ArticlePDF Available

A genomewide oscillation in transcription gates DNA replication and cell cycle

Authors:

Abstract and Figures

Microarray analysis from a yeast continuous synchrony culture system shows a genomewide oscillation in transcription. Maximums in transcript levels occur at three nearly equally spaced intervals in this approximately 40-min cycle of respiration and reduction. Two temporal clusters (4,679 of 5,329) are maximally expressed during the reductive phase of the cycle, whereas a third cluster (650) is maximally expressed during the respiratory phase. Transcription is organized functionally into redox-state superclusters with genes known to be important in respiration or reduction being synthesized in opposite phases of the cycle. The transcriptional cycle gates synchronous bursts in DNA replication in a constant fraction of the population at 40-min intervals. Restriction of DNA synthesis to the reductive phase of the cycle may be an evolutionarily important mechanism for reducing oxidative damage to DNA during replication.
Content may be subject to copyright.
A genomewide oscillation in transcription gates DNA
replication and cell cycle
Robert R. Klevecz*, James Bolen, Gerald Forrest, and Douglas B. Murray
Dynamics Group, Department of Biology, Beckman Research Institute of the City of Hope Medical Center, Duarte, CA 91010
Edited by Steven L. McKnight, University of Texas Southwestern Medical Center, Dallas, TX, and approved November 26, 2003 (received for review
October 7, 2003)
Microarray analysis from a yeast continuous synchrony culture
system shows a genomewide oscillation in transcription. Maxi-
mums in transcript levels occur at three nearly equally spaced
intervals in this 40-min cycle of respiration and reduction. Two
temporal clusters (4,679 of 5,329) are maximally expressed during
the reductive phase of the cycle, whereas a third cluster (650) is
maximally expressed during the respiratory phase. Transcription is
organized functionally into redox-state superclusters with genes
known to be important in respiration or reduction being synthe-
sized in opposite phases of the cycle. The transcriptional cycle gates
synchronous bursts in DNA replication in a constant fraction of the
population at 40-min intervals. Restriction of DNA synthesis to
the reductive phase of the cycle may be an evolutionarily impor-
tant mechanism for reducing oxidative damage to DNA during
replication.
U
ntil recently, dynamic analyses of the regulation of gene
expression were necessarily confined to data from a few
protein products or messages or were approached indirectly by
means of perturbation analysis supported by computer simula-
tions (1–3). Cell structure and the spatial compartmentalization
of metabolic and macromolecular processes are well understood
and have been for some time. In contrast, the temporal orga-
nization of cells, the origins of order and periodicity in pheno-
type, while intriguing several generations of theoretical biolo-
gists (46), have only recently become accessible to definitive
experimentation. The impediments to such experimentation
have, in part, been removed by microchip and expression array
technologies (7–11), which allow one to focus not so much on the
proximal function of a few genes but on the search for a global
mechanism capable of organizing a stable phenotype and timing
cellular events.
The gating of cell cycle events such as DNA replication and
cell division by an oscillator that is an integral submultiple of the
cell cycle was first observed as ‘‘quantized’’ generation times in
mammalian cells (12) and subsequently extended to a wide
variety of organisms (13, 14). Recently, evidence has been
obtained in fission and budding yeast that is consistent with the
idea that generation times are quantized, or that there is a
free-running oscillator underlying the cell cycle (15, 16). One
expectation that follows from this observation is that the entry
into S phase or cell division would occur only at times that are
equal to or multiples of the fundamental period. The question of
whether such timing extends to transcription as a whole as
suggested by recent analyses (10, 11, 17), or whether the oscil-
lation will be limited to genes already identified as important in
cell cycle progression, should be examined.
Much of our success in the genetic and molecular genetic
dissection of the genome has been accomplished by setting aside
concerns about the moment-to-moment, or hour-to-hour,
changes in the transcriptome. This facility with genetic dissection
has led to an emphasis on steady-state relationships in the
connectivity among genes. In consequence, it may be surprising
to find evidence for genomewide oscillations. Computational
studies have suggested that periodic gating might be modeled as
a population of coupled attractors, where each attractor element
is taken to represent the dynamic behavior of individual genes or
clusters of coregulated genes (18–21). If the cell is a coupled
system, oscillations already identified in ‘‘cell cycle-regulated’’
genes might imply genomewide oscillations because behavior of
a transcript can be readily tuned or moved to different patterns
of expression internally by interactions with other transcripts, or
externally by cell-to-cell signaling or perturbations. Indeed, it is
difficult in coupled complex systems to maintain some elements
or genes at a constant level while others are oscillating.
Materials and Methods
Media and Culture Conditions. The WT parent strain used through-
out this study was the diploid Saccharomyces cerevisiae IFO 0233.
Details of culturing methods were as described (22–26). Fer-
menters were from B. Braun Biotech, Surrey, U.K. (Model
Biolab; working volume 750 ml). The basic medium consisted of
5gliter (NH
4
)
2
SO
4
,2gliter KH
2
PO
4
, 0.5 gliter MgSO
4
7H
2
O,
0.1 gliter CaCl
2
2H
2
O, 0.02 gliter FeSO
4
7H
2
O, 0.01 gliter
ZnSO
4
7H
2
O, 0.005 gliter CuSO
4
5H
2
O, 0.001 gliter
MnCl
2
4H
2
O,1mlliter 70% H
2
SO
4
,and1gkg Difco yeast
extract. Glucose medium was supplemented with 22 gliter
glucose monohydrate and 1 mlliter Sigma Antifoam A. Unless
otherwise stated, the fermenters were operated at an agitation
rate of 750 rpm, an aeration rate of 150 mlmin
1
, a temperature
of 30°C, and a pH of 3.4. Dissolved oxygen (DO) levels in the
medium, carbon dioxide production, and oxygen consumption
were measured every 10 s. Cultures were not nutrient limited,
and glucose levels oscillated between 50 and 100
M in each
cycle.
Total RNA Preparation. RNA samples were collected every 4 min
(Fig. 1). Five hundred microliters of yeast sample was pipetted
into 1 ml of ice-cold absolute ethanol. Samples were stored at
72°C overnight and centrifuged at 10,000 g for 2 min, and the
RNA was isolated by using glass beads to rupture the cells
according to the small-scale RNA procedure of Brown (27). The
RNA was dissolved in diethyl pyrocarbonate-treated milli-Q
water and DNase treated with DNA-free (Ambion) in buffer.
The final RNA samples were analyzed in a Uvikon spectropho-
tometer (Kontron Instruments, Zurich), giving 260280 ratios of
2.0–2.2. Samples were also analyzed on an Agilent (Palto Alto,
CA) 2100 Bioanalyzer, giving 25S18S ratios of 1.8–2.0.
Target PreparationProcessing for Affymetrix GeneChip Analysis.
Purified total RNA samples were processed as recommended in
the Affymetrix GeneChip Expression Analysis Technical Man-
ual (Affymetrix, Santa Clara, CA). RNA samples were adjusted
to a final concentration of 1
g
l. Typically, 25–250 ng was
loaded onto an RNA Lab-On-A-Chip (Caliper Technologies.,
This paper was submitted directly (Track II) to the PNAS office.
Abbreviation: DO, dissolved oxygen.
See Commentary on page 1118.
*To whom correspondence should be addressed. E-mail: rklevecz@coh.org.
© 2004 by The National Academy of Sciences of the USA
1200–1205
PNAS
February 3, 2004
vol. 101
no. 5 www.pnas.orgcgidoi10.1073pnas.0306490101
Mountain View, CA) and analyzed in an Agilent 2100 Bioana-
lyzer. Double-stranded cDNA was synthesized from 10
gof
total RNA by using a SuperScript double-stranded cDNA syn-
thesis kit (Invitrogen) and oligo(dT) primers containing a T7
RNA polymerase promoter. Double-stranded cDNA was used as
a template to generate cRNA by using a BioArray High-Yield
RNA Transcript Labeling kit (Enzo Diagnostics). The biotin-
labeled cRNA was fragmented to 35200 bases following the
Affymetrix protocol. Ten micrograms of fragmented cRNA was
hybridized to yeast S98 Affymetrix arrays at 45°Cfor16hinan
Affymetrix GeneChip Hybridization Oven 320. The GeneChip
arrays were washed and then stained with streptavidin-
phycoerythrin on an Affymetrix Fluidics Station 400, followed by
scanning on a HewlettPackard GeneArray scanner. Results
were quantified and analyzed by using
MICROARRAY SUITE 5.0
software (Affymetrix). EXCEL files were created to permit fur-
ther processing in
MATHCAD (Mathsoft, Cambridge, MA), SIG-
MAPLOT,orMATLAB (Mathworks, Natick, MA). Intensity values
for each of the 6,317 ORFs in the chip were linked to the
Saccharomyces Genome Database site, and both their genetic
and physical map locations were associated with the intensity
values for each gene. The results for all ORFs scored as present
by using the default Affymetrix settings were identified accord-
ing to the original sample number and the phase in the DO
oscillation to which they are mapped for presentation. Further
analysis was performed for any ORF present in at least one
sample in each of the three cycles. Of the ORFs scored as present
by these criteria, 41 were removed from the analysis based on the
recent sequence reanalysis (28).
Northern Analyses. Total RNA (57
g) was denatured in glyoxal
and analyzed on a 1.0% agarose gel according to the procedure
of Burnett (29). The RNA was transferred to a Nytran nylon
membrane by using high-efficiency transfer solution (Tel-Test,
Friendswood, TX) in a rapid downward flow turboblotter
(Schleicher & Schuell), washed in 20 mM Tris buffer, pH 8 at
60°C for 10 min to remove glyoxal and baked (1 h, 80°C). The
remaining gel was stained and checked for complete RNA
transfer. Probes were labeled with
32
P dCTP by using the
Prime-It RmT random primer labeling kit (Stratagene). Blots
were prehybridized in QuikHyb solution (Stratagene) at 68°C for
30 min. Hybridization was carried out at 68°C in QuikHyb for 23
h. The blots were washed, exposed to a Molecular Dynamics
phosphor screen, and analyzed on a Typhoon 9410 variable mode
imager (Amersham Pharmacia).
Flow Cytometry. Samples were taken from fermenter cultures
(300
l; 1 10
8
cells) and fixed by adding 700
lof95%
ethanol and stored at 4°C. Before staining for flow cytometric
analyses aliquots of 50
l (containing 1 10
7
cells) were
washed twice in 1 ml of PBS, centrifuged at 5,000 g, and
resuspended in PBS. Samples were then treated with RNase (150
g enzyme; 37°C for 2 h) and stained overnight with propidium
iodide (PI) (5
g for 1 10
7
cells). The samples were analyzed
on a MoFlo MLS (Dako Cytomation, Fort Collins, CO) flow
cytometer. Data were acquired by using dual laser excitation.
Light scatter signals (forward and side) were acquired with
excitation from a HeNe laser (Melles Griot, Carlsbad, CA). PI
was excited with 500 mW at 488-nm wavelength from an Innova
90 Argon laser (Coherent Radiation, Santa Clara, CA). The PI
signal was split through a 580DRLP and a 630DRLP dichroic
beam splitting filters and collected with a 640LP filter (Omega
Optical, Brattleboro, VT).
Results
The Benchmark Oscillation in DO. To begin an experimental analysis
of the connectivity relationships among genes, it was essential to
have a very reproducible and precise biological system. Oscilla-
tions in respiratory activity in budding yeast and other organisms
have been known for many years (30, 31). At sufficient cell
densities in batch or continuous cultures, cell-to-cell signaling
synchronizes the culture with respect to oxidative and reductive
functions (22). These changes are mirrored by changes in the
mitochondrial structure and function (23), flux through the citric
acid cycle (24), and redox state of the cell measured by
NAD(P)H fluorescence and glutathione levels (24). Synchroni-
zation of the respiratory oscillation in the population of cells
appears, at a minimum, to involve respiratory inhibition by the
release of small amounts of H
2
S (25) and phase shifts induced by
acetaldehyde (26). Caused presumably by initial condition de-
pendence, the period of the oscillation can vary from 40 to 44
min, but the coefficient of variation in period or amplitude in a
given culture, once the oscillation is established, is 2% (32). In
Fig. 1, the DO level, indicating the respiratory (low DO) and
reductive phases (high DO levels) is shown for one of the
continuous cultures sampled for these experiments, together
with the points in the cycle when samples were taken.
Expression Microarray Analysis. Using Affymetrix chips, the pat-
terns of gene expression for 5,329 expressed genes were followed
by sampling at 4-min intervals through three cycles of the DO
oscillation. The reproducibility of the global pattern of expres-
sion can be seen in the color contour map of expression levels
through the three cycles where orange-red indicates high levels
of expression and blue indicates low levels. Genes are arranged
in Fig. 2 according to their time of maximum expression in the
cycle. The time of maximum gene expression in the cycle was
determined by averaging gene expression in the three replicates
(Fig. 2A), assigning the phase of maximum expression to each
transcript then examining how well this assignment is repro-
duced through the three cycles (Fig. 7, which is published as
supporting information on the PNAS web site). More than 87%
of all expressed genes are expressed maximally in the reductive
Fig. 1. Sampling for expression microarray analysis from continuous syn-
chrony cultures. Respiratory-reductive synchrony is monitored by using DO
levels in the media of aerobic continuous cultures. The sample times are shown
imposed on the DO oscillation. High DO levels are associated with the reduc-
tive phase and low DO levels with the respiratory phase of the cycle. Samples
were taken at 4-min intervals through three complete cycles each from two
independent cultures. A total of 68 RNA samples were isolated and analyzed.
All RNA samples were analyzed for quality by using an Agilent Bioanalyzer
capillary electrophoresis system. Thirty-two samples were selected from the
best RNA proles and aligned based solely on their phase position on the DO
curve, before Affymetrix expression analysis. Sampling for ow cytometry is
shown as red squares and for RNA as yellow triangles.
Klevecz et al. PNAS
February 3, 2004
vol. 101
no. 5
1201
CELL BIOLOGY SEE COMMENTARY
phase of the cycle. Of these, 2,429 genes are maximally expressed
early in the reductive phase, whereas 2,250 genes are maximally
expressed late in the reductive phase of the cycle. The remaining
650 genes are maximally expressed in the respiratory phase of the
cycle (Fig. 2B). In general, the genes maximally expressed in the
respiratory phase show higher-amplitude oscillations than do
those maximally expressed in the reductive phase. The average
peak-to-trough ratio in the respiratory phase is 2.40, whereas in
the reductive phase it is 1.89. Only 41 transcripts show a 6-fold
change, whereas just 7 show a 20-fold change. It appears that
low-amplitude oscillations in expression are a near universal
property of this system.
Northern Blot Analysis. Probes for six genes were prepared from
yeast cDNA by PCR using primers that amplify the entire coding
regions of the expressed genes (Fig. 3). These genes were chosen
as representative based on their patterns of expression and time
of maximum expression as indicated: early respiratory phase
(STR3), mid respiratory phase (MET17), late respiratoryearly
reductive phase (CRC1), early reductive phase (CYB5), and late
reductive phase (YRO2 and YDR070c). A total of 18 RNA time
series samples taken from the second and third cycles of Fig. 7
were subjected to Northern blot analysis for each of the six genes.
Comparison between the Affymetrix intensity results and the
Northern blots was accomplished by dividing the radioactivity
after background subtraction by the average radioactivity in all
18 samples. This scaling to the average expression of all mea-
Fig. 2. Average expression levels from three cycles of the respiratory oscil-
lation. Color contour (intensity) maps of the expression levels of the 5,329
expressed genes are shown for all 32 RNA samples through three cycles of the
DO oscillation. (A) The average expression level for the three biological
replicates are shown. High levels of expression are orange, and low levels are
blue. Genes were scored as present based on the Affymetrix default settings
and included in the analysis if at least 1 of the 32 was scored as present in each
of the three cycles. Genes are dened according to the most recent assign-
ments (23). Values shown here were scaled by dividing the average expression
level for each gene into each of the time-series samples for that gene.
Transcripts were ordered according to their phase of maximum expression in
the average of the three replicates. This same scaling and ordering was used
in Fig. 7. Samples are identied according to their phase in the cycle (0360°
cycle). Sample phases are shown in reference to the DO curve (thick black line,
B). The reductive phase is taken to be the period of minimal oxygen consump-
tion (maximum DO, yellow background) in the interval between the minimum
DO levels, and the respiratory phase is shown against a blue-green back-
ground. (B) Summary of the results for the time of maximum expression
(red line) for the transcripts of A. Color scale: orange-red ⫽⬎1.6 and dark blue
⫽⬍0.8.
Fig. 3. Comparison of Affymetrix microchip data with Northern blots.
Northern blot analyses of transcripts found to be maximally expressed at
differing points in the redoxtranscriptional cycle were compared with the
results from the microarray analyses. Six clones of representative genes were
examined in 18 time-series samples representing 1.5 cycles of the DO oscilla-
tion and were scaled by dividing each time point by the average of all of the
points. Red lines indicate Northern blot data, and black lines indicate
Affymetrix intensity data.
1202
www.pnas.orgcgidoi10.1073pnas.0306490101 Klevecz et al.
surements for each transcript is the same scaling as was per-
formed on the Affymetrix fluorescence intensity measures
throughout this study. In Fig. 3, the results of the two assays are
compared. The patterns of expression found in the Affymetrix
microchips are reproduced to a considerable degree by the
results from the Northern blot analysis.
Temporal Organization of Functionally Related Transcripts. Prior
emphasis in microarray analysis has been on paired treated
versus control samples examining those genes showing differ-
ences 2-fold, although it has been reported that changes in the
1.3- to 1.4-fold range can be significant (33). Of the 5,329
expressed genes only 161 show oscillations in the range of 1.3- to
1.4-fold peak-to-trough differences through the cycle. These fall
into two classes with 118 showing intensities 2,000 per slide.
Most of these are known highly expressed genes such as actin,
ribosomal proteins (52 of 118), genes of intermediary metabo-
lism such as PGK, and other structural or maintenance proteins.
This low-amplitude oscillation in maintenance genes is consis-
tent with the report of Warrington et al. (34), who found
evidence that maintenance genes in mammalian systems show
oscillations. If the time course of actin (ACT1) is followed, there
is an intensity oscillation of 1,350 fluorescent units superimposed
on a high actin level of 5,862 intensity units within the cycle.
Thus, although the fold change in expression is low, the pattern
of oscillations in ACT1 and the others in the cluster with high
average expression levels is clear. The pattern shows three peaks
of expression, as though these constitutive expressers are tran-
scribed at every permitted phase, whereas expression of other
functionally related groups is limited to one phase. This high
transcript level, low-amplitude behavior, is shown for the cyto-
solic ribosomal proteins in Fig. 4 Upper Right.
To look at the temporal organization of functionally related
transcripts (Fig. 4), we compared transcripts for genes previously
identified as important in respiration with those identified with
respiratory inhibition (reduction). For example, transcripts for
mitochondrial ribosomal proteins are maximally expressed dur-
ing the reductive phase when mitochondrial function is minimal
(23) and the cristae are more clearly defined (orthodox config-
uration). Conversely, transcripts such as those involved in me-
thionine and sulfur metabolism, important in the establishment
of the reductive phase and DNA replication are made early in the
respiratory phase, before their putative maximal function at the
beginning of DNA replication (Fig. 4). As with all transcripts
examined, there is a 4- to 12-min lag between the peak in
transcript level and the expected time of maximum function of
the protein product. This phase anticipation is functionally
significant if protein synthesis is similarly temporally coherent
and occurs with a fixed-phase relationship to transcription. In
support of this idea, the ubiquitin-constituent transcripts of the
proteosome are clustered and maximally expressed in the re-
ductive phase (Fig. 4 and Table 1, which is published as
supporting information on the PNAS web site).
The
-DNA-polymeraseprimase complex is made up of four
subunits, each of which is an independent transcript with non-
contiguous physical map locations. Although it could be other-
wise, one might expect that the biological function would make
simultaneous transcription likely and thus offers an opportunity
to apply a biological rationale to the examination of the patterns
in the transcriptional cycle. The average of the three replicates
of the four subunits are shown together with the SD of intensity
change (Fig. 5). The timing of expression shows the same 4- to
12-min phase anticipation as mentioned above, consistent with
the timing of DNA synthesis from the flow cytometric analysis
of the cell cycle presented below. Similarly, transcripts function-
ally related to sulfur metabolism also show very reproducible
patterns (Fig. 5). As an alternative to the color map of Fig. 2, a
subset of the early reductive-phase transcripts representative of
the range of expression in this group have been coplotted with
the DO oscillation (Fig. 8, which is published as supporting
information on the PNAS web site). Table 1 contains the
expression patterns for the transcripts shown in Figs. 4 and 5.
Flow Cytometry. The relationship between DNA replication and
the transcriptional cycle was studied by flow cytometric analysis.
Fig. 4. Temporal organization of functionally related transcripts. Subsets of
functionally related transcripts were color-mapped as in Fig. 1 to show the
time of expression maximums as described above. The les used to generate
these color maps are available Table 1. For sulfur metabolism transcripts
(Lower Right) the scale is orange-red ⫽⬎2.2 and dark blue ⫽⬍0.8. For all
other panels, orange-red ⫽⬎1.4 and dark blue ⫽⬍0.8.
Fig. 5. Cycle-to-cycle reproducibility in selected functional groupings. Sub-
tracting the minimum expression and scaling the result to an average of 1
determines the change in average intensity for each gene. (Left) Six sulfur-
associated genes (see Table 1) were averaged, and the SD was determined.
(Right) The genes for the four subunits of the
DNA polymeraseprimase
complex were averaged, and the SD was determined. The error bars represent
1 SD.
Klevecz et al. PNAS
February 3, 2004
vol. 101
no. 5
1203
CELL BIOLOGY SEE COMMENTARY
By sampling at 4-min intervals through four cycles of the DO
oscillation (Fig. 6C) and then aligning multiple flow cytometric
DNA histogram analyses with the respiratory oscillation, we
found that DNA replication in these cells begins abruptly at the
end of the respiratory phase, as oxygen consumption decreases
and H
2
S levels rise. To make this synchronous gating of cells into
S phase clear, all 39 of the 1D propidium iodide-DNA histo-
grams (Fig. 6A) were stacked, and the values for S were
determined by removal of G
1
and G
2
by Gaussian deconvolution.
Although 10% of the cells are gated into S phase in any turn
of the respiratorytranscriptional cycle, this synchronous gating
is so precisely timed that a track representing the movement of
cells from early to late S can be followed in the 3D reconstruc-
tions. In Fig. 6B, the increase in DNA content is shown by the
green track running from left to right and diagonally upward
from early to late S phase, DNA replication continues through-
out the reductive phase, ending just as respiration begins. The
duration of S phase and DNA synthesis in these cells is 24 min,
close to that observed in the most rapidly growing high glucose
cultures. The movement of the synchronous cohort through S
can also be followed in animations viewed at www.talandic.com.
The behavior of DNA replication, the most readily identifiable
cell cycle event in this culture system, is dynamically similar to
the circadian time-scale gating seen in animal model systems (35)
and humans (36) where a correlation between glutathione levels
and circadian gating of cells into S phase has been observed (36).
The role of reactive oxygen intermediates and the cellular
mechanisms preventing damage to single-stranded DNA during
replication have been extensively studied (37, 38). However, the
possibility that temporal organization in the cell has evolved to
prevent or minimize such damage has received little attention. In
the case of RNA synthesis where repair pathways for oxidative
damage are unknown, it is of interest to note that 87% of the
transcripts examined here are maximally expressed during the
reductive phase.
Discussion
Although the respiratory oscillation has been investigated in a
number of laboratories, neither the relationship between tran-
scription and respiratory metabolism, nor its connection to the
cell cycle and to the gating of cells into S phase and mitosis has
previously been explored. Under the growth conditions com-
monly used in this system, the population doubling time is 8h,
and it has been assumed that the 40- to 44-min respiratory
oscillation was uncoupled from the cell cycle. It appears rather
that this short period cycle might be the fundamental timekeep-
ing oscillator because transcription and the gating of DNA
replication and the duration of the cell cycle show similar
quantized distributions. In the strain of yeast used in these
studies a series of deletion mutations have been developed that
show DO oscillations with either half or twice the period (20 and
80 min) of the WT (39). It will be of interest to determine
whether these period additions or period doublings of the DO
oscillation and the respiratory cycle (39) are manifested as well
in the period of transcription. A reanalysis of population dou-
bling time data in yeast cultures growing on different media with
differing carbon sources from rather poor to rich found that
population doubling times cluster at multiples of the 40-min
fundamental (32, 40), further urging upon us the view that the
cell cycle is a downstream manifestation of a timekeeping
attractor (17). The spatio-temporal dynamics and cell-to-cell
signaling in these vigorously stirred cultures that gives rise to
precise periodic outputs is the subject of ongoing investigations.
Alter et al. (10, 41) and others (11, 19) in their reanalysis of
the original Stanford cell cycle data using singular value decom-
position or wavelet decomposition have all come to a similar
conclusion: that there is evidence for a genomewide, low-
amplitude oscillation in transcription. Here, using a yeast con-
tinuous culture system exhibiting a very high level of precision
and stability, genomewide oscillations in transcription have been
demonstrated by using Affymetrix chips and replicate sampling
through three cycles of the respiratory oscillation. The long-
standing assumptions that the oscillator is uncoupled from the
cell cycle and that the oscillation is confined to a small subset of
genes intimately connected to energy metabolism is refuted by
the finding of genomewide oscillations in transcription. We find
sufficient evidence to support the claim that housekeeping,
maintenance, or constitutive genes are not expressed at a
constant level through the transcriptional cycle. As a practical
matter, the canonical-treated versus control paradigm is made
suspect when the assumption of steady-state conditions is shown
to be incorrect. Given that the patterns of expression shown here
hold through multiple cycles even when transcripts showing very
low-amplitude expression patterns are considered, the question
is raised whether these oscillatory patterns, unacknowledged in
most expression array experiments, represent an uncontrolled
variable and contribute to the biological variability and hence the
apparent 2-fold limitations to sensitivity.
Fig. 6. Flow cytometric analysis of S-phase gating in the reductive phase of
the cycle. The 1D frequency histogram (A) showing DNA content of the
population for 39 samples taken at 4-min intervals have been stacked (B)in
relationship to the time in the DO oscillation at which they were sampled (C)
and color-mapped to show numbers of cells in early, middle, and late S phase
(B). In the resulting 2D color map, with red indicating more cells and blue
representing fewer cells, the track of cells through S phase appears as a series
of green bands moving left to right and upward on the diagonal. Red 350
cells, light green 200 cells, and dark blue 0 cells.
1204
www.pnas.orgcgidoi10.1073pnas.0306490101 Klevecz et al.
Continuous culture techniques allow the examination of
genomewide expression in a manner that avoids the pitfalls of
conventional synchrony that is, synchronization-associated per-
turbations to the system, lack of cycle-to-cycle reproducibility,
and rapid decay of blockage-induced synchrony, while permit-
ting sampling to be as frequent and enduring as required. In
other words, we can sample for expression array analysis from a
biologically stable platform as frequently as we can afford to
sample, and we can do so under conditions where the benchmark
oscillation in respiration guides the sampling regimen. Using this
system it should be possible to better define and eliminate
sources of variability in expression arrays.
The most interesting realization to come out of these studies
is that the separation in time between oxidative and reductive
phases propagates throughout the transcriptome and is coordi-
nated with the initiation of DNA replication. This may be an
important strategy evolved in cells to prevent oxidative damage
to DNA during replication. Because transcription is mirrored in
the oscillations in oxidative and reductive genes, there is the
implication that protein synthesis and catabolism must be sim-
ilarly organized in time. Given the totality of this coordinate
expression of such central and yet disparate elements as DNA
replication, the machinery for protein synthesis and degradation,
and mitochondrial ribosomes, it seems unavoidable to expect
that similar temporal organization will be found in mammalian
and other systems.
We thank the Analytical Cytometry Core facility for help in flow
cytometric analysis and J. Denis Heck and the University of California-
Irvine DNA MicroArray Facility for help with the expression array
studies. This work is supported by a Beckman Research Institute grant.
1. Klevecz, R. R. & Ruddle, F. H. (1968) Science 159, 634636.
2. Kauffman, S. A. & Wille, J. J. (1975) J. Theor. Biol. 55, 4793.
3. Mackey, M. C. & Glass, L. (1977) Science 197, 287289.
4. Waddington, C. D. (1957) The Strategy of the Gene (Allen and Unwin, London)
5. Eigen, M. & Schuster, P. (1977) Naturwissenschaften 64, 541565.
6. Kauffman, S. A. (1993) The Origins of Order (Oxford Univ. Press, Cambridge,
U.K.)
7. Wodicka, L., Dong, H., Mittman, M., Ho, M. H. & Lockhart, D. (1997) Nat.
Biotechnol. 15, 13591367.
8. Cho, R. J., Campbell, M. J., Winzeler, E. A., Steinmetz, L., Conway, A.,
Wodicka, L., Wolfsberg, T. G, Gabrielian, A. E., Landsman, D., Lockhart, D. J.
& Davis, R. W. (1998) Mol. Cell. 2, 6573.
9. Spellman, P. T., Sherlock, G., Zhang, M. Q., Iyer, V. R., Anders, K., Eisen,
M. B., Brown, P. O., Botstein, D. & Futcher, B. (1998) Mol. Biol. Cell. 9,
32733297.
10. Alter, O., Brown, P. O. & Botstein, D. (2000) Proc. Natl. Acad. Sci. USA 97,
1010110106.
11. Klevecz, R. R. (2000) Funct. Integr. Genomics 1, 186192.
12. Klevecz, R. R. (1976) Proc. Natl. Acad. Sci. USA 73, 40124016.
13. Chisholm, S. W. & Costello, J. H. (1980) J. Phycol. 16, 375383.
14. Homma, K. & Hastings, J. W. (1988) J. Biol. Rhythms 3, 4958.
15. Haase, S. B. & Reed S. I. (1999) Nature 401, 394397.
16. Sveiczer, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. J. & Novak, B. (2000)
Proc. Natl. Acad. Sci. USA 97, 78657870.
17. Rifkin, S. A. & Kim, J. (2002) Bioinformatics 18, 11761183.
18. Klevecz, R. R., Kauffman, S. A. & Shymko, R. M. (1984) Int. Rev. Cytol. 86,
97128.
19. Klevecz, R. R. (1998) Physica D 124, 110.
20. Klevecz, R. R., Bolen, J. L. & Dura´n, O. (1992) Int. J. Bifurcation Chaos 2,
941953.
21. Bolen, J. L., Dura´n, O. & Klevecz, R. R. (1993) Physica D 67, 245256.
22. Satroutdinov, A. D., Kuriyama, H. & Kobayashi, H. (1992) FEMS Microbiol.
Lett. 77, 261267.
23. Lloyd, D., Salgado, L. E., Turner, M. P., Suller, M. T. & Murray, D. B. (2002)
Microbiology 148, 37153724.
24. Murray, D. B., Engelen, F. A. A., Lloyd, D. & Kuriyama, H. (1999) Microbiology
145, 27392745.
25. Sohn, H. Y., Murray, D. B. & Kuriyama, H. (2000) Yeast 16, 11851190.
26. Murray, D. B., Klevecz, R. R. & Lloyd, D. (2003) Exp. Cell Res. 287, 1015.
27. Brown, A. J. P. (1996) in Yeast Protocols, ed. Evans, I. H. (Humana, Totowa,
NJ), pp. 269276.
28. Kellis, M., Patterson, N., Endrizzi, M., Birren, B. & Lander, E. S. (2003) Nature
423, 241254.
29. Burnett, W. V. (1997) BioTechniques 22, 668671.
30. Mochan, E. & Pye, E. K. (1973) Nat. New Biol. 242, 177179.
31. Poole, R. K., Lloyd, D. & Kemp, R. B. (1973) J. Gen. Microbiol. 77, 209.
32. Klevecz, R. R. & Murray, D. B. (2001) Mol. Biol. Rep. 28, 7382.
33. Yue, H., Eastman, P. S., Wang, B. B., Minor, J., Doctolero, M. H., Nuttall,
R. L., Stack, R., Becker, J. W., Montgomery, J. R., Vainer, M. & Johnston, R.
(2001) Nucleic Acids Res. 29, E41.
34. Warrington, J. A., Nair, A., Mahadevappa, M. & Tsyganskaya, M. (2000)
Physiol. Genomics 2, 143147.
35. Potten, C. S., Booth, D., Cragg, N. J., OShea, J. A., Tudor, G. L. & Booth, C.
(2002) Cell Prolif. 35, 1621.
36. Smaaland, R., Abrahamsen, J. F., Svardal, A. M., Lote, K. & Ueland, P. M.
(1992) Br. J. Cancer 66, 3945.
37. Georgiou, G. (2002) Cell 111, 607610.
38. Koerkamp, M. G., Rep, M., Bussemaker, H. J., Hardy, G. P., Mul, A.,
Piekarska, K., Szigyarto, C. A., De Mattos, J. M. & Tabak, H. F. (2002) Mol.
Biol. Cell. 13, 27832794.
39. Adams, C. A., Kuriyama, H., Lloyd, D. & Murray D. B. (2003) Yeast 20,
463470.
40. Wheals, A. E., Tyson, C. B. & Lord, P. G. (1979) J. Bacteriol. 138, 9298.
41. Alter, O. Brown, P. O. & Botstein, D. (2003) Proc. Natl. Acad. Sci. USA 100,
33513356.
Klevecz et al. PNAS
February 3, 2004
vol. 101
no. 5
1205
CELL BIOLOGY SEE COMMENTARY
... While traditionally described as a result of carbon limitation, limitations by other essential nutrients like phosphate [5] and ammonium, ethanol, phosphate, glucose, and sulfur [6,7] can lead to metabolic oscillations. Under the growth conditions commonly used in this system, the population doubling time and thus the length of the cell division cycle is about 8 h, and the metabolic oscillations have period 40 − 44 min [8]. ...
... The oscillations were first observed as periodic oscillations in the oxygen consumption of continuous, glucose-limited cultures growing in a chemostat, but were later also observed in batch cultures [9]. The MC has two distinct phases : low oxygen consumption (LOC) phase when dissolved oxygen in the medium is high and high oxygen consumption phase (HOC) when the oxygen in the medium drops to low levels [2,3,8]. Using experimental techniques ranging from micorarray analysis [2,8] to short-life luciferase fluorescent reporters [4] researchers were able assign transcription arXiv:2305.07643v1 ...
... The MC has two distinct phases : low oxygen consumption (LOC) phase when dissolved oxygen in the medium is high and high oxygen consumption phase (HOC) when the oxygen in the medium drops to low levels [2,3,8]. Using experimental techniques ranging from micorarray analysis [2,8] to short-life luciferase fluorescent reporters [4] researchers were able assign transcription arXiv:2305.07643v1 [math.DS] 12 May 2023 of particular genes to these phases. ...
Preprint
We introduce two time-delay models of metabolic oscillations in yeast cells. These oscillations arise as a result of resource deprivation. Our model tests a hypothesis that the oscillations occur as multiple pathways share a limited resource which we equate to the number of available ribosomes. We initially explore a single-protein model with a constraint equation governing the total resource available to the cell. The model is then extended to include three coupled proteins that share the same resource pool. Three approaches are considered to numerically detect existence of a limit cycle. First, numerical stability diagrams are generated in the parameter space by plotting response features resulting from nonlinear simulations. We use features such as the standard deviation and maximum 1D persistent homology to analyze the solutions. The second method uses a spectral element approach to solve a boundary value problem to estimate the period of oscillation. Finally, we use the spectral element method to approximate the system as a high dimensional discrete map and evaluate the stability of the linearized system by examining its eigenvalues. Our results show that for certain combinations of total resource available to the cell and the time delay lead to limit cycles whereby the equilibrium solution loses stability via Hopf bifurcation. Further, for certain protein production times (delays) in the three protein model, we observe a phase shift in the protein production rates suggesting that our model correctly captures the shared resource pool dynamics.
... Our conclusion that the temporal segregation of biosynthetic processes dictates the primary metabolic dynamics has a number of direct consequences. First, the earlier conjectures on the causes of metabolic dynamics during the cell cycle, such as respiratory activity 34,49,50 and carbohydrate-storage turnover 51-53 , should not be correct. Second, as the temporal segregation of biosynthetic processes is likely a condition-independent behavior, metabolic dynamics should occur across all nutrients on which cells grow and divide. ...
... By measuring NAD(P)H levels in single cells as a readout of biosynthetic and primary metabolic dynamics, we tested whether the above-mentioned consequences of our finding are correct. First, the conjectures that metabolic oscillations are caused by dynamics in respiration 34,49,50 or carbohydrate-storage metabolism 51-53 are expected to be incorrect. Indeed, decreasing the oxygen content in the microfluidic device, confirmed by a drop in the level of mCherry-tagged γ-subunit Atp3 of the ATP synthase 54 , did not affect the NAD(P)H oscillations ( Fig. 5a and Extended Data Fig. 9). ...
... An early work based on glucose-limited synchronous cultures 59 and a recent multi-omics study with α-factor-synchronized cells 9 has generated important indications along these lines, but we can now (based on direct activity measurements) provide actual evidence to this notion. Together with the fact that we have observed metabolic oscillations under a broad range of experimental conditions whenever cells divided, this indicates that the metabolic oscillations do not emerge in specific primary metabolic pathways, such as respiration-or storage-related pathways, as earlier conjectured 34,[49][50][51][52][53] . Thus, primary metabolism is dynamic likely because it has to fulfill the temporarily changing demands for precursors, redox and energy cofactors to supply the different biosynthetic processes. ...
Article
Full-text available
Many cell biological and biochemical mechanisms controlling the fundamental process of eukaryotic cell division have been identified; however, the temporal dynamics of biosynthetic processes during the cell division cycle are still elusive. Here, we show that key biosynthetic processes are temporally segregated along the cell cycle. Using budding yeast as a model and single-cell methods to dynamically measure metabolic activity, we observe two peaks in protein synthesis, in the G1 and S/G2/M phase, whereas lipid and polysaccharide synthesis peaks only once, during the S/G2/M phase. Integrating the inferred biosynthetic rates into a thermodynamic-stoichiometric metabolic model, we find that this temporal segregation in biosynthetic processes causes flux changes in primary metabolism, with an acceleration of glucose-uptake flux in G1 and phase-shifted oscillations of oxygen and carbon dioxide exchanges. Through experimental validation of the model predictions, we demonstrate that primary metabolism oscillates with cell-cycle periodicity to satisfy the changing demands of biosynthetic processes exhibiting unexpected dynamics during the cell cycle.
... Despite major insights, such as the dependence of cell division on cell growth [7,8,16] and the identification of participating pathways [17,18], many aspects of the coordination between cell growth and division remain unclear [3,19,20]. To investigate this coordination, we exploited the spontaneous metabolic synchronization of yeast cultures: when a glucose-starved culture of budding yeast is refed with a glucose-limited medium at a constant rate, the culture can begin a respiratory cycle, manifested by periodic oscillation in the levels of dissolved oxygen in the culture medium [21,22] and periodic expression of thousands of genes [23][24][25][26][27], reviewed in [28][29][30][31]. Recently we observed similar oscillations in phosphate-starved nondividing yeast cells [26], demonstrating that these phenomena of metabolic synchronization are not restricted to carbon-source-limited or continuous yeast cultures. ...
... This metabolic cycling in single cells, however, does not exactly parallel the metabolic cycling in synchronized cultures since measurements of the DNA content of single cells from the synchronized cultures [8,23,24,[35][36][37] have demonstrated at least two sub-populations during each period of the synchronized cultures: dividing (2C DNA content) and non-dividing (1C DNA content). Thus, distinguishing between metabolic cycling in synchronized cultures and in single cells is crucial to the analysis throughout this article. ...
... This dependence is commonly observed in models of quorum-sensing based synchronization [44,45] and likely relevant to the metabolic synchronization as well. It may provide the unifying principle behind the "short" respiratory cycle of 45min [23] observed in IFO cultures fed with a medium containing high glucose concentration (22g/L) and the "longer" cycles of 4 − 5h [24] observed in CEN.PK cultures fed with a medium containing lower glucose concentration (10g/L). The difference in frequency is also likely to be strain dependent, i.e., different sensitivity for, and/or a different rate of secretion of, the quorum sensing chemical mediating synchrony. ...
Preprint
Full-text available
Yeast cells grown in culture can spontaneously synchronize their respiration, metabolism, gene expression and cell division. Such metabolic oscillations in synchronized cultures reflect single-cell oscillations, but the relationship between the oscillations in single cells and synchronized cultures is poorly understood. To understand this relationship and the coordination between metabolism and cell division, we collected and analyzed DNA-content, gene-expression and physiological data, at hundreds of time-points, from cultures metabolically-synchronized at different growth rates, carbon sources and biomass densities. The data enabled us to extend and generalize our mechanistic model, based on ensemble average over phases (EAP), connecting the population-average gene-expression of asynchronous cultures to the gene-expression dynamics in the single-cells comprising the cultures. The extended model explains the carbon-source specific growth-rate responses of hundreds of genes. Our physiological data demonstrate that the frequency of metabolic cycling in synchronized cultures increases with the biomass density, suggesting that this cycling is an emergent behavior, resulting from the entraining of the single-cell metabolic cycle by a quorum-sensing mechanism, and thus underscoring the difference between metabolic cycling in single cells and in synchronized cultures. Measurements of constant levels of residual glucose across metabolically synchronized cultures indicate that storage carbohydrates are required to fuel not only the G1/S transition of the division cycle but also the metabolic cycle. Despite the large variation in profiled conditions and in the scale of their dynamics, most genes preserve invariant dynamics of coordination with each other and with the rate of oxygen consumption. Similarly, the G1/S transition always occurs at the beginning, middle or end of the high oxygen consumption phases, analogous to observations in human and drosophila cells. These results highlight evolutionary conserved coordination among metabolism, cell growth and division.
... ATP generation from fermentation is much faster than from oxidative phosphorylation when glucose is available. In yeast, growth is also characterized by DNA replication, mainly without mitochondrial oxidative activity (aerobic glycolysis), most probably to protect DNA from mutations (Klevecz et al., 2004). In M. oryzae MoIsw2 could thus be an ATP regulated switch (Machné and Murray, 2012) between fast aerobic glycolysis, DNA-synthesis growth and quality control,and slower but more high ATP yield e cient oxidative phosphorylation growth, oxidative defenses, and interaction with the abiotic and biotic environment that makes up the M. oryzae ecological niche. ...
... oxygen is consumed and the substrates oxidized (catabolism)(Klevecz et al., 2004;Machné and Murray, 2012). In addition, the binding sequences close to avirulence genes and retrotransposons, so their regulation can shift depending on retrotransposon transpositions. ...
... Most of the downregulated genes in the DMoisw2 are genes involved in mitochondrial electron transport and other mitochondrial processes. In other words, our results indicate that MoIsw2 is involved in regulating the balance between DNA synthesis, which is safest without ATP generation by mitochondrial respiration that generates Radical Oxygen Species (ROS) that is a necessary biproduct of respiratory metabolism(Klevecz et al., 2004).Regulation of DNA-binding genes closest to the MoIsw2 binding site ...
Preprint
Full-text available
Isw2 proteins are conserved in eukaryotes and are known to bind to DNA and dynamically influence local chromosome condensation close to the DNA binding site in an ATP-dependent manner making genes close to the binding sites more accessible for transcription and repression. A putative MoISW2 gene was deleted with large effects on pathogenicity as a result, complemented and combined with a ChIP-sec to identify binding sites, RNAsec to test effects on regulation of genes along the chromosomes and compared with RNAseq from 55 downloaded RNA-seq datasets from the same strain. MoIsw2 binding and activities create regions of multiple binding sites with high gene expression variability close to the binding sites while surrounding regions have lower gene expression variability. We show that genes affected by the MoIsw2 activity are niche-determinant genes (secreted proteins, secondary metabolites and stress-coping genes) and avirulence genes. We further show that MoIsw2 binding sites coincide with known transposable elements (TE) making it likely that TE-transposition of the binding sites can affect the transcription profile of M. oryze in a strain-specific manner. We discuss that since avirulence genes are close to the MoIsw2 sites, TE transposition likely changes not just the expression and eventual silencing of the avirulence genes but the expression of a whole range of niche determinant genes. We conclude that MoIsw2 is a likely candidate for a master regulator that regulates the dynamic balance between biomass growth and nutrient uptake combined with reactions to the biotic and abiotic environment, i.e., the balance between anabolism and catabolism. Since TE transposition is known to be stress-induced the overall effect of TE activity together with MoIsw2 activity is likely to be a mechanism that creates more mutations and faster evolution of niche determinant genes than for housekeeping genes.
... In bacteria, the net result is a precise and reproducible timing of DNA synthesis in the cell cycle across a wide range of nutritional conditions and growth rates [3][4][5][6]. In eukaryotes, the metabolic control is thought to confine DNA synthesis to the reduction phase of a redox metabolic cycle, reiterated several times per cell cycle [7][8][9][10][11][12][13]. The mechanism of the metabolic control of replication remains largely unknown as well as its interface with "classical" replication control functions and importance for cell survival. ...
Article
Full-text available
Background In all living organisms, DNA replication is exquisitely regulated in a wide range of growth conditions to achieve timely and accurate genome duplication prior to cell division. Failures in this regulation cause DNA damage with potentially disastrous consequences for cell viability and human health, including cancer. To cope with these threats, cells tightly control replication initiation using well-known mechanisms. They also couple DNA synthesis to nutrient richness and growth rate through a poorly understood process thought to involve central carbon metabolism. One such process may involve the cross-species conserved pyruvate kinase (PykA) which catalyzes the last reaction of glycolysis. Here we have investigated the role of PykA in regulating DNA replication in the model system Bacillus subtilis. Results On analysing mutants of the catalytic (Cat) and C-terminal (PEPut) domains of B. subtilis PykA we found replication phenotypes in conditions where PykA is dispensable for growth. These phenotypes are independent from the effect of mutations on PykA catalytic activity and are not associated with significant changes in the metabolome. PEPut operates as a nutrient-dependent inhibitor of initiation while Cat acts as a stimulator of replication fork speed. Disruption of either PEPut or Cat replication function dramatically impacted the cell cycle and replication timing even in cells fully proficient in known replication control functions. In vitro, PykA modulates activities of enzymes essential for replication initiation and elongation via functional interactions. Additional experiments showed that PEPut regulates PykA activity and that Cat and PEPut determinants important for PykA catalytic activity regulation are also important for PykA-driven replication functions. Conclusions We infer from our findings that PykA typifies a new family of cross-species replication control regulators that drive the metabolic control of replication through a mechanism involving regulatory determinants of PykA catalytic activity. As disruption of PykA replication functions causes dramatic replication defects, we suggest that dysfunctions in this new family of universal replication regulators may pave the path to genetic instability and carcinogenesis.
Article
Full-text available
The recently observed circadian oscillations of the intestinal microbiota underscore the profound nature of the human-microbiome relationship and its importance for health. Together with the discovery of circadian clocks in non-photosynthetic gut bacteria and circadian rhythms in anucleated cells, these findings have indicated the possibility that virtually all microorganisms may possess functional biological clocks. However, they have also raised many essential questions concerning the fundamentals of biological timekeeping, its evolution, and its origin. This narrative review provides a comprehensive overview of the recent literature in molecular chronobiology, aiming to bring together the latest evidence on the structure and mechanisms driving microbial biological clocks while pointing to potential applications of this knowledge in medicine. Moreover, it discusses the latest hypotheses regarding the evolution of timing mechanisms and describes the functions of peroxiredoxins in cells and their contribution to the cellular clockwork. The diversity of biological clocks among various human-associated microorganisms and the role of transcriptional and post-translational timekeeping mechanisms are also addressed. Finally, recent evidence on metabolic oscillators and host-microbiome communication is presented.
Article
We introduce two time-delay models of metabolic oscillations in yeast cells. Our model tests a hypothesis that the oscillations occur as multiple pathways share a limited resource which we equate to the number of available ribosomes. We initially explore a single-protein model with a constraint equation governing the total resource available to the cell. The model is then extended to include three proteins that share a resource pool. Three approaches are considered at constant delay to numerically detect oscillations. First, we use a spectral element method to approximate the system as a discrete map and evaluate the stability of the linearized system about its equilibria by examining its eigenvalues. For the second method, we plot amplitudes of the simulation trajectories in 2D projections of the parameter space. We use a history function that is consistent with published experimental results to obtain metabolic oscillations. Finally, the spectral element method is used to convert the system to a boundary value problem whose solutions correspond to approximate periodic solutions of the system. Our results show that certain combinations of total resource available and the time delay, lead to oscillations. We observe that an oscillation region in the parameter space is between regions admitting steady states that correspond to zero and constant production. Similar behavior is found with the three-protein model where all proteins require the same production time. However, a shift in the protein production rates peaks occurs for low available resource suggesting that our model captures the shared resource pool dynamics.
Article
Metabolism and DNA replication are the two most fundamental biological functions in life. The catabolic branch of metabolism breaks down nutrients to produce energy and precursors used by the anabolic branch of metabolism to synthesize macromolecules. DNA replication consumes energy and precursors for faithfully copying genomes, propagating the genetic material from generation to generation. We have exquisite understanding of the mechanisms that underpin and regulate these two biological functions. However, the molecular mechanism coordinating replication to metabolism and its biological function remains mostly unknown. Understanding how and why living organisms respond to fluctuating nutritional stimuli through cell-cycle dynamic changes and reproducibly and distinctly temporalize DNA synthesis in a wide-range of growth conditions is important, with wider implications across all domains of life. After summarizing the seminal studies that founded the concept of the metabolic control of replication, we review data linking metabolism to replication from bacteria to humans. Molecular insights underpinning these links are then presented to propose that the metabolic control of replication uses signalling systems gearing metabolome homeostasis to orchestrate replication temporalization. The remarkable replication phenotypes found in mutants of this control highlight its importance in replication regulation and potentially genetic stability and tumorigenesis.
Article
As centers for energy production and essential biosynthetic activities, mitochondria are vital for cell growth and proliferation. Accumulating evidence suggests an integrated regulation of these organelles and the nuclear cell cycle in distinct organisms. In budding yeast, a well-established example of this coregulation is the coordinated movement and positional control of mitochondria during the different phases of the cell cycle. The molecular determinants involved in the inheritance of the fittest mitochondria by the bud also seem to be cell cycle-regulated. In turn, loss of mtDNA or defects in mitochondrial structure or inheritance often lead to a cell cycle delay or arrest, indicating that mitochondrial function can also regulate cell cycle progression, possibly through the activation of cell cycle checkpoints. The up-regulation of mitochondrial respiration at G2/M, presumably to fulfil energetic requirements for progression at this phase, also supports a mitochondria-cell cycle interplay. Cell cycle-linked mitochondrial regulation is accomplished at the transcription level and through post-translational modifications, predominantly protein phosphorylation. Here, we address mitochondria-cell cycle interactions in the yeast Saccharomyces cerevisiae and discuss future challenges in the field.
Article
H 2 S belongs to the class of molecules known as gasotransmitters, which also includes nitric oxide (NO) and carbon monoxide (CO). Three enzymes are recognized as endogenous sources of H 2 S in various cells and tissues: cystathionine g-lyase (CSE), cystathionine β-synthase (CBS) and 3-mercaptopyruvate sulfurtransferase (3-MST). The current article reviews the regulation of these enzymes as well as the pathways of their enzymatic and non-enzymatic degradation and elimination. The multiple interactions of H 2 S with other labile endogenous molecules (e.g. NO) and reactive oxygen species are also outlined. The various biological targets and signaling pathways are discussed, with special reference to H 2 S and oxidative posttranscriptional modification of proteins, the effect of H 2 S on channels and intracellular second messenger pathways, the regulation of gene transcription and translation and the regulation of cellular bioenergetics and metabolism. The pharmacological and molecular tools currently available to study H 2 S physiology are also reviewed, including their utility and limitations. In subsequent sections, the role of H 2 S in the regulation of various physiological and cellular functions is reviewed. The physiological role of H 2 S in various cell types and organ systems are overviewed. Finally, the role of H 2 S in the regulation of various organ functions is discussed as well as the characteristic bell-shaped biphasic effects of H 2 S. In addition, key pathophysiological aspects, debated areas, and future research and translational areas are identified A wide array of significant roles of H 2 S in the physiological regulation of all organ functions emerges from this review.
Article
Full-text available
SUMMARY Large-scale synchronous cultures of the fission yeast Schizosaccharot?iyces/ionzbe 972h- were prepared by a sedimentation-velocity selection method in a zonal rotor. Oxygen uptake was measured polarographically on samples withdrawn at frequent intervals from the culture vessel. Rates of oxygen uptake, expressed per ml culture, increased exponentially, doubling over each cell cycle, but rose to maxima twice per cycle, once during cell division and once at approximately one-half of the cycle. Oxygen uptake at the maxima was stimulated by carbonyl-cyanide m-chlorophenyl hydrazone; that at the troughs was insensitive to this compound. Oxygen uptake was inhibited at all points in the cycle by antimycin A and cyanide at low concentrations; inhibition by these compounds did not alter the periodicity of the oscillations. Heat evolution increased at a constant rate during two cycles of synchronous growth, but in the presence of carbonyl-cyanide m-chlorophenyl hydrazone, peaks of heat evolution in phase with peaks of oxygen uptake were observed. The discontinuous respiratory activity of mitochondria through the cell cycle is discussed with reference to previous contrasting data, and possible control mechanisms are suggested .
Book
Stuart Kauffman here presents a brilliant new paradigm for evolutionary biology, one that extends the basic concepts of Darwinian evolution to accommodate recent findings and perspectives from the fields of biology, physics, chemistry and mathematics. The book drives to the heart of the exciting debate on the origins of life and maintenance of order in complex biological systems. It focuses on the concept of self-organization: the spontaneous emergence of order widely observed throughout nature. Kauffman here argues that self-organization plays an important role in the emergence of life itself and may play as fundamental a role in shaping life's subsequent evolution as does the Darwinian process of natural selection. Yet until now no systematic effort has been made to incorporate the concept of self-organization into evolutionary theory. The construction requirements which permit complex systems to adapt remain poorly understood, as is the extent to which selection itself can yield systems able to adapt more successfully. This book explores these themes. It shows how complex systems, contrary to expectations, can spontaneously exhibit stunning degrees of order, and how this order, in turn, is essential for understanding the emergence and development of life on Earth. Topics include the new biotechnology of applied molecular evolution, with its important implications for developing new drugs and vaccines; the balance between order and chaos observed in many naturally occurring systems; new insights concerning the predictive power of statistical mechanics in biology; and other major issues. Indeed, the approaches investigated here may prove to be the new center around which biological science itself will evolve. The work is written for all those interested in the cutting edge of research in the life sciences.
Article
In yeast and somatic cells, mechanisms ensure cell-cycle events are initiated only when preceding events have been completed¹. In contrast, interruption of specific cell-cycle processes in early embryonic cells of many organisms does not affect the timing of subsequent events², indicating that cell-cycle events are triggered by a free-running cell-cycle oscillator. Here we present evidence for an independent cell-cycle oscillator in the budding yeast Saccharomyces cerevisiae. We observed periodic activation of events normally restricted to the G1 phase of the cell cycle, in cells lacking mitotic cyclin-dependent kinase activities that are essential for cell-cycle progression. As in embryonic cells, G1 events cycled on schedule, in the absence of S phase or mitosis, with a period similar to the cell-cycle time of wild-type cells. Oscillations of similar periodicity were observed in cells responding to mating pheromone in the absence of G1 cyclin (Cln)- and mitotic cyclin (Clb)-associated kinase activity, indicating that the oscillator may function independently of cyclin-dependent kinase dynamics. We also show that Clb-associated kinase activity is essential for ensuring dependencies by preventing the initiation of new G1 events when cell-cycle progression is delayed.
Article
The cDNA microarray is one technological approach that has the potential to accurately measure changes in global mRNA expression levels. We report an assessment of an optimized cDNA microarray platform to generate accurate, precise and reliable data consistent with the objective of using microarrays as an acquisition platform to populate gene expression databases. The study design consisted of two independent evaluations with 70 arrays from two different manufactured lots and used three human tissue sources as samples: placenta, brain and heart. Overall signal response was linear over three orders of magnitude and the sensitivity for any element was estimated to be 2 pg mRNA. The calculated coefficient of variation for differential expression for all non-differentiated elements was 12–14% across the entire signal range and did not vary with array batch or tissue source. The minimum detectable fold change for differential expression was 1.4. Accuracy, in terms of bias (observed minus expected differential expression ratio), was less than 1 part in 10 000 for all non-differentiated elements. The results presented in this report demonstrate the reproducible performance of the cDNA microarray technology platform and the methods provide a useful framework for evaluating other technologies that monitor changes in global mRNA expression.
Article
Planktonic foraminiferal tests are not formed in isotopic equilibrium with seawater; the deviation is species dependent.
Article
Large discrete arrays of chaotic attractors, coupled by diffusion, organize into asynchronous periodic spiral waves, synchronous periodic bands, turbulent fields or synchronous chaos as a function of coupling strength and array size. Self-organization of periodic spirals in both two and three dimensional arrays of nonexcitable systems appears to require the early establishment of an antipodal phase relationship between the few cells that will form the vortex. Cells within or close to the vortex maintain low z amplitude, near limit cycle trajectories, with stable, well-defined phase relationships. In periodic banding structures, initial antipodal phase seeds evolve to isochrons that form nested periodic trajectories. The likelihood that biological systems are fundamentally oscillatory and chaotic is discussed.
Article
The dynamic origins of phenotypic heterogeneity and genotypic instability and hypermutation have been investigated in simulated tissues comprised of 900-25 600 cells each represented by initially identical Rossler attractors running in the chaotic domain. This attractor, representing the cell cycle behavior of individual cells in the array, has previously been used to model the dynamic behavior of mammalian cells in culture. In these tissue constructs, the behavior of an individual cell is modified by its interactions with its immediate neighbors as a consequence of diffusive coupling through one of the variables. Differentiation within the initially identical population of attractors is manifested as a position dependent set of novel stable trajectories in phase space that are revealed through the use of return maps. These self-mapping patterns, which we define as the phenotype of the cell, are periodic and stable over a considerable period of time. A comparison of tissues whose individual cell cycle attractor phases describe an archimedean spiral with those that exhibit S-T chaos, or turbulence, suggests that the heterogeneous phenotype of tumor tissues is better modeled by turbulence. Instability in the spiral array exists primarily at the boundary between periodic regions of differing phase and trajectory, and involves infrequent excursions by these boundary cells away from their stables trajectories. Such instabilities are hypothesized to play an important role in the amplification, hypermutation, and gene conversion events seen in certain normal biological tissues and tumors.