomotion: identifica- cence. Biophotonics Int. 2001;8:42–51.
tion of phenotypes for mutants in genes encoding muscle focal adhesion 34. Feng F-K, E L-L, Kong X-P, Wang D-S, Liu H-C. Lipofuscin in saliva and
components. Methods San Diego Calif. 2012;56:95–102. doi:10.1016/j. plasma and its association with age in healthy adults. Aging Clin Exp Res.
ymeth.2011.11.005 2015;27:573–580. doi:10.1007/s40520-015-0326-3
22. Luo S, Kleemann GA, Ashraf JM, Shaw WM, Murphy CT. TGF-β and 35. Hosokawa H, Ishii N, Ishida H, Ichimori K, Nakazawa H, Suzuki K.
insulin signaling regulate reproductive aging via oocyte and germline qual- Rapid accumulation of fluorescent material with aging in an oxygen-
ity maintenance. Cell. 2010;143:299–312. doi:10.1016/j.cell.2010.09.013 sensitive mutant mev-1 of Caenorhabditis elegans. Mech Ageing Dev.
23. Glenn CF, Chow DK, David L, et  al. Behavioral deficits during early
1994;74:161–170. doi:10.1016/0047-6374(94)90087-6
stages of aging in Caenorhabditis elegans result from locomotory defi- 36. Balch WE, Morimoto RI, Dillin A, Kelly JW. Adapting proteostasis

cits possibly linked to muscle frailty. J Gerontol A  Biol Sci Med Sci. for disease intervention. Science. 2008;319:916–919. doi:10.1126/
2004;59:1251–1260. science.1141448
24. Herndon LA, Schmeissner PJ, Dudaronek JM, et al. Stochastic and genetic 37. López-Otín C, Blasco MA, Partridge L, Serrano M, Kroemer G.

factors influence tissue-specific decline in ageing C.  elegans. Nature. The hallmarks of aging. Cell. 2013;153:1194–1217. doi:10.1016/j.
2002;419:808–814. doi:10.1038/nature01135 cell.2013.05.039
25. Tsalik EL, Hobert O. Functional mapping of neurons that control locomo- 38. Ben-Zvi A, Miller EA, Morimoto RI. Collapse of proteostasis represents
tory behavior in Caenorhabditis elegans. J Neurobiol. 2003;56:178–197. an early molecular event in Caenorhabditis elegans aging. Proc Natl Acad
doi:10.1002/neu.10245 Sci USA. 2009;106:14914–14919.
26. Wakabayashi T, Kitagawa I, Shingai R. Neurons regulating the dura-
39. Labbadia J, Morimoto RI. Repression of the heat shock response is a pro-
tion of forward locomotion in Caenorhabditis elegans. Neurosci Res. grammed event at the onset of reproduction. Mol Cell. 2015;59:639–650.
2004;50:103–111. doi:10.1016/j.neures.2004.06.005 doi:10.1016/j.molcel.2015.06.027
27. Sutphin GL, Kaeberlein M. Measuring Caenorhabditis elegans life span on 40. Koopman JJ, Wensink MJ, Rozing MP, van Bodegom D, Westendorp RG.
solid media. J Vis Exp JoVE. 2009;27. doi:10.3791/1152 Intrinsic and extrinsic mortality reunited. Exp Gerontol. 2015;67:48–53.
28. Coburn C, Allman E, Mahanti P, et  al. Anthranilate fluorescence marks a doi:10.1016/j.exger.2015.04.013
calcium-propagated necrotic wave that promotes organismal death in C. ele- 41. Redecke L, Binder S, Elmallah MI, et al. UV-light-induced conversion and
gans. PLoS Biol. 2013;11:e1001613. doi:10.1371/journal.pbio.1001613 aggregation of prion proteins. Free Radic Biol Med. 2009;46:1353–1361.
29. Avery L, Horvitz HR. Effects of starvation and neuroactive drugs on
doi:10.1016/j.freeradbiomed.2009.02.013
feeding in Caenorhabditis elegans. J Exp Zool. 1990;253:263–270. 42. Arnesen T, Nord E. The value of DALY life: problems with ethics and
doi:10.1002/jez.1402530305 validity of disability adjusted life years. Lepr Rev. 2000;71:123–127.
30. Zeckhauser R, Shepard D. Where now for saving lives? Law Contemp 43. Landefeld JS, Seskin EP. The economic value of life: linking theory

Probl. 1976:5–45. to practice. Am J Public Health. 1982;72:555–566. doi:10.2105/
31. Zembroń-Łacny A, Dziubek W, Rogowski Ł, Skorupka E, Dąbrowska G. AJPH.72.6.555
Sarcopenia: monitoring, molecular mechanisms, and physical interven- 44. Mrozek JR, Taylor LO. What determines the value of life? A meta-analysis.
tion. Physiol Res. 2014;63:683–691. J Policy Anal Manage. 2002;21:253–270. doi:10.1002/pam.10026

Das könnte Ihnen auch gefallen