Sei sulla pagina 1di 13

Environmental Microbiology (2017) 19(11), 4493–4505 doi:10.1111/1462-2920.

13898

Testing the metabolic theory of ecology with marine


bacteria: different temperature sensitivity of major
phylogenetic groups during the spring phytoplankton
bloom

Nestor Arandia-Gorostidi,1* than 0.65 eV, the value predicted by the Metabolic
Tamara Megan Huete-Stauffer ,1,2 Theory of Ecology (MTE). Contrary to MTE predic-
Laura Alonso-Sa ez1,3 and Xose  Anxelu G. Mora n2 tions, carrying capacity tended to increase with
1 warming for all bacterial groups. Our analysis con-
Plankton Ecology and Pelagic Ecosystem Dynamics
Division, Instituto Espan~ ol de Oceanografı́a, Centro firms that resource availability is key when
Oceanogra fico de Gijo
 n/Xixo n, Gijo
 n/Xixo
 n, Asturias, addressing the temperature response of heterotro-
Spain. phic bacterioplankton. We further show that even
2
Biological and Environmental Sciences and under nutrient-sufficient conditions, warming differ-
Engineering Division, King Abdullah University of entially affected distinct bacterioplankton taxa.
Science and Technology (KAUST), Red Sea Research
Center, Thuwal, Saudi Arabia.
3
Marine Research Division, AZTI, Sukarrieta, Bizkaia, Introduction
Spain.
One of the current challenges in marine ecology is to pre-
dict how global change will affect the structure and
Summary metabolism of marine biota. Ocean warming may increase
Although temperature is a key driver of bacterio- surface temperature between 1 and 68C depending on the
plankton metabolism, the effect of ocean warming on CO2 emissions scenario (Meehl et al., 2007; Collins et al.,
different bacterial phylogenetic groups remains 2013), posing a threat to future marine food webs
unclear. Here, we conducted monthly short-term (Edwards and Richardson, 2004; Wiltshire and Manly,
incubations with natural coastal bacterial communi- 2004; Sarmento et al., 2010) and biogeochemical pro-
ties over an annual cycle to test the effect of cesses (Sarmiento et al., 1998; Laws et al., 2000). Since
experimental temperature on the growth rates and heterotrophic prokaryotes represent the largest living
carrying capacities of four phylogenetic groups: biomass in the oceans (Whitman et al., 1998) and play a
SAR11, Rhodobacteraceae, Gammaproteobacteria key role in all biogeochemical cycles, the effects of ocean
and Bacteroidetes. SAR11 was the most abundant warming on these planktonic organisms may have large
consequences for ecosystem functioning (Kirchman et al.,
group year-round as analysed by CARD-FISH, with
2005; Degerman et al., 2013; Huete-Stauffer et al., 2015).
maximum abundances in summer, while the other
However, there is still no consensus about the metabolic
taxa peaked in spring. All groups, including SAR11,
response of heterotrophic bacteria to increasing tempera-
showed high temperature-sensitivity of growth rates
tures (e.g., Li et al., 2004; Sarmento et al., 2010; Mora n
and/or carrying capacities in spring, under phyto-
et al., 2011), and to what extent it is comparable to that of
plankton bloom or post-bloom conditions. In that
other organisms.
season, Rhodobacteraceae showed the strongest
In the last decade, the Metabolic Theory of Ecology
temperature response in growth rates, estimated
(MTE, Brown et al., 2004) has emerged as a valuable
here as activation energy (E, 1.43 eV), suggesting an
framework to test the metabolic response to changes in
advantage to outcompete other groups under warmer
temperature of different organisms, including heterotrophic
conditions. In summer E values were in general lower
bacteria (Daufresne et al., 2009). The MTE assumes that
the metabolic rates of all organisms are a combination of
Received 3 February, 2017; revised 16 August, 2017; accepted 17
August, 2017. *For correspondence. E-mail n.arandia86@gmail. the allometric scaling of their body mass (West, 1997) and
com; Tel. 134 985 309 780; Fax 134 985 326 277. biochemical kinetics, which are temperature dependent as
C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd
V
ez and X. A. G. Mora
4494 N. Arandia-Gorostidi, T. M. Huete-Stauffer, L. Alonso-Sa n

described by the Van’t Hoff-Arrhenius relation (Arrhenius, groups of coastal heterotrophic bacteria, hereinafter
1889; Gillooly et al., 2001). Therefore, the MTE merges referred to as bacteria, while assessing their experimental
the effect of temperature and mass into a single equation responses to temperature. In order to test the predictions
that can be used from organisms to communities, allowing of the MTE with specific taxa, we incubated surface natural
the effect of temperature to be tested experimentally. samples from the southern Bay of Biscay continental shelf
Although the MTE has successfully predicted the increase at three different temperatures encompassing 68C variabil-
in metabolic rates with temperature in different organisms ity around the ambient value. We targeted in the analysis
pez-Urrutia et al., 2006; Bailly et al., 2014), the pre-
(Lo typical oligotrophs such as SAR11, as well as more
dicted activation energy for heterotrophic metabolism (0.65 resource-dependent (i.e., copiotrophic) bacteria such as
eV, Gillooly et al., 2001; Brown et al., 2004) has not been Bacteroidetes or Rhodobacteraceae, in order to assess
agreed upon in bacteria, for which variations between 0 whether a systematic differential response to temperature,
and >1 eV have been reported (Sinsabaugh and Shah, according to their trophic strategy, could be identified. Our
2010; Mora  n et al., 2011; Huete-Stauffer et al., 2015). Yet,
results indicate that the taxonomic composition of marine
studies attempting to measure activation energies of envi- bacteria communities may be highly relevant for predicting
ronmental microorganisms are still scarce. the microbial response to warming in a future ocean.
The MTE aims also at predicting the response to warm-
ing of other biological properties, such as the carrying Results
capacity (i.e., the maximum density or biomass that the
environment can sustain with the available resources). The Environmental setting
carrying capacity has been hypothesized to decrease with Surface seawater samples for temperature controlled incu-
temperature (Brown et al., 2004; Savage et al., 2004), bations were collected monthly between January and
since at higher metabolic rates, and thus higher resource
December 2012 at a continental shelf station located in the
requirements, the same energy supply will sustain a lower
southern Bay of Biscay, where effects of coastal pollution
organism abundance. However, the response of the carry-
and river discharges are negligible. After collection, seawa-
ing capacity to temperature is not free of controversy
ter was filtered by 0.8 mm pore-size filters to remove
(Jiang and Morin, 2004), as external forcing, especially
predators. In situ temperature ranged from 12.88C in
resource availability and predation pressure, may affect

the outcome of this relationship (Eiler et al., 2003; Solić
et al., 2009).
In general, substrate availability has been described as
a key factor governing the temperature dependence of
heterotrophic bacterial metabolism (Lo pez-Urrutia and
Mora  n, 2007). A recent study has shown that the overall
dependence of heterotrophic bacterial communities on
temperature has a seasonal component, with stronger
temperature effects in periods with high nutrients and
phytoplankton biomass (Huete-Stauffer et al., 2015). Yet,
marine bacterial communities are highly diverse, and
some of the widespread, abundant groups are markedly
different ranging from oligotrophs to copiotrophs (Koch,
2001; Lauro et al., 2009; Yooseph et al., 2010). Although
disparity in the in situ growth rates of different phylogenetic
groups of marine bacteria, such as SAR11, Rhodobactera-
ceae, Gammaproteobacteria and Bacteroidetes, has been
reported previously in a few studies (see review in Kirch-
man, 2016), the temperature-sensitivity of specific taxa is
unclear and it has been addressed mostly with bacterial
isolates (Ratkowsky et al., 1982; Cho and Giovannoni,
2004). Thus, our knowledge of how temperature affects
the growth and carrying capacity of dominant phylogenetic
groups of environmental bacterial communities is still very
limited. Fig. 1. In situ temperature, chlorophyll a concentration and cell-
specific bacterial heterotrophic production (sBP) (A), and DAPI-
In this study, we address the seasonal variability in the based abundances of total bacteria (line) and of each of the
growth rates and carrying capacities of broad phylogenetic phylogenetic groups identified with CARD-FISH (B).

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
Metabolic theory of ecology with marine bacteria 4495
March to 21.28C in August (Fig. 1A). In situ inorganic biovolumes year-round, while cell-size of Rhodobactera-
nutrients concentration (Supporting Information Fig. S1), ceae was significantly higher (0.049 6 0.018 mm3, ANOVA,
peaked during winter (January and March for phosphate p-value < 0.001). Non-significant differences were found
and dissolved inorganic nitrogen, DIN) and also in late between Bacteroidetes and the other analysed groups,
autumn (December for silicate). Cell-specific bacterial and its average size was very similar to that found for Rho-
heterotrophic production (sBP) was also measured at the dobacteraceae (0.048 6 0.016 mm3). The average cell-size
onset of the experiments as an estimation of carbon of the bacterial community (for all DAPI-stained cells) was
bioavailability (Fig. 1A). sBP showed maximum values in negatively correlated with in situ temperature (Pearson
spring (with a pronounced peak in April) and minimum val- r 5 20.60, p-value 5 0.037, n 5 12), with maximum values
ues in summer and early winter. Total chlorophyll a in spring (May, 0.044 mm3), and minimum values in sum-
concentration was measured in unfiltered seawater sam- mer (August, 0.027 mm3, Supporting Information Fig. S2).
ples, as an estimation of the state of the phytoplankton
community in environmental conditions. Maximum
Seasonal dynamics of in situ bacterial growth rates and
chlorophyll a concentrations were found in late autumn
carrying capacities
(December, 1.8 mg L21) and spring (March and April, 1.4
mg L21), while minimum values were found in summer Every month, seawater was incubated at in situ tempera-
(July, 0.1 mg L21), coinciding with the increase in water col- ture conditions during ca. 1 week in order to measure the
umn stratification. We also took samples for estimating the growth rates and carrying capacities of each phylogenetic
abundance of heterotrophic nanoflagellates (HNF) by flow group. To calculate bacterial growth rates, we took 2 or 3
cytometry (Christaki et al., 2011) before and after the pre- samples every day during the exponential growth phase of
filtration treatment in order to assess the efficiency of 0.8 the bacterial community for CARD-FISH analysis (see
mm pore-size filters in removing potential bacterial preda- experimental procedures). SAR11 showed the lowest
tors. We found that filtration removed on average growth rates at in situ temperature (Fig. 2E–H, ANOVA, p-
98.4% 6 0.8% of HNFs but only 8.4% 6 15.5% of bacteria. value 5 0.002, n 5 12), with a mean value of 0.50 6 0.38
d21. Their growth rates showed low variability year-round,
but unusually high values (>1 d21) were found in spring
Seasonal dynamics of the in situ abundance and
(April and May). Average growth rates of Rhodobactera-
cell-size of bacterial groups
ceae, Gammaproteobacteria and Bacteroidetes were
Abundance of heterotrophic bacteria at the initial time of higher, generally close to 1 d21, but showed different
the incubations showed a high variability, ranging from seasonalities. While maximum growth rates in Gammapro-
2.60 6 0.04 105 cells mL21 in May to 1.02 6 0.01 106 teobacteria and Bacteroidetes were found in spring (2.10
cells mL21 in September (Fig. 1B). The contribution of d21 in May for both groups), Rhodobacteraceae had maxi-
SAR11, Rhodobacteraceae, Gammaproteobacteria and mum growth rates in winter (1.50 d21 in January).
Bacteroidetes to the bacterial community (measured as Occasionally, growth rates were very low at the end of the
the percentage of hybridized cells by CARD-FISH) was on stratified season (September and October) for these three
average 54%, with a minimum in January (25%) and a groups (0.25 d21 to non-detectable growth).
maximum in March (68%). SAR11 was the most abundant The carrying capacity (K) was measured as the maxi-
group for most of the months, reaching up to 51% of total mum bacteria abundance reached by the different
cells in August, but it dropped to 8% in May. Bacteroidetes phylogenetic groups during the incubations (see Supporting
was the second most abundant group ranging from 4% in Information Fig S3). K showed contrasting seasonalities
August to 29% in May. Rhodobacteraceae and Gammap- between SAR11 and the rest of the groups at in situ tem-
roteobacteria showed minimum contributions in November perature. While SAR11 K values were minimum in Spring
(1.5% and 3.2% respectively), and maxima in April (16%). (May) and maximum in summer (July, Fig. 2A), the rest of
Seasonality of SAR11 abundance was clearly different the groups showed minimum K values in late summer-
from the other groups, as maximum abundances were autumn (September for Gammaproteobacteria and October
found in summer rather than in spring. Accordingly, tem- for Rhodobacteraceae and Bacteroidetes), and maximum
perature was positively correlated with SAR11 abundance K values in Spring (March for Gammaproteobacteria and
(Supporting Information Table S1, Pearson r 5 0.63, April for Rhodobacteraceae and Bacteroidetes).
p-value 5 0.017, n 5 12) and negatively with Bacteroidetes
(Pearson r 5 20.60, p-value 5 0.023, n 5 12).
Temperature dependence of growth rates and carrying
Significant differences in the initial cell-size were found
capacities of target bacterial groups
between different phylogenetic groups (Fig. 2I–L). SAR11
(0.038 6 0.007 mm3 excluding size of May) and Gammap- During the incubation experiments, we also determined
roteobacteria (0.038 6 0.010 mm3) showed the lowest the growth rates and carrying capacities of the different
C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
ez and X. A. G. Mora
4496 N. Arandia-Gorostidi, T. M. Huete-Stauffer, L. Alonso-Sa n

Fig. 2. Initial and maximum cell abundance (A–D), growth rates at in situ temperature (E–H) and initial biovolume (I–L) of SAR11,
Rhodobacteraceae (Rhodo), Gammaproteobacteria (Gamma) and Bacteroidetes (Bctd). Dashed line in Biovolume panels indicate year-round
average cell-size.

groups under three experimental temperatures (in situ Rhodobacteraceae (1.46 6 0.06 eV) and December for
temperature and 38C below and above in situ tempera- SAR11 and Bacteroidetes (0.86 6 0.12 eV and 1.03 6 0.11
ture). To test the effect of temperature on these eV), but non-significant differences were observed
parameters, we followed the equations by the MTE (see between the mean annual E values of the different phylo-
Experimental Procedures). The effect of temperature on genetical groups (ANOVA, p-value 5 0.06, n 5 12).
growth rates (Fig. 3E–H) was summarized by the activa- Minimum E values (not statistically different from 0, indica-
tion energy (E) of specific growth rates. All phylogenetic tive of no effects of temperature on growth rate) were
groups consistently showed their highest E values in usually found in summer. In spring, significant differences
months with high chlorophyll a concentrations: April for were found among the E of all groups (ANOVA,
Gammaproteobacteria (0.89 6 0.04 eV), May for p-value < 0.01, n 5 12). The lowest E values were found
C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
Metabolic theory of ecology with marine bacteria 4497

Fig. 3. Dependence of different variables on temperature for the four phylogenetic groups; (A–D) carrying capacity (K-RT) and (E–H) growth
rates represented as activation energy (E). Error bars represent SD between two replicates. Black lines represent the absence of temperature
dependence (coinciding with 0 value of E and K-RT). The dashed line represents the E value predicted by the MTE (0.65 eV).

for SAR11 (0.55 6 0.12 eV) and the highest for Rhodobac- for SAR11 and Rhodobacteraceae and in April and Octo-
teraceae (1.43 6 0.05 eV), with intermediate values for ber for Gammaproteobacteria. K-RT was significantly
Gammaproteobacteria and Bacteroidetes (0.81 6 0.13 eV correlated with E for Rhodobacteraceae (Pearson r 5 0.87,
and 0.89 6 0.04 eV respectively). Activation energies of p-value 5 0.026) and cell-size for SAR11 (Pearson
Gammaproteobacteria and Bacteroidetes were positively r 5 0.74, p-value 5 0.006, n 5 12).
correlated with their initial mean cell-sizes (Supporting
Information Table S1, Pearson’s r 5 0.60, p-value 5 0.023, Discussion
n 5 12 and r 5 0.55, p-value 5 0.038, n 5 12 respectively).
The ratio between the growth rates in the 138C and The response of heterotrophic bacterioplankton to increas-
238C treatments was also calculated for each group in ing temperatures may largely influence biogeochemical
order to better visualize the magnitude of change within
the temperature range assessed (Fig. 4). In accordance
with the activation energies, this ratio showed conspicuous
seasonal variations for all groups. Maximum values were
found in spring, yet with significant differences between
phylogenetic groups (ANOVA, p-value < 0.001, n 5 4).
Rhodobacteraceae showed the highest ratios (3.4) and
SAR11 showed the lowest (1.7). The latter group, together
with Gammaproteobacteria, yielded mean ratios below 1 in
summer (i.e., decrease rather than increase in growth rate
with a 68C temperature rise) (Fig. 4).
Regarding the carrying capacities, either non-significant
effect or an increase in K with temperature (i.e., positive K-
RT values) was found for all bacterial groups, but some
negative values (close to 0) were sporadically found (Fig.
3A–D). Similar to other variables, K-RT values tended to
Fig. 4. Boxplot of the ratio between the specific growth rates in the
show seasonal variability. Maximum K-RT was found in 138C and 238C temperature treatments of each phylogenetic
winter for Bacteroidetes (January), spring (April and May) group in the different seasons.

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
ez and X. A. G. Mora
4498 N. Arandia-Gorostidi, T. M. Huete-Stauffer, L. Alonso-Sa n

processes in the future ocean (Cho and Azam, 1990; agree with the general assumption that some members of
Azam and Long, 2001; Regaudie-de-Gioux and Duarte, the latter three taxonomic groups are predominately copio-
2012). However, the metabolic response of widespread trophs (Kirchman, 2002; Lo pez-Pe rez et al., 2012; Nelson
phylogenetic groups of marine bacteria to ocean warming and Carlson, 2012; Buchan et al., 2014), and support the
is largely unknown. In a companion paper assessing the view that trophic strategies strongly determine the sea-
temperature-dependence of heterotrophic bacteria growth sonal composition of bacterial communities in coastal
rates from a community level perspective, a high temporal environments (Pinhassi and Hagstro € m, 2000; Bunse and
variability was found (Huete-Stauffer et al., 2015). A plausi- Pinhassi, 2017).
ble explanation for this finding was that distinct bacterial The measured growth rates of specific taxa agree with
groups dominating different periods of the year exhibited the relatively few previous studies carried out in natural
different temperature-responses, an idea that, to our conditions (Yokokawa et al., 2004; Yokokawa and Nagata,
knowledge, has been never tested in a complete annual 2005; Teira et al., 2009; Lankiewicz et al., 2016). Thus,
cycle. This hypothesis was addressed in the present work, SAR11 consistently showed the minimum growth rates of
in which we targeted four broad bacterioplankton groups the groups targeted (Fig. 2). However, it is remarkable that
widely distributed in marine waters in order to identify this group occasionally also had high growth rates (April
which taxa are potentially more sensitive to ocean warm- and May), reaching values comparable to those of the
ing. As our aim was to quantify the growth rates of different other groups (>1 d21). Similar high growth rates of SAR11
taxa under a range of temperature treatments, we used a have been previously found only in predator-free as well as
single-cell approach (CARD-FISH), which allows a direct in virus-free incubations (Ferrera et al., 2011), supporting
quantification of cell abundances of target phylogenetic the idea that at least some SAR11 lineages may be tightly
groups (Amann and Fuchs, 2008) along the growth curves. controlled by top-down factors (Zhao et al., 2013). The ele-
It should be taken into account that the taxonomical resolu-
vated growth rates of SAR11 cells measured in May were
tion provided by CARD-FISH is restricted by the specificity
also accompanied by unusually large cell-sizes, in the
and coverage of the phylogenetic probes used. In our
higher end of the few previous estimates of SAR11 bacte-
case, we used specific probes for four abundant, major
ria sizes (0.05–0.12 mm3, Malmstrom et al., 2004; Elifantz
phylogenetic groups (SAR11, Rhodobacteraceae,
et al., 2005; Schattenhofer et al., 2009). These results sug-
Gammaproteobacteria and Bacteroidetes), yet, the growth
gest that a distinct SAR11 lineage, more responsive to
rates measured were likely the result of a combined
high resource availability, was dominant in the post-bloom
response of different taxa within each of them. On average,
conditions of May. At the same site, according to Alonso-
these four groups represented more than half of the initial
Saez et al. (2015), two dominant SAR11 operational taxo-
bacterial abundance in our study site, but increased up to
nomic units (OTU) are typically found, with one of them
90% of the cells (71% 6 16% on average) at the time of
contributing on average 69% to total reads (OTU 1) while
maximum abundance in the incubations. Therefore, we
the second one was generally much less abundant (21%).
feel confident that we targeted the bacterial groups that
were actively growing in the incubations and, most proba- Interestingly, although the OTU 1, affiliated with the broadly
bly, also in situ (Campbell and Kirchman, 2013; Kirchman, distributed surface 1 clade, was clearly dominant over
2016). most of 2012, a clear switch was found in May, when OTU
The bacterial groups analysed showed contrasting sea- 4, affiliated with the SAR11 surface 2 clade, contributed up
sonal dynamics, likely related to their trophic strategies to 83% of SAR11 reads (Supporting Information Fig. S4).
and specific adaptations to occupy different ecological While we cannot confirm which particular SAR11 phylo-
niches (Gifford et al., 2013). For instance, SAR11 domi- type grew in our incubations, these results suggest the
nated in summer, which is characterized by a stratified presence of different SAR11 lineages with distinct eco-
water column and oligotrophic conditions at the site of the physiological properties, with the potential to attain high
study (Calvo-Dı́az and Mora  n, 2006; Mora  n et al., 2007). growth rates and large cell-size under resource-replete
The seasonal dynamics of SAR11 is consistent with the conditions, comparable to those of copiotrophic taxa. Inter-
well-known adaptation of this taxon to low nutrient concen- estingly, a higher activity of some SAR11 phylotypes
trations (Giovannoni et al., 2005; 2017; Sowell et al., during periods of high primary production and BP has
2009). The other analysed groups (Rhodobacteraceae, been detected by metatranscriptomics (Gifford et al.,
Gammaproteobacteria and Bacteroidetes) showed oppo- 2014). Such elevated activity was related with a higher
site seasonal trends, with low abundance in summer expression of genes associated to the use and transport of
stratified conditions and peaking in spring, coinciding with algal-derived low molecular weight compounds, suggesting
recurrent phytoplankton blooms in the area of study that the activity of at least some SAR11 OTUs may
(Calvo-Dı́az, 2008) due to high inorganic nutrient concen- respond to the large inputs of labile organic matter
tration (Supporting Information Fig. S1). These results released after phytoplankton blooms.
C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
Metabolic theory of ecology with marine bacteria 4499
Since temperature governs enzyme kinetics (Elias et al., temperature-sensitivity usually decreases after a gradual
2014), ocean warming is expected to have a generalized, thermal adaptation (Hall et al., 2010). However, the lack of
direct impact on variables related to metabolism (Gillooly statistical differences between summer and winter in the
et al., 2001; Kingsolver and Huey, 2008; O’Connor et al., temperature-sensitivity of both growth rates and carrying
2009; Huete-Stauffer et al., 2015; Mora  n et al., 2015). capacities, suggests that the seasonal variation of in situ
However, we report here large discrepancies with the MTE temperature was not as relevant as resource availability in
predictions of four widespread phylogenetic groups of determining the bacterial response to temperature.
coastal bacterioplankton. While an overall increase in Among the groups targeted, Rhodobacteraceae showed
growth rates with increasing temperature (i.e., positive E the strongest response to temperature in spring, conferring
values) was found in agreement with the MTE expectations this group with a potential competitive advantage under
(Gillooly et al., 2001), a large variation in activation ener- warming conditions. On the contrary, SAR11 showed the
gies was observed (Figs 2 and 3E–H) and E values were lowest temperature-sensitivity of all groups, possibly
often lower than that predicted by the MTE for heterotro- related to the limited ability of members of this clade to
phic organisms (0.65 eV, Brown et al., 2004). On the other respond to environmental stimuli due to their streamlined
hand, the carrying capacity showed an opposite trend to genomes (Giovannoni et al., 2005; Grote et al., 2012). The
MTE predictions (Brown et al., 2004; Savage et al., 2004), other groups, presumably dominated by copiotrophic taxa
as the predicted decrease of K with temperature was only with larger genomes (Yooseph et al., 2010), may have a
observed occasionally for some of the groups, and even faster response to changes of environmental parameters
then, values were close to zero. The pattern of increasing (Lauro et al., 2009) like temperature, as we found experi-
growth rates and maximum standing stocks with increasing mentally. While the reasons for the higher temperature-
temperature seems general for the broad taxonomic bacte- sensitivity of Rhodobacteraceae as compared with
rial groups found in coastal temperate waters, which may Gammaproteobacteria or Bacteroidetes are elusive, we
ultimately lead to increasing bacterial biomass in a warmer hypothesize that the higher ability of Rhodobacteraceae to
ocean, as previously reported (Mora  n et al., 2015). adapt to changing environmental conditions (Moran et al.,
However, it should be taken into account that protistan 2004; Polz et al., 2006; Beier et al., 2015), may lead the
grazing on bacteria, which was avoided in our experimen- observed higher response to temperature. Another plausi-
tal setup through pre-filtration of the samples, will also ble explanation may be related to the composition of the
probably increase in warmer conditions (Vaque  et al., substrates preferentially used by each taxa, which may be
2009; Sarmento et al., 2010), ultimately limiting the characterized by different energetic demands, therefore
observed increase of bacterial biomass with warming. Yet, yielding distinct mean activation energies in response to
increasing temperatures seem to have a larger impact on temperature. Interestingly, differences in marine bacterial
heterotrophic bacteria than on protistan grazers (Sarmento community activation energies with water column depth
et al., 2010; Von Scheibner et al., 2014), suggesting that have been related to the quality of organic carbon com-
the increase in bacterial biomass may exceed mortality pounds (Lønborg et al., 2016), as less labile C would
caused by grazing under warmer ocean conditions. require a higher demand of energy for its breakdown and
Regardless of the potential effect of grazing, the utilization according to the Carbon Quality Theory (CQT,
observed enhancement of bacterial metabolism with tem- Davidson et al., 2000; Davidson and Janssens, 2006).
perature was mostly observed in spring, with E values for Although C consumption by the copiotrophic groups in our
Rhodobacteraceae, Gammaproteobacteria and Bacteroi- short incubations would likely depend solely on labile C
detes even exceeding the theoretical value predicted by substrates, several works have shown that different taxa
the MTE (Brown et al., 2004). The sBP values, used in this are specialized in the use of particular compounds (Cottrell
work as a proxy of carbon bioavailability (Fouilland et al., and Kirchman, 2000; Alonso-Sa ez and Gasol, 2007; Teel-
2014), were also maximum in spring, which suggests that ing et al., 2012), which may require different enzymes for
the response to temperature of all phylogenetic groups their uptake and degradation. Evidence that some
was favoured in periods of resource replete conditions. biochemical pathways are more sensitive to temperature
Our results agree with previous works reporting a tight than others was recently demonstrated for extracellular
relation between the temperature-dependence of bacterial enzymes used for degrading specific DOM compounds in
metabolism and resource availability (Lo pez-Urrutia and a global study in the tropical and subtropical ocean (Ayo
Mora  n 2007; Berggren et al., 2010). Yet, we cannot rule et al., 2017). We speculate that such differences in the pat-
out that the observed higher temperature-response in terns of substrate utilization by distinct bacterial groups
spring may also be due to other processes, such as sea- may ultimately influence their growth activation energies, a
sonal acclimation over the annual cycle. Accordingly, hypothesis that will be interesting to test in future studies.
under the relatively low temperatures in spring, warming In summary, our work agrees with previous studies sug-
may have had a higher impact than in summer, when the gesting that warming may lead to important changes in
C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
ez and X. A. G. Mora
4500 N. Arandia-Gorostidi, T. M. Huete-Stauffer, L. Alonso-Sa n

bacterial community composition (Lindh et al., 2013; FLX1 and the resulting sequences were analysed by the
Bergen et al., 2016), and further illustrates that such alter- MOTHUR platform (Schloss et al., 2009).
ation may be related to the different metabolic response to Samples were transported to the laboratory in darkened
20 L polycarbonate transparent bottles (Nalgene, Rochester,
warming observed for distinct bacterial taxa. Thus,
NY, USA), avoiding direct sunlight exposure, within 6 h of col-
resource availability and taxonomic composition may be
lection. Once in the laboratory, 6 polycarbonate bottles of 4 L
key factors in the final response of bacterial communities (Nalgene, Rochester, NY, USA) were filled with 2 L of sampled
to increasing temperatures. Although here we tested the water, and placed in duplicates in temperature-controlled incu-
effect of temperature on broad phylogenetic groups, the bators set at in situ temperature, 38C above and 38C below
role of more specific taxa, including low abundance and the in situ value of each month. The incubations were carried
rare phylotypes should also be considered in future works, out with the natural photoperiod of the sampling date and
under saturating photosynthetically active radiation (PAR) irra-
as both can have significant roles in biogeochemical cycles
diance (ca., 150 mmol photons m22 s21). The experiments
(Morris et al., 2005; Sauret et al., 2014). On the other lasted between 3 and 7 days (Supporting Information Table
hand, predicting the effect of gradual ocean warming on S3) until total bacterial abundance declined after the station-
the future composition of bacterial communities will require ary phase and the exponential growth phase was clearly
considering other factors such as long-term acclimation, distinguishable. We measured the abundances of heterotro-
interactions with other microorganisms and species evolu- phic bacteria (distinguishing between low and high nucleic
acid content cells) and heterotrophic nanoflagellate abun-
tion. Thus, in order to unveil which species substitution will
dance during the incubation with a FACSCalibur flow-
be dominant in a future ocean scenario, sustained
cytometer (BD Biosciences, Heidelberg, Germany) using
long-term observations on the evolution of bacterioplank- common protocols (Gasol and Mora  n, 2015), specifically
ton assemblages will be required. The differential following Calvo-Dı́az and Mora  n (2006) for bacteria and
temperature sensitivity of some of the main phylogenetic Christaki et al. (2011) for HNFs. The efficiency of HNF
groups present in coastal, temperate waters along the removal by the 0.8 mm cartridge was assessed by counting
productive season found in this study, clearly indicates before and after pre-filtration. During the incubations, we sam-
pled once or twice per day, depending on the growth phase, to
that taxon-specific studies may be key to understand the
determine the abundance of target bacterial groups by CARD-
future dynamics and biogeochemical function of marine FISH analysis.
bacterioplankton.

Experimental procedures Bacterial heterotrophic production


Bacterial heterotrophic production was measured based on
Sample collection and temperature incubations
[3H]-leucine incorporation rates. Leucine was added at satu-
The sampling site is located 13 km off the coast of Gijon/Xixon rating concentration (40 nmol L21) to three experimental
(Spain), in the Southern Bay of Biscay (43.6758N, 5.5788W). replicates (1.2 mL) and two controls, treated with 120 mL 50%
This station is part of a transect regularly monitored within the TCA prior to isotope addition. Samples were incubated from 1
programme Radiales of the Spanish Institute of Oceanogra- to 3 h at in situ temperature in temperature-controlled incuba-
phy (IEO). Seawater was sampled monthly between January tors. The incorporation was stopped with the addition of 120
and December 2012 from 5 m depth with 5 L Niskin bottles mL 50% TCA and the tubes were frozen until analysis following
(Nalgene, Rochester, NY, USA). Seawater collections were the centrifugation method (Smith and Azam, 1992). Finally,
carried out during the first hours of the morning, in order to 1 mL of scintillation cocktail was added to the tubes and they
avoid diurnal variations of the microbial community. After the were counted on a Beckman scintillation counter. Leucine
collection, seawater samples were pre-filtered by 0.8 mm pore- incorporation rates were normalized by total bacterial abun-
size cartridges (PALL corporation, East Hills, NY, USA) to iso- dance to estimate cell-specific bacterial production (sBP, pmol
late the heterotrophic bacteria from predators, as used in Leu cell21 h21)
many previous works (e.g., Williams, 1981; Malmstrom et al.,
2007; Lami and Kirchman, 2014). In situ oceanographic condi-
Catalysed reporter deposition fluorescent in situ
tions were measured by a SeaBird 25 CTD. Ambient
hybridization (CARD-FISH)
chlorophyll a concentrations were determined in the lab prior
to the pre-filtration steps by filtering 200 ml subsamples on 0.2 CARD-FISH analysis (Pernthaler et al., 2002) was carried out
mm pore-size polycarbonate filters. Filters were frozen at to estimate the abundance of specific phylogenetic groups of
2208C and processed within 2 weeks as explained in Calvo- bacterioplankton. Samples were fixed with 4% formaldehyde
Dı́az and Mora  n (2006). Samples for DNA sequencing were solution (Sigma-Aldrich, St Louis, MO, USA) for 3 h at room
collected onto 0.2 mm filters as explained in Alonso-Sa ez et al. temperature, and filtered through 0.2 mm pore-size polycar-
(2015). DNA extraction was carried out using the PowerWater bonate filters (GTTP type, 25 mm; Millipore). Filters were
DNA isolation Kit (Mobio, Carlsbad, CA, USA). 16S rDNA dipped in 0.1% (w/w) agarose and permeabilized in lysozyme
genes were amplified using the general bacterial primers (10 mg mL21, Sigma-Aldrich, St Louis, MO, USA) at 378C for
341F and 805R (Herlemann et al., 2011) following Alonso- 1 h, and achromopeptidase (60 U mL21, 0.01 M NaCl, 0.01 M
Saez et al. (2015). Amplicons were sequenced using a 454 Tris-HCl, pH 7.6, Sigma-Aldrich), for 30 min at 378C.

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
Metabolic theory of ecology with marine bacteria 4501
Hybridizations were carried out overnight at 358C using Horse- can be described as the sum of such individual rates with
radish Peroxidase (HRP) labelled probes (50 ng mL21) diluted mass and temperature correction:
in 900 mL of hybridization buffer, with different formamide
concentrations. The probes and formamide concentrations X I M b 2E=k T
used are described below; SAR11–441R (Morris et al., 2002, l5 ’ e (2)
M M
45% formamide) for SAR11, Ros537 (Eilers et al., 2001, 55%
formamide) for Rhodobacteraceae, Gam42a (Manz et al., where the activation energy is the slope of the ln transformed
1996) with its Betaproteobacteria competitor Bet42a (Manz temperature-corrected growth rates against the temperature
et al., 1992, 55% formamide) for Gammaproteobacteria, and factor (1/kT). Regarding mass correction, the allometric scal-
CF319 (Manz et al., 1996, 55% formamide) for Bacteroidetes. ing exponent (b) has been empirically estimated as b 5 1/4 for
Hybridized samples were amplified (468C, 30 min) using tyra- population growth (Slobodkin, 1962; Blueweiss et al., 1978)
mide labelled with Alexa 488 dye (1 mg mL21, 468C, 30 min, and as b 5 3/4 for carrying capacity (Damuth, 1987; Belgrano
ThermoFisher, Germany), dried and stored at 2808C. Filters et al., 2002). However, such universal values of the exponent
were transferred onto slides and stained with DAPI at 1 have been questioned (Glazier, 2015), and due to the small
mg mL21. DAPI and CARD-FISH stained bacteria were quanti- size variability of bacteria (usually less than 0.01 mm3) we did
fied using an epifluorescence microscope Leica DM5500B, a not perform any mass-correction as discussed by Huete-
monochromatic camera Leica DFC360 FX and the ACMETool Stauffer et al. (2015).
(www.technobiology.ch) automatized image analysis software The effect of temperature on the carrying capacity (K-TR)
(Zeder and Pernthaler, 2009; Zeder et al., 2011). Counts of was measured as the linear slope between the maximum bac-
hybridized bacteria are reported as percentage of total DAPI terial abundance (ln-transformed) and temperature (in 8C).
stained cells, calculated from 10 randomly chosen micro- Mass-correction was omitted for calculating carrying capacity
scopic fields. temperature dependence. Within this linear relationship, posi-
tive values indicate higher carrying capacity under warmer
conditions, while negative values indicate a decrease in carry-
Growth rates, carrying capacity, cell-size and ing capacity with increasing temperature.
temperature dependence
We used the slope of ln-transformed abundance versus time Acknowledgements
during the exponential growth phase to calculate the growth
We are grateful to Basque Government for supporting
rates in duplicate seawater incubations. The carrying capacity
N.A.G.’s PhD fellowship and to the Spanish Ministry of
(K) was calculated as the maximum bacterial abundance Economy and Competitiveness (MINECO) for supporting
reached at stationary grow phase by each phylogenetic group T.M.H-S.’s PhD fellowship, L.A.S.’s Juan de la Cierva and
and temperature treatment. Cell-sizes were calculated for the Ramon y Cajal fellowships and the COMITE project (CTM-
initial time of the incubations for the different phylogenetic 2010-15840). The Spanish Institute of Oceanography (IEO)
groups. Microscopy images were transformed to binary time-series programme Radiales provided logistic support for
images in order to minimize the brightness of DAPI signal in seawater collection. We are especially thankful to all the staff
the surrounding area of the cell, as described in Massana of the R/V ‘Jose de Rioja’ for their help during the sampling
et al. (1997). After transformation, biovolume was calculated collection and L. Dı́az, A. Calvo-Dı́az and E. Nogueira for their
by an automatic analysis system based in R programming lan- help during the experiments. The authors have declared that
guage, which calculates the size of CARD-FISH positive cells no competing interests exist.
for DAPI transformed images using the algorithm by Baldwin
and Bankston (1988). Cells with volumes smaller than 0.0042
mm3 and larger than 0.344 mm3 were excluded from the analy-
References
sis, as they were considered out of the range of typical Alonso-Saez, L., and Gasol, J.M. (2007) Seasonal variations
bacterioplankton cells (Straza et al., 2009). The image pre- in the contributions of different bacterial groups to the
processing method applied in this work was different from uptake of low-molecular-weight compounds in northwestern
than that reported in Mora n et al. (2015), producing some var- Mediterranean coastal waters. Appl Environ Microbiol 73:
iations in the SAR11 cell-size reported for the same samples 3528–3535.
between the current work and the one by Mora  n et al., (2015). Alonso-Saez, L., Diaz-Perez, L., and Mora  n, X.A.G. (2015)
To calculate the temperature dependence of bacterial The hidden seasonality of the rare biosphere in coastal
growth rates, we used the activation energy (E) as in the MTE marine bacterioplankton. Environ Microbiol 17: 3766–3780.
basic equation: Amann, R., and Fuchs, B.M. (2008) Single-cell identification
in microbial communities by improved fluorescence in situ
I5I 0 M b e 2E=k T (1) hybridization techniques. Nat Rev Microbiol 6: 339–348.
Arrhenius, S. (1889) Uber die Reaktionsgeschwindigkeit bei
Where I is the individual metabolic rate, I0 is the normalization der Inversion von Rohrzucker durcj Sauren. Z Phys Chem
constant, b is an allometric scaling exponent, M and T corre- 4: 226–248.
spond to the body-size and temperature (in K) and k Ayo, B., Abad, N., Artolozaga, I., Azua, I., Bana, Z., Unanue,
represents the Boltzmanns constant. According to Savage M., et al. (2017) Imbalanced nutrient recycling in a warmer
et al. (2004), the growth rate of an entire population is strictly ocean driven by differential response of extracellular enzy-
dependent of individual metabolic rates. Thus, growth rates matic activities. Glob Chang Biol.

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
ez and X. A. G. Mora
4502 N. Arandia-Gorostidi, T. M. Huete-Stauffer, L. Alonso-Sa n

Azam, F., and Long, R.A. (2001) Sea snow microcosms. Collins, M., Knutti, R., Arblaster, J., Dufresne, J.-L., Fichefet,
Nature 414: 495, 497–498. T., Friedlingstein, P., et al. (2013) Long-term climate
Bailly, D., Cassemiro, F.A.S., Agostinho, C.S., Marques, E.E., change, projections, commitments and irreversibility. In Cli-
and Agostinho, A.A. (2014) The metabolic theory of ecology mate Change 2013, The Physical Science Basis. Contribu-
convincingly explains the latitudinal diversity gradient of tion of Working Group I to the Fifth Assessment Report of
Neotropical freshwater fish. Ecology 95: 553–562. the Intergovernmental Panel on Climate Change. Stocker,
Baldwin, W.W, and Rivers, A.R. (1988) Measurement of live T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K.,
bacteria by nomarski interference microscopy and ste- reo- Boschung, J., et al. (eds). New York, NY: Cambridge Uni-
logic methods as tested with macroscopic rod-shaped mod- versity Press, pp. 1029–1136.
els. Appl Environ Microbiol 54: 105–109. Cottrell, M.T., and Kirchman, D.L. (2000) Natural assemblages
Beier, S., Rivers, A.R., Moran, M.A., and Obernosterer, I. of marine proteobacteria and members of the Cytophaga-
(2015) Phenotypic plasticity in heterotrophic marine Flavobacter cluster consuming low- and high-molecular-
microbial communities in continuous cultures. ISME J 9: weight dissolved organic matter. Appl Environ Microbiol 66:
1141–1151. 1692–1697.
Belgrano, A., Allen, A.P., Enquist, B.J., and Gillooly, J.F. Damuth, J. (1987) Interspecific allometry of population-density
(2002) Allometric scaling of maximum population density: a in mammals and other animals: the independence of body-
common rule for marine phytoplankton and terrestrial mass and population energy-use. Bot J Linnean Soc 31:
plants. Ecol Lett 5: 611–613. 193–246.
Bergen, B., Endres, S., Engel, A., Zark, M., Dittmar, T., Daufresne, M., Lengfellner, K., and Sommer, U. (2009) Global
Sommer, U., and Jurgens, K. (2016) Acidification and warm- warming benefits the small in aquatic ecosystems. Proc
ing affect prominent bacteria in two seasonal phytoplankton Natl Acad Sci USA 106: 12788–12793.
bloom mesocosms. Environ Microbiol 18: 4579–4595. Davidson, E.A., and Janssens, I.A. (2006) Temperature sensi-
Berggren, M., Laudon, H., Jonsson, A., and Jansson, M. tivity of soil carbon decomposition and feedbacks to climate
(2010) Nutrient constraints on metabolism affect the tem- change. Nature 440: 165–173.
perature regulation of aquatic bacterial growth efficiency. Davidson, E.A., Trumbore, S.E., and Amundson, R. (2000) Soil
Microb Ecol 60: 894–902. warming and organic carbon content. Nature 408: 789–790.
Blueweiss, L., Fox, H., Kudzma, V., Nakashima, D., Peters, Degerman, R., Dinasquet, J., Riemann, L., Sjo €stedt de Luna,
R., and Sams, S. (1978) Relationships between body size S., and Andersson, A. (2013) Effect of resource availability
and some life history parameters. Oecologia 37: 257–272. on bacterial community responses to increased tempera-
Brown, J.H., Gillooly, J.F., Allen, A.P., Savage, V.M., and West, ture. Aquat Microb Ecol 68: 131–142.
G.B. (2004) Toward a metabolic theory of ecology. Ecology Edwards, M., and Richardson, A.J. (2004) Impact of climate
85: 1771–1789. change on marine pelagic phenology and trophic mismatch.
Buchan, A., LeCleir, G.R., Gulvik, C.A., and Gonzalez, J.M. Nature 430: 881–884.
(2014) Master recyclers: features and functions of bacteria Eiler, A., Langenheder, S., Bertilsson, S., and Tranvik, L.J.
associated with phytoplankton blooms. Nat Rev Microbiol (2003) Heterotrophic bacterial growth efficiency and com-
12: 686–698. munity structure at different natural organic carbon concen-
Bunse, C., and Pinhassi, J. (2017) Marine bacterioplankton trations. Appl Environ Microbiol 69: 3701–3709.
seasonal succession dynamics. Trends Microbiol 25: Eilers, H., Pernthaler, J., Peplies, J., Glockner, F.O., Gerdts,
494–505. G., and Amann, R. (2001) Isolation of novel pelagic bacteria
Calvo-Dı́az, A. (2008) Estructura de la Comunidad Pico- from the German bight and their seasonal contributions to
planctonica Y Produccio n Heterotro fica Bacteriana En La surface picoplankton. Appl Environ Microbiol 67: 5134–
Plataforma Continental Asturiana. PhD Thesis. Oviedo, 5142.
Spain: University of Oviedo. Elias, M., Wieczorek, G., Rosenne, S., and Tawfik, D.S.
Calvo-Dı́az, A., and Mora  n, X.A.G. (2006) Seasonal dynamics (2014) The universality of enzymatic rate- temperature
of picoplankton in shelf waters of the southern Bay of Bis- dependency. Trends Biochem Sci 39: 1–7.
cay. Aquat Microb Ecol 42: 159–174. Elifantz, H., Malmstrom, R.R., Cottrell, M.T., and Kirchman,
Campbell, B.J., and Kirchman, D.L. (2013) Bacterial diversity, D.L. (2005) Assimilation of polysaccharides and glucose by
community structure and potential growth rates along an major bacterial groups in the Delaware Estuary. Appl Envi-
estuarine salinity gradient. ISME J 7: 210–220. ron Microbiol 71: 7799–7805.
Cho, B.C., and Azam, F. (1990) Biogeochemical significance Ferrera, I., Gasol, J.M., Sebastian, M., Hojerova, E., and
of bacterial biomass in the ocean’s euphotic zone. Mar Ecol Koblizek, M. (2011) Comparison of growth rates of aerobic
Prog Ser 63: 253–259. anoxygenic phototrophic bacteria and other bacterioplank-
Cho, J.C., and Giovannoni, S.J. (2004) Cultivation and Growth ton groups in coastal Mediterranean waters. Appl Environ
Characteristics of a Diverse Group of Oligotrophic Marine Microbiol 77: 7451–7458.
Gammaproteobacteria. Appl Environ Microbiol 70: Fouilland, E., Tolosa, I., Bonnet, D., Bouvier, C., Bouvier, T.,
432–440. Bouvy, M., et al. (2014) Bacterial carbon dependence on
Christaki, U., Courties, C., Massana, R., Catala, P., Lebaron, freshly produced phytoplankton exudates under different
P., Gasol, J.M., and Zubkov, M.V. (2011) Optimized routine nutrient availability and grazing pressure conditions in
flow cytometric enumeration of heterotrophic flagellates coastal marine waters. FEMS Microbiol Ecol 87: 757–769.
using SYBR Green I. Limnol Oceanogr Methods 9: Gasol, J.M., and Mora  n, X.A.G. (2015) Flow cytometric deter-
329–339. mination of microbial abundances and its use to obtain

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
Metabolic theory of ecology with marine bacteria 4503
indices of community structure and relative activity. In pure cultures and in the Delaware estuary. ISME J 10: 823–
Hydrocarbon and Lipid Microbiology Protocols. McGenity, 832.
T.J., Timmis, K.N., and Nogales B. (eds). Berlin, Heidel- Lauro, F.M., McDougald, D., Thomas, T., Williams, T.J., Egan,
berg: Springer, pp. 159–187. S., Rice, S., et al. (2009) The genomic basis of trophic strat-
Gifford, S.M., Sharma, S., Booth, M., and Moran, M.A. (2013) egy in marine bacteria. Proc Natl Acad Sci USA 106:
Expression patterns reveal niche diversification in a marine 15527–15533.
microbial assemblage. ISME J 7: 281–298. Laws, E.A., Falkowski, P.G., Smith, W.O., Ducklow, H., and
Gifford, S.M., Sharma, S., and Moran, M.A. (2014) Linking McCarthy, J.J. (2000) Temperature effects on export pro-
activity and function to ecosystem dynamics in a coastal duction in the open ocean. Glob Biogeochem Cycles 14:
bacterioplankton community. Front Microbiol 5: 185. 1231–1246.
Gillooly, J.F., Brown, J.H., West, G.B., Savage, V.M., and Li, W.K.W., Head, E.J.H., and Glen Harrison, W. (2004) Mac-
Charnov, E.L. (2001) Effects of size and temperature on roecological limits of heterotrophic bacterial abundance in
metabolic rate. Science 293: 2248–2251. the ocean. Deep Sea Res Part 1 Oceanogr Res Pap 51:
Giovannoni, S.J. (2017) SAR11 bacteria: the most abundant 1529–1540.
plankton in the oceans. Ann Rev Mar Sci 9: 231–255. Lindh, M.V., Riemann, L., Baltar, F., Romero-Oliva, C.,
Giovannoni, S.J., Bibbs, L., Cho, J.C., Stapels, M.D., Salomon, P.S., Graneli, E., and Pinhassi, J. (2013) Conse-
Desiderio, R., Vergin, K.L., et al. (2005) Proteorhodopsin in quences of increased temperature and acidification on bac-
the ubiquitous marine bacterium SAR11. Nature 438: terioplankton community composition during a mesocosm
82–85. spring bloom in the Baltic Sea. Environ Microbiol Rep 5:
Glazier, D.S. (2015) Is metabolic rate a universal ’pacemaker’ 252–262.
for biological processes?. Biol Rev Camb Philos Soc 90: Lønborg, C., Cuevas, L.A., Reinthaler, T., Herndl, G.J., Gasol,
377–407. J.M., Mora  n, X.A.G., et al. (2016) Depth dependent rela-
Grote, J., Thrash, J.C., Huggett, M.J., Landry, Z.C., Carini, P., tionships between temperature and ocean heterotrophic
Giovannoni, S.J., and Rappe, M.S. (2012) Streamlining and prokaryotic production. Front Mar Sci 3: 90.
core genome conservation among highly divergent mem- pez-Pe
Lo rez, M., Gonzaga, A., Martin-Cuadrado, A.B.,
bers of the SAR11 clade. mBio 3: e00252–12. Onyshchenko, O., Ghavidel, A., Ghai, R., et al. (2012)
Hall, E.K., Singer , G.A., Kainz, M.J., and Lennon, J.T. (2010) Genomes of surface isolates of Alteromonas macleodii: the
Evidence for a temperature acclimation mechanism in bac- life of a widespread marine opportunistic copiotroph. Sci
teria: an empirical test of a membrane-mediated trade-off. Rep 2: 696.
Funct Ecol 24: 898–908. pez-Urrutia, A.,
Lo  and Mora  n, X.A.G. (2007) Resource limita-
Herlemann, D.P., Labrenz, M., Ju €rgens, K., Bertilsson, S., tion of bacterial production distorts the temperature depen-
Waniek, J.J., and Andersson, A.F., (2011) Transitions in dence of oceanic carbon cycling. Ecology 88: 817–822.
bacterial communities along the 2000 km salinity gradient pez-Urrutia, A., San Martin, E., Harris, R.P., and Irigoien, X.
Lo
of the Baltic Sea. ISME J 5: 1571–1579. (2006) Scaling the metabolic balance of the oceans. Proc
Huete-Stauffer, T.M., Arandia-Gorostidi, N., Diaz-Perez, L., Natl Acad Sci USA 103: 8739–8744.
and Moran, X.A. (2015) Temperature dependences of Malmstrom, R.R., Kiene, R.P., Cottrell, M.T., and Kirchman,
growth rates and carrying capacities of marine bacteria D.L. (2004) Contribution of SAR11 bacteria to dissolved
depart from metabolic theoretical predictions. FEMS Micro- dimethylsulfoniopropionate and amino acid uptake in the
biol Ecol 91: fiv11. North Atlantic Ocean. Appl Environ Microbiol 70: 4129–4135.
Jiang, L., and Morin, P.J. (2004) Temperature-dependent inter- Malmstrom, R.R., Straza, T.R.A., Cottrell, M.T., and Kirchman,
actions explain unexpected responses to environmental D.L. (2007) Diversity, abundance, and biomass production
warming in communities of competitors. J Anim Ecol 73: of bacterial groups in the western Arctic Ocean. Aquat
569–576. Microb Ecol 47: 45–55.
Kingsolver, J.G., and Huey, R.B. (2008) Size, temperature, Manz, W., Amann, R., Ludwig, W., Wagner, M., and Schleifer,
and fitness: three rules. Evol Ecol Res 10: 251–268. K.-H. (1992) Phylogenetic oligodeoxynucleotide probes for
Kirchman, D. (2002) The ecology of Cytophaga–Flavobacte- the major subclasses of proteobacteria: problems and solu-
ria in aquatic environments. FEMS Microbiol Ecol 39: tions. Syst Appl Microbiol 15: 593–600.
91–100. Manz, W., Amann, R., Ludwig, W., Vancanneyt, M., and
Kirchman, D.L. (2016) Growth rates of microbes in the Schleifer, K.H. (1996) Application of a suite of 16S rRNA-
oceans. Ann Rev Mar Sci 8: 285–309. specific oligonucleotide probes designed to investigate bac-
Kirchman, D.L., Malmstrom, R.R., and Cottrell, M.T. (2005) teria of the phylum cytophaga-flavobacter-bacteroides in
Control of bacterial growth by temperature and organic mat- the natural environment. Microbiology 142: 1097–1106.
ter in the Western Arctic. Deep Sea Res Part 2 Top Stud Massana, R., Murray, A.E., Preston, C.M., and DeLong, E.F.
Oceanogr 52: 3386–3395. (1997) Vertical distribution and phylogenetic characteriza-
Koch, A.L. (2001) Oligotrophs versus copiotrophs. Bioessays tion of marine planktonic Archaea in the Santa Barbara
23: 657–661. Channel. Appl Environ Microbiol 63: 50–56.
Lami, R., and Kirchman, D.L. (2014) Diurnal expression of Meehl, G.A., Stocker, T.F., Collins, W.D., Friedlingstein, P.,
SAR11 proteorhodopsin and 16S rRNA genes in coastal Gaye, A.T., Gregory, J.M., et al. (2007) Global climate pro-
North Atlantic waters. Aquat Microb Ecol 73: 185–194. jections. In Solomon, S., Qin, M.M.D., Chen Z., Marquis,
Lankiewicz, T.S., Cottrell, M.T., and Kirchman, D.L. (2016) M., Averyt, K.B., Tignor, M., Miller, H.L. (eds) Climate
Growth rates and rRNA content of four marine bacteria in change 2007: the physical science basis. Contribution of

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
ez and X. A. G. Mora
4504 N. Arandia-Gorostidi, T. M. Huete-Stauffer, L. Alonso-Sa n

working group I to the fourth assessment report of the inter- Sauret, C., Severin, T., Vetion, G., Guigue, C., Goutx, M.,
governmental panel on climate change. Cambridge: Cam- Pujo-Pay, M., et al. (2014) ’Rare biosphere’ bacteria as key
bridge University Press, pp. 747–845. phenanthrene degraders in coastal seawaters. Environ
Moran, M.A., Buchan, A., Gonzalez, J.M., Heidelberg, J.F., Pollut 194: 246–253.
Whitman, W.B., Kiene, R.P., et al. (2004) Genome Savage, V.M., Gilloly, J.F., Brown, J.H., and Charnov, E.L.
sequence of Silicibacter pomeroyi reveals adaptations to (2004) Effects of body size and temperature on population
the marine environment. Nature 432: 910–913. growth. Am Nat 163: 429–441.
Mora n, X.A.G., Bode, A., Sua rez, L.A., and Nogueira, E. Schattenhofer, M., Fuchs, B.M., Amann, R., Zubkov, M.V.,
(2007) Assessing the relevance of nucleic acid content as Tarran, G.A., and Pernthaler, J. (2009) Latitudinal distribu-
an indicator of marine bacterial activity. Aquat Microb Ecol tion of prokaryotic picoplankton populations in the Atlantic
46: 141–152. Ocean. Environ Microbiol 11: 2078–2093.
Mora n, X.A.G., Ducklow, H.W., and Erickson, M. (2011) Sin- Schloss, P.D., Westcott, S.L., Ryabin, T., Hall, J.R., Hartmann,
gle-cell physiological structure and growth rates of hetero- M., Hollister, E.B., et al. (2009) Introducing mothur:
trophic bacteria in a temperate estuary (Waquoit Bay, open-source, platform-independent, community-supported
Massachusetts). Limnol Oceanogr 56: 37–48. software for describing and comparing microbial communi-
Mora n, X.A., Alonso-Sa ez, L., Nogueira, E., Ducklow, H.W., ties. Appl Environ Microbiol 75: 7537–7541.
Gonzalez, N., Lopez-Urrutia, A., et al. (2015) More, smaller Sinsabaugh, R.L., and Shah, J.J.F. (2010) Integrating
bacteria in response to ocean’s warming? Philos Trans R resource utilization and temperature in metabolic scaling of
Soc Lond B Biol Sci 282: 20150371. riverine bacterial production. Ecology 91: 1455–1465.
Morris, R.M., Rappe, M.S., Connon, S.A., Vergin, K.L., Slobodkin, L.B. (1962) Growth and Regulation of Animal
Siebold, W.A., Carlson, C.A., et al. (2002) SAR11 clade Populations. New York, NY: Holt, Reinhart, and Winston.
dominates ocean surface bacterioplankton communities. Smith, D.C., and Azam, F. (1992) A simple, economical
Nature 420: 806–810. method for measuring bacterial protein 717 synthesis rates
Morris, R.M., Vergin, K.L., Cho, J.-C., Rappe , M.S., Carlson, in seawater using 3H-leucine. Mar Microb Food Webs 6:
C.A., and Giovannoni, S.J. (2005) Temporal and spatial 107–114.
response of bacterioplankton lineages to annual convective 
Solić, M., Krstulovic, N., Vilibic, I., Bojanic, N., Kuspipilic, G.,
overturn at the Bermuda Atlantic Time-series Study site. Sestanovic, S., et al. (2009) Variability in the bottom-up and
Limnol Oceanogr 50: 1687–1696. top-down controls of bacteria on trophic and temporal
Nelson, C.E., and Carlson, C.A. (2012) Tracking differential scales in the middle Adriatic Sea. Aquat Microb Ecol 58:
incorporation of dissolved organic carbon types among 15–29.
diverse lineages of Sargasso Sea bacterioplankton. Environ Sowell, S.M., Wilhelm, L.J., Norbeck, A.D., Lipton, M.S.,
Microbiol 14: 1500–1516. Nicora, C.D., Barofsky, D.F., et al. (2009) Transport func-
O’Connor, M.I., Piehler, M.F., Leech, D.M., Anton, A., and tions dominate the SAR11 metaproteome at low-nutrient
Bruno, J.F. (2009) Warming and resource availability shift extremes in the Sargasso Sea. ISME J 3: 93–105.
food web structure and metabolism. PLOS Biol 7: Straza, T.R., Cottrell, M.T., Ducklow, H.W., and Kirchman, D.L.
e1000178. (2009) Geographic and phylogenetic variation in bacterial
Pernthaler, A., Pernthaler, J., and Amann, R. (2002) Fluores- biovolume as revealed by protein and nucleic acid staining.
cence in situ hybridization and catalyzed reporter deposition Appl Environ Microbiol 75: 4028–4034.
for the identification of marine bacteria. Appl Environ Micro- Teeling, H., Fuchs, B.M., Becher, D., Klockow, C.,
biol 68: 3094–3101. Gardebrecht, A., Bennke, C.M., et al. (2012) Substrate-
Pinhassi, J., and Hagstro € m, Å. (2000) Seasonal succes- controlled succession of marine bacterioplankton popula-
sion in marine bacterioplankton. Aquat Microb Ecol 21: tions induced by a phytoplankton bloom. Science 336:
245–256. 608–611.
Polz, M.F., Hunt, D.E., Preheim, S.P., and Weinreich, D.M. Teira, E., Martinez-Garcia, S., Lonborg, C., and Alvarez-
(2006) Patterns and mechanisms of genetic and phenotypic Salgado, X.A. (2009) Growth rates of different phylogenetic
differentiation in marine microbes. Philos Trans R Soc Lond bacterioplankton groups in a coastal upwelling system.
B Biol Sci 361: 2009–2021. Environ Microbiol Rep 1: 545–554.
Ratkowsky, D.A., Olley, J., McMeekin, T.A., and Ball, A. Vaque  Peters, F., Felipe, J., Malits, A., and
, D., Guadayol, O.,
(1982) Relationship between temperature and growth rate Pedro  s-Alio
, C. (2009) Differential response of grazing and
of bacterial cultures. J Bacteriol 149: 1–5. bacterial heterotrophic production to experimental warming
Regaudie-de-Gioux, A., and Duarte, C.M. (2012) Temperature in Antarctic waters. Aquat Microb Ecol 54: 101–112.
dependence of planktonic metabolism in the ocean. Glob von Scheibner, M., Dorge, P., Biermann, A., Sommer, U.,
Biogeochem Cycles 26: 1015–1024. Hoppe, H.G., and Jurgens, K. (2014) Impact of warming on
Sarmento, H., Montoya, J.M., Vazquez-Dominguez, E., phyto-bacterioplankton coupling and bacterial community
Vaque, D., and Gasol, J.M. (2010) Warming effects on composition in experimental mesocosms. Environ Microbiol
marine microbial food web processes: how far can we go 16: 718–733.
when it comes to predictions? Philos Trans R Soc Lond B West, G.B. (1997) A general model for the origin of allometric
Biol Sci 365: 2137–2149. scaling laws in biology. Science 276: 122–126.
Sarmiento, J.L., Hughes, T.M.C., Stouffer, R.J., and Manabe, Whitman, W.B., Coleman, D.C., and Wiebe, W.J. (1998)
S. (1998) Simulated response of the ocean carbon cycle to Prokaryotes: the unseen majority. Proc Natl Acad Sci U S A
anthropogenic climate warming. Nature 393: 245–249. 95: 6578–6583.

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V
Metabolic theory of ecology with marine bacteria 4505
Williams, P.J.L. (1981) Microbial contribution to overall marine
plankton metabolism: direct measurements of respiration. Fig. S1. In situ concentrations of inorganic nutrients; dis-
Oceanol Acta 4: 359–364. solved inorganic nitrogen (DIN, orange), phosphate (PO4,
Wiltshire, K.H., and Manly, B.F.J. (2004) The warming trend at green) and silicate (SiO4, black)
Helgoland Roads, North Sea: phytoplankton response. Hel- Fig. S2. Initial mean cell-size of the bacterial community
gol Mar Res 58: 269–273. during the 2012 annual cycle. Inset show correlation
Yokokawa, T., and Nagata, T. (2005) Growth and grazing mor- between bacterial biovolume and temperature.
tality rates of phylogenetic groups of bacterioplankton in Fig. S3. Abundance changes with time at three incubation
coastal marine environments. Appl Environ Microbiol 71: temperatures (in situ, 38C below and above in situ) for
6799–6807. SAR11, Rhodobacteraceae, Gammaproteobacteria and
Yokokawa, T., Nagata, T., Cottrell, M.T., and Kirchman, D.L. Bacteroidetes for the experiment of April.
(2004) Growth rate of the major phylogenetic bacterial groups Fig. S4. Contribution of the two most abundant SAR11
in the Delaware estuary. Limnol Oceanogr 49: 1620–1629. OTUs (OTU1 and OTU4) to total SAR11 reads (in percent-
Yooseph, S., Nealson, K.H., Rusch, D.B., McCrow, J.P., age) during the year 2012. Data were obtained by pyrose-
Dupont, C.L., Kim, M., et al. (2010) Genomic and functional quencing as explained in Alonso-Sa ez and colleagues
adaptation in surface ocean planktonic prokaryotes. Nature
(2015)
468: 60–66.
Table S1. Pearson correlation coefficients of the relation-
Zeder, M., and Pernthaler, J. (2009) Multispot live-image auto-
ships between the initial abundance, carrying capacity (K),
focusing for high-throughput microscopy of fluorescently
growth rate (m), activation energy (E) and temperature effect
stained bacteria. Cytometry A 75: 781–788.
on carrying capacity (K-RT) of the different groups, and the
Zeder, M., Ellrott, A., and Amann, R. (2011) Automated sam-
environmental variables temperature (T), initial cell-size and
ple area definition for high- throughput microscopy. Cytome-
cell-specific bacterial heterotrophic production (LIR). Num-
try A 79: 306–310.
bers in parentheses indicate p-values.
Zhao, Y., Temperton, B., Thrash, J.C., Schwalbach, M.S.,
Table S2. Summary statistics of selected parameters of the
Vergin, K.L., Landry, Z.C., et al. (2013) Abundant SAR11
viruses in the ocean. Nature 494: 357–360. four phylogenetic groups considered.
Table S3. Detailed conditions of each incubation
experiment.
Supporting information
Additional Supporting Information may be found in the
online version of this article at the publisher’s web-site:

C 2017 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 19, 4493–4505
V

Potrebbero piacerti anche