ArticlePDF Available

Abstract and Figures

Viruses affect biogeochemical cycling, microbial mortality, gene flow, and metabolic functions in diverse environments through infection and lysis of microorganisms. Fundamental to quantitatively investigating these roles is the determination of viral abundance in both field and laboratory samples. One current, widely-used method to accomplish this in aquatic samples is the 'filter-mount' method in which samples are filtered onto costly 0.02 μm-pore-size ceramic filters for enumeration of viruses with epifluorescence microscopy. Here we describe a cost-effective (ca. 500-fold lower materials cost) alternative virus enumeration method in which fluorescently-stained samples are wet-mounted directly onto slides, after optional chemical flocculation of viruses in samples with viral concentrations <5×10(7) mL(-1). The concentration of viruses in the sample is then determined from the ratio of viruses to a known concentration of added microsphere beads via epifluorescence microscopy. Virus concentrations obtained using this wet-mount method, with and without chemical flocculation, were significantly correlated with, and had equivalent precision to, those from the filter-mount method across concentrations ranging from 2.17×10(6) to 1.37×10(8) viruses mL(-1) when tested using cultivated viral isolates and natural samples from marine and freshwater environments. In summary, the wet-mount method is significantly less expensive than the filter-mount method, and is appropriate for rapid, precise and accurate enumeration of aquatic viruses over a wide range of viral concentrations (≥1×10(6) viruses mL(-1)) encountered in field and laboratory samples. Copyright © 2015, American Society for Microbiology. All Rights Reserved.
This content is subject to copyright. Terms and conditions apply.
An Inexpensive, Accurate, and Precise Wet-Mount Method for
Enumerating Aquatic Viruses
Brady R. Cunningham,
a
Jennifer R. Brum,
b
Sarah M. Schwenck,
b
Matthew B. Sullivan,
b
Seth G. John
a
Department of Earth and Ocean Sciences, University of South Carolina, Columbia, South Carolina, USA
a
; Department of Ecology and Evolutionary Biology, University of
Arizona, Tucson, Arizona, USA
b
Viruses affect biogeochemical cycling, microbial mortality, gene flow, and metabolic functions in diverse environments through
infection and lysis of microorganisms. Fundamental to quantitatively investigating these roles is the determination of viral
abundance in both field and laboratory samples. One current, widely used method to accomplish this with aquatic samples is the
“filter mount” method, in which samples are filtered onto costly 0.02-m-pore-size ceramic filters for enumeration of viruses by
epifluorescence microscopy. Here we describe a cost-effective (ca. 500-fold-lower materials cost) alternative virus enumeration
method in which fluorescently stained samples are wet mounted directly onto slides, after optional chemical flocculation of vi-
ruses in samples with viral concentrations of <510
7
viruses ml
1
. The concentration of viruses in the sample is then deter-
mined from the ratio of viruses to a known concentration of added microsphere beads via epifluorescence microscopy. Virus
concentrations obtained by using this wet-mount method, with and without chemical flocculation, were significantly correlated
with, and had precision equivalent to, those obtained by the filter mount method across concentrations ranging from 2.17 10
6
to 1.37 10
8
viruses ml
1
when tested by using cultivated viral isolates and natural samples from marine and freshwater envi-
ronments. In summary, the wet-mount method is significantly less expensive than the filter mount method and is appropriate
for rapid, precise, and accurate enumeration of aquatic viruses over a wide range of viral concentrations (>110
6
viruses ml
1
)
encountered in field and laboratory samples.
Viruses are the most abundant biological entities in aquatic sys-
tems, and their infection of microorganisms has substantial in-
fluences on microbial ecology, biogeochemical cycling, and gene
transfer in aquatic environments (reviewed in references 1and 2). An
accurate method to quantify aquatic viruses is thus essential for use in
field and laboratory studies to investigate the roles of viruses in
aquatic environments. Enumeration of viruses in aquatic samples has
previously been accomplished by using transmission electron mi-
croscopy (TEM) (3), epifluorescence microscopy (reviewed in refer-
ence 4), and flow cytometry (reviewed in reference 5).
While each of the above-mentioned methods requires the use of
relatively expensive laboratory equipment, the per-sample cost of the
widely used epifluorescence microscopy method has recently in-
creased dramatically. This method involves filtering the sample onto
0.02-m-pore-size ceramic filters, staining viruses on the filters by
using one of several available nucleic acid dyes, mounting the filter
onto a slide, and visually enumerating the deposited viruses by epi-
fluorescence microscopy (reviewed in reference 4). However, the fil-
ters used for this “filter mount” method have risen in cost to ca. $10
each in the United States (with increased costs in some other coun-
tries), creating a significant financial burden for researchers pursuing
studies of environmental viruses. To address this, we have developed
a new, less costly “wet-mount” epifluorescence microscopy method
to enumerate aquatic viruses, in which fluorescently stained samples
are wet mounted directly onto a slide, with quantification of viral
concentrations based on the relative abundance of viruses and silica
beads in the sample.
MATERIALS AND METHODS
Comparison of the wet-mount and filter mount methods for virus enu-
meration. The wet-mount method was tested by comparing viral concen-
trations obtained with the wet-mount and filter mount methods in trip-
licate samples collected from a variety of marine and freshwater
environments as well as in cultivated viral lysates (described in Table S1 in
the supplemental material). Briefly, field samples included those from a
6-depth profile (5 to 300 m) from the Eastern Tropical North Pacific
Ocean (using whole seawater samples); 8 surface ocean locations through-
out the Pacific, Atlantic, and Southern Oceans chosen for their range of
chlorophyll concentrations (collected on the Tara Oceans Expedition [6],
using 0.2-m-filtered samples); and a freshwater location in South Caro-
lina. All field samples were preserved with glutaraldehyde (0.5% final
concentration), flash frozen in liquid nitrogen, and stored at 80°C until
analysis. Lysate samples included the Synechococcus virus S-WHM1 (7),
two dilutions of the Synechococcus virus S-SM1 (8), and the Prochlorococ-
cus virus P-HM2 (9). Triplicate independent 1-ml samples were processed
by using each of the filter mount and wet-mount methods, as described
below. Statistical comparison of viral concentrations obtained by using
each method was then performed by using two-tailed ttests and Pearson
correlation (SigmaPlot v12.5; Systat Software Inc.).
Filter mount sample preparation and analysis. The filter mount
method was performed according to methods described previously by
Received 4 November 2014 Accepted 12 February 2015
Accepted manuscript posted online 20 February 2015
Citation Cunningham BR, Brum JR, Schwenck SM, Sullivan MB, John SG. 2015. An
inexpensive, accurate, and precise wet-mount method for enumerating aquatic
viruses. Appl Environ Microbiol 81:2995–3000. doi:10.1128/AEM.03642-14.
Editor: K. E. Wommack
Address correspondence to Seth G. John, sjohn@geol.sc.edu.
B.R.C. and J.R.B. contributed equally to this work.
This article is contribution number 0015 of the Tara Oceans Expedition
2009 –2012.
Supplemental material for this article may be found at http://dx.doi.org/10.1128
/AEM.03642-14.
Copyright © 2015, American Society for Microbiology. All Rights Reserved.
doi:10.1128/AEM.03642-14
May 2015 Volume 81 Number 9 aem.asm.org 2995Applied and Environmental Microbiology
Suttle and Fuhrman (4). Briefly, samples were filtered onto 0.02-m-
pore-size ceramic filters (Whatman Anodisc), stained with SYBR gold
(Invitrogen) for 15 min, and mounted onto a glass slide with an anti-
fade solution (Acros Organics). Viruses were viewed under blue exci-
tation using a Nikon TS100 inversion microscope or a Zeiss Axio Im-
ager epifluorescence microscope at a 1,000 magnification. The viral
concentration was determined by using the average number of fluo-
rescent viruses within a given area of the microscope reticle in 20 fields
of view and the total volume of sample filtered through a measured
area on the filter.
Wet-mount sample preparation and analysis. The wet-mount
method for enumerating viruses involves an optional virus concentration
step followed by combining a known volume of stained sample with a
known volume and concentration of silica beads for relative enumeration
of viruses and beads to calculate the virus concentration in the sample
(Fig. 1). The reagents for assessing viral concentrations using the wet-
mount method are fully described in Table 1, and the protocol is as fol-
lows.
1. If the virus concentration is expected to be 510
7
viruses ml
1
,
concentrate the viruses by chemical flocculation, as follows. Add 1
l iron chloride solution to a 1-ml sample in a microcentrifuge
tube, mix the sample by inversion, and centrifuge the mixture for
20 min at 14,000 g. Remove the supernatant and resuspend the
pellet in 10 l ascorbate-EDTA buffer. (If a lower concentration
factor is desired, the pellet may be resuspended in a larger volume
of ascorbate-EDTA buffer. In this case, increase the amounts of
SYBR gold, glycerol, and silica beads accordingly in subsequent
steps.)
2. Add 2 l SYBR gold working stock to the concentrated sample,
vortex the mixture, and incubate the mixture for 15 min in the
dark. If the sample was not concentrated with chemical floccula-
tion (i.e., the concentration is expected to exceed 5 10
7
viruses
ml
1
), mix 10 l unconcentrated sample with 2 l SYBR gold
working stock in a microcentrifuge tube.
3. Add 5 l glycerol, vortex the mixture, and add 2 l working bead
solution (thoroughly vortex the working bead solution before ad-
dition to the sample to ensure accurate pipetting of the beads). If
the sample was not concentrated with chemical flocculation, add 1
l ascorbic acid antifade solution as well.
4. Mix the prepared sample thoroughly by pipetting up and down and
then immediately pipette 10 l onto a glass microscope slide and
place a coverslip over the sample (both the glass slide and coverslip
should be cleaned with isopropanol). Avoid trapping air under the
coverslip.
5. Using an epifluorescence microscope, count the number of viruses
in a given area within the microscope reticle under blue (495-
nm) excitation at a 1,000 magnification. Within the same field of
view, count the beads under white light. Continue counting fields
of view until at least 100 each of viruses and beads have been
counted.
6. The virus concentration is calculated as c
virus
n
virus
/n
beads
v
beads
/v
sample
c
beads
, where c
virus
is the virus concentration in the
sample (viruses ml
1
), n
virus
is the total number of viruses counted,
n
beads
is the total number of beads counted, v
beads
is the volume of
working bead solution added (l), v
sample
is the sample volume
FIG 1 Overview of the wet-mount method for enumeration of aquatic viruses.
TABLE 1 Reagent preparation for the wet-mount virus enumeration protocol
Reagent Prepn method
SYBR gold working stock Dilute SYBR gold (Invitrogen) (10,000stock) into PBS to prepare a 1,000solution
Ascorbic acid antifade solution Dissolve ascorbic acid into PBS to create a 10% (wt/vol) solution
a
Working bead solution Thoroughly vortex the stock bead solution (2.34-m silica spheres) (catalog no.
SS04N/4186; Bangs Laboratories), dilute it 10-fold into PBS to obtain a concn of
10
8
beads ml
1
, and store it at 4°C
Iron chloride solution Dissolve FeCl
3
·6H
2
O into ultrapure water to form a solution of 10 g Fe liter
1
; the
solution has expired if a cloudy precipitate forms
b
Ascorbate-EDTA buffer Combine equal parts of 0.4 M Mg
2
EDTA and 0.8 M ascorbic acid and adjust with 10
N NaOH to reach a pH of 6–7; prepare fresh within 48 h of use
c
a
See reference 11.
b
See reference 10.
c
An alternative ascorbate-EDTA buffer can be made with MgCl
2
and Na
2
EDTA if Mg
2
EDTA is unavailable (10).
Cunningham et al.
2996 aem.asm.org May 2015 Volume 81 Number 9Applied and Environmental Microbiology
(l) (the volume prior to concentration if chemical flocculation is
used), and c
beads
is the bead concentration in the working bead
solution (beads ml
1
).
A full, illustrated protocol describing this method is also available
online for convenience (http://eebweb.arizona.edu/faculty/mbsulli
/protocols/). For analysis of samples with 510
7
viruses ml
1
, viruses
must be concentrated with chemical flocculation by using a method
adapted from methods described previously by John et al. (10), to obtain
a sufficient concentration of viruses for analysis (Fig. 1A). Samples in this
study with viral concentrations below that threshold were first concen-
trated 100-fold with this chemical flocculation method (step 1 in the list of
procedures above) before being stained with SYBR gold for 15 min and
then combined with silica beads and glycerol (added to create a more
viscous solution and reduce clumping of beads) (steps 2 and 3 in the list of
procedures above) (Fig. 1B). The silica bead size (2.34 m) was selected
due to the ease of visually counting the beads under white light. Due to the
relatively large size of the beads (compared to the size of viruses), the bead
solution must be vortexed thoroughly prior to the addition of beads to the
sample to ensure the addition of an accurate concentration of beads.
These concentrated samples did not require the addition of an antifade
solution since they were resuspended in a buffer containing ascorbic acid,
which reduces fading of the fluorescent signal (11). Samples were then
pipetted directly onto an isopropanol-cleaned glass microscope slide and
covered with a cleaned glass coverslip (step 4 in the list of procedures
above). Viruses and beads were enumerated in multiple fields of view on a
Nikon TS100 inversion microscope or a Zeiss Axio Imager epifluores-
cence microscope at a 1,000 magnification until at least 100 each of
viruses and beads were enumerated to calculate the virus concentration
(steps 5 and 6 in the list of procedures above). For each field of view, the
total number of fluorescent viruses was determined under blue excitation,
after which the total number of beads within the same field of view was
determined under white light (Fig. 2).
For analysis of samples with 510
7
viruses ml
1
, chemical floccu-
lation of viruses prior to wet-mount sample preparation was not necessary
to obtain a sufficient concentration of viruses for enumeration. These
samples (S-SM1 lysates) were prepared by staining 10 l of sample with
SYBR gold, followed by the addition of an ascorbic acid solution (to act as
an antifade solution), glycerol, and silica beads (steps 2 and 3 in the list of
procedures above) (Fig. 1B). These samples were then wet mounted onto
slides and enumerated exactly as described above for the samples that had
been concentrated with chemical flocculation. To compare the wet-
mount method with the filter mount method for these samples with high
viral concentrations, they were diluted 10-fold in phosphate-buffered sa-
line (PBS) prior to filtering 1 ml of sample for the filter mount method, as
the undiluted sample would have resulted in an excessive viral density on
the filter, preventing analysis. We also note that while p-phenylenedi-
FIG 2 Images of samples prepared by use of the filter mount and wet-mount virus enumeration methods. Shown are epifluorescence images of purified S-SM1
lysate obtained by using the filter mount (A) and wet-mount (B) methods, seawater from 30 m in the Pacific Ocean depth profile by using the filter mount (D)
and wet-mount (E) methods, freshwater from Lake Murray by using the filter mount (F) and wet-mount (G) methods, and unpurified S-SM1 lysate with
Synechococcus cells by using the filter mount (H) and wet-mount (I) methods. These epifluorescence images include arrows pointing to two of the viruses in each
image. Under white light (C), beads are visible in the same field of view as for the wet-mount sample (B), with arrows pointing to two of the beads in the image.
Bar, 10 m.
A Wet-Mount Method for Enumerating Aquatic Viruses
May 2015 Volume 81 Number 9 aem.asm.org 2997Applied and Environmental Microbiology
amine is a popular antifade chemical (4), it reacted with glutaraldehyde to
form a precipitate in these wet-mount samples and thus should not be
used in the wet-mount method with fixed samples.
The minimum number of beads and viruses enumerated per sample is
justified as follows. Counting statistics (also known as shot noise) dictates
that the error in the quantity of viruses or beads counted is given by 1⁄n
(12), where nis the number of objects enumerated, and therefore, the total
error in viral abundance is total 1⁄nvirus1⁄nbeads, where n
virus
and
n
beads
are the total numbers of viruses and beads counted, respectively.
When at least 100 each of viruses and beads are enumerated, the maxi-
mum error is 14%.
Storage conditions for samples prepared by using the wet-mount
method. Storage conditions were assessed by using two different lysates
concentrated 100using the flocculation method described above. To
assess storage after samples were mounted onto slides, triplicate samples
(S-SM1 lysate) were prepared and analyzed by using the full protocol
described above, with slides being stored vertically at 20°C immediately
after enumeration, and viruses and beads were recounted 7 days later.
Additional triplicate samples (S-WHM1 lysates) were prepared through
step 3 in the list of procedures above, with 10 l of the prepared sample
being analyzed immediately and the remaining sample (10 l) being
stored in a microcentrifuge tube at 20°C until analysis 7 days later.
RESULTS AND DISCUSSION
The wet-mount method resulted in fluorescently stained viruses
with an intensity similar to those of the filter-mount method (Fig.
2). While there was typically a lower density of viruses in the im-
ages derived from samples prepared by using the wet-mount
method, this is favorable because viruses are enumerated in larger
fields of view with the wet-mount than with the filter mount
method. However, images depicting a greater density of viruses
and cells can be obtained with more concentrated samples (Fig.
2I). Viral concentrations obtained by using the wet-mount
method were strongly correlated (Pearson correlation coefficient
of 0.986; P0.001) with those obtained by using the filter mount
method for all sample types tested, including viral lysates and
samples from a variety of oceanic and freshwater regions (Fig. 3).
There was no significant difference in viral concentrations ob-
tained by the use of these methods for the majority of samples (13
of 19 samples; two-tailed ttests) (see Table S1 in the supplemental
material). For the remaining samples with significantly different
viral concentrations, neither method consistently resulted in
higher or lower viral concentrations, nor were these differences
restricted to a specific range of viral concentrations (i.e., high ver-
sus low) or sample type (i.e., freshwater versus marine sample,
natural sample versus lysate, or low versus high chlorophyll con-
centration), indicating stochastic variability inherent to analyses
of samples (Fig. 3; see also Table S1 in the supplemental material).
Furthermore, we consider the low magnitude of the differences in
FIG 3 Viral concentrations in natural samples and lysates obtained by using
the filter mount and wet-mount enumeration methods. Error bars are stan-
dard deviations of the means of data from triplicate samples. Closed symbols
represent samples in which there was no significant difference in virus concen-
trations obtained by using the filter mount and wet-mount methods (P0.05
by two-tailed ttests), while open symbols represent samples in which there was
a significant difference (see Table S1 in the supplemental material). Average
viral concentrations for all samples obtained by using each method were
strongly and positively correlated (Pearson correlation coefficient, 0.986; P
0.001). The solid lines represent a 1:1 relationship, and dashed lines represent
an interval of 70% agreement between methods around the 1:1 relationship to
facilitate visual comparison of results.
Cunningham et al.
2998 aem.asm.org May 2015 Volume 81 Number 9Applied and Environmental Microbiology
average viral concentrations for the few significantly different
samples to be acceptable for studies of aquatic viruses.
It is important to note that there is no available standard used
in aquatic virus enumeration methods. Previous studies compar-
ing aquatic viral concentrations determined by using different
methods (i.e., TEM, filter mount, and flow cytometry) have
shown discrepancies between methods, with one method usually
resulting in consistently higher viral concentrations (13–17).
However, we observed no such consistent differences in our com-
parison of the wet-mount and filter mount methods. Further-
more, the comparison in this study showed that most of the sam-
ples had at least 70% agreement between virus concentrations
obtained by use of the wet-mount method and those obtained by
use of the filter mount method (Fig. 3), which is similar to data
from previously reported comparisons of methods used to enu-
merate viruses (16,17). The wet-mount method also had high
precision; standard deviations of the means for triplicate samples
were 2 to 18% (average, 7% 4%) of the mean virus concentra-
tion and were not significantly different from those obtained by
using the filter mount method (P0.531 by two-tailed ttest).
Thus, the wet-mount method and the filter mount method can be
used with equal confidence.
When enumerating viruses, it is sometimes advantageous to
store prepared samples for enumeration at a later date. For exam-
ple, samples prepared with the filter mount method can be stored
at 20°C for at least 4 months, with no significant change in viral
concentrations (4). For wet-mount samples, we tested storage of
prepared samples both in microcentrifuge tubes and on slides at
20°C (Fig. 4). While the calculated virus concentration was
higher after storage of the prepared samples under both storage
conditions, these differences were not significant (P0.210 and
P0.083, respectively, by two-tailed ttests). Thus, samples pre-
pared by use of the wet-mount method can be stored frozen either
before or after mounting the sample onto slides, with no signifi-
cant change in the calculated virus concentration.
The wet-mount method had one major drawback compared to
the filter mount method, which was the inability to efficiently
enumerate samples with viral concentrations of 110
6
viruses
ml
1
. Attempted analysis of samples with lower viral concentra-
tions (i.e., samples below 300 m in the Pacific Ocean depth profile)
using the wet-mount method resulted in 1 virus per field of
view, even after maximum concentration (100-fold) with chemi-
cal flocculation. Thus, the wet-mount method is not recom-
mended for samples with viral concentrations of 110
6
viruses
ml
1
because the low density of viruses on the slide significantly
extends the time for analysis of a sample. Although this limitation
prevented analysis of the deep-sea samples (300 m) in the Pacific
Ocean depth profile in this study, many deep-sea samples have
viral concentrations above this limit (e.g., see reference 18), and
thus, the wet-mount method should be useful for a wide range of
environmental samples.
The available methods to enumerate aquatic viruses each have
benefits and limitations that are worth considering when planning
research projects. For example, TEM-based analyses of aquatic
samples can generate information about the morphological char-
acteristics of viruses (e.g., see reference 21) in addition to viral
abundance (e.g., see reference 3) but can potentially underesti-
mate the number of viruses because they may be obscured by
debris in the sample (16). Fluorescence-based methods for viral
enumeration (i.e., epifluorescence microscopy and flow cytom-
etry) are significantly faster than TEM but can potentially falsely
include gene transfer agents or cell debris as viruses (reviewed in
reference 19) while excluding single-stranded DNA (ssDNA) vi-
ruses that have very faint fluorescence (20). One additional advan-
tage of fluorescence-based methods is the ability to enumerate
both viruses and bacteria (if present) by using the same prepared
sample (e.g., see reference 16). However, the wet-mount method
presented here has not yet been evaluated for accuracy in counting
of bacterial cells. Among the available epifluorescence-based
methods, the filter mount method also provides an opportunity to
obtain images with a high density of viruses and cells, while the
flow cytometry method does not. The viral density in images ob-
tained by use of the wet-mount method is generally much lower
than that for filter mount samples, although the density of viruses
and cells increases when more concentrated samples are used.
While each of these variables is important when evaluating poten-
tial virus enumeration methods for a given project, we offer the
wet-mount method as a cost-effective alternative to the widely
used filter mount epifluorescence method.
A significant advantage of the wet-mount method over the
filter mount method is the lack of a requirement for costly 0.02-
m-pore-size ceramic filters. Currently, these filters are available
from only one supplier and are expensive ($10 each). Instead,
the wet-mount method uses microsphere silica beads that can be
purchased from several suppliers at a 500-fold-lower cost ($0.02
for 20 lofa10
8
-bead ml
1
working solution per sample, calcu-
lated based on $150 for 15 ml of a 10
9
-bead ml
1
stock solution).
Even after accounting for the cost of other reagents and slides, the
per-sample materials cost for the wet-mount method is much
lower ($0.10 per sample). Thus, the wet-mount method is rec-
ommended as an equivalently accurate and precise but cheaper
alternative for enumerating viruses in field and laboratory sam-
ples with viral concentrations of 110
6
viruses ml
1
.
Conclusion. Enumeration of viruses in field and laboratory
samples is an important tool for investigating the numerous in-
FIG 4 Storage of samples prepared by using the wet-mount method. Concen-
trations of viruses in triplicate samples (S-SM1 lysate for tube storage and
S-WHM1 lysate for slide storage) prepared according to the wet-mount pro-
tocol and stored at 20°C in microcentrifuge tubes (tube storage) or wet
mounted onto slides (slide storage) are shown. Viruses were enumerated im-
mediately after preparation (time [T]0 days) and after 7 days of storage
(T7 days). Error bars are standard deviations of the means for triplicate
samples.
A Wet-Mount Method for Enumerating Aquatic Viruses
May 2015 Volume 81 Number 9 aem.asm.org 2999Applied and Environmental Microbiology
fluences of viruses in aquatic environmental systems. However,
the high cost of enumerating viruses in aquatic samples using the
established filter mount epifluorescence microscopy method can
be a significant burden in conducting aquatic virus research. In
this study, we present a new, less expensive wet-mount method for
aquatic virus enumeration that can be used with accuracy and
precision equivalent to those of the filter mount method for a
variety of environmental and laboratory samples.
ACKNOWLEDGMENTS
We thank Bonnie Poulos from the Tucson Marine Phage Lab at the Uni-
versity of Arizona for her help troubleshooting initial problems with this
method. We also thank the crew and scientists of the R/V New Horizon for
their assistance when sampling in the Eastern Tropical North Pacific
Ocean, as well as the coordinators and members of the Tara Oceans con-
sortium for organizing sampling.
This publication was funded in part by Gordon and Betty Moore
Foundation grant GBMF3305 to S.G.J. and M.B.S. and by grants
GBMF2631 and GBMF3790 to M.B.S. We also acknowledge the following
sponsors for their support in the Tara Oceans Expedition: CNRS, EMBL,
Genoscope/CEA, VIB, Stazione Zoologica Anton Dohrn, UNIMIB,
ANR (projects POSEIDON/ANR-09-BLAN-0348, BIOMARKS/ANR-08-
BDVA-003, PROMETHEUS/ANR-09-GENM-031, and TARA-GIRUS/
ANR-09-PCS-GENM-218), EU FP7 (MicroB3/no. 287589), FWO, BIO5,
Biosphere 2, agnès b., the Veolia Environment Foundation, Region
Bretagne, World Courier, Illumina, Cap L’Orient, the EDF Foundation
EDF Diversiterre, FRB, the Prince Albert II de Monaco Foundation, Eti-
enne Bourgois, and the captain and crew of the Tara schooner. Tara
Oceans would not exist without continuous support from 23 institutes.
REFERENCES
1. Suttle CA. 2005. Viruses in the sea. Nature 437:356 –361. http://dx.doi.org
/10.1038/nature04160.
2. Breitbart M. 2012. Marine viruses: truth or dare. Annu Rev Mar Sci
4:425–448. http://dx.doi.org/10.1146/annurev-marine-120709-142805.
3. Bergh O, Borsheim K, Bratbak G, Heldal M. 1989. High abundance of
viruses found in aquatic environments. Nature 340:467–468. http://dx
.doi.org/10.1038/340467a0.
4. Suttle CA, Fuhrman J. 2010. Enumeration of virus particles in aquatic or
sediment samples by epifluorescence microscopy, p 145–153. In Wilhelm
SW, Weinbauer MG, Suttle CA (ed), Manual of aquatic viral ecology.
ASLO, Waco, TX.
5. Brussaard C, Payet J, Winter C, Weinbauer M. 2010. Quantification of
aquatic viruses by flow cytometry, p 102–109. In Wilhelm SW, Weinbauer
MG, Suttle CA (ed), Manual of aquatic viral ecology. ASLO, Waco, TX.
6. Karsenti E, Acinas SG, Bork P, Bowler C, De Vargas C, Raes J, Sullivan
MB, Arendt D, Benzoni F, Claverie J-M, Follows M, Gorsky G,
Hingamp P, Iudicone D, Jaillon O, Kandels-Lewis S, Krzic U, Not F,
Ogata H, Pesant S, Reynaud EG, Sardet C, Sieracki ME, Speich S,
Velayoudon D, Weissenbach J, Wincker P. 2011. A holistic approach to
marine eco-systems biology. PLoS Biol 9:e1001177. http://dx.doi.org/10
.1371/journal.pbio.1001177.
7. Millard A, Clokie MRJ, Shub DA, Mann NH. 2004. Genetic organization
of the psbAD region in phages infecting marine Synechococcus strains. Proc
Natl Acad SciUSA101:11007–11012. http://dx.doi.org/10.1073/pnas
.0401478101.
8. Sullivan MB, Waterbury JB, Chisholm SW. 2003. Cyanophages infecting
the oceanic cyanobacterium Prochlorococcus. Nature 424:1047–1051. http:
//dx.doi.org/10.1038/nature01929.
9. Sullivan MB, Huang KH, Ignacio-Espinoza JC, Berlin AM, Kelly L,
Weigele PR, DeFrancesco AS, Kern SE, Thompson LR, Young S, Yan-
dava C, Fu R, Krastins B, Chase M, Sarracino D, Osburne MS, Henn
MR, Chisholm SW. 2010. Genomic analysis of oceanic cyanobacterial
myoviruses compared with T4-like myoviruses from diverse hosts and
environments. Environ Microbiol 12:3035–3056. http://dx.doi.org/10
.1111/j.1462-2920.2010.02280.x.
10. John SG, Mendez CB, Deng L, Poulos B, Kauffman AKM, Kern S, Brum
J, Polz MF, Boyle EA, Sullivan MB. 2011. A simple and efficient method for
concentration of ocean viruses by chemical flocculation. Environ Microbiol
Rep 3:195–202. http://dx.doi.org/10.1111/j.1758-2229.2010.00208.x.
11. Patel A, Noble RT, Steele JA, Schwalbach MS, Hewson I, Fuhrman JA.
2007. Virus and prokaryote enumeration from planktonic aquatic envi-
ronments by epifluorescence microscopy with SYBR green I. Nat Protoc
2:269–276. http://dx.doi.org/10.1038/nprot.2007.6.
12. John SG, Adkins JF. 2010. Analysis of dissolved iron isotopes in seawater.
Mar Chem 119:65–76. http://dx.doi.org/10.1016/j.marchem.2010.01.001.
13. Bettarel Y, Sime-Ngando T, Amblard C, Laveran H. 2000. A comparison
of methods for counting viruses in aquatic systems. Appl Environ Micro-
biol 66:2283–2289. http://dx.doi.org/10.1128/AEM.66.6.2283-2289.2000.
14. Hennes KP, Suttle CA. 1995. Direct counts of viruses in natural waters
and laboratory cultures by epifluorescence microscopy. Limnol Oceanogr
40:1050–1055. http://dx.doi.org/10.4319/lo.1995.40.6.1050.
15. Weinbauer MG, Suttle CA. 1997. Comparison of epifluorescence and
transmission electron microscopy for counting viruses in natural marine
waters. Aquat Microb Ecol 13:225–232. http://dx.doi.org/10.3354/ame
013225.
16. Noble R, Fuhrman J. 1998. Use of SYBR green I for rapid epifluorescence
counts of marine viruses and bacteria. Aquat Microb Ecol 14:113–118.
http://dx.doi.org/10.3354/ame014113.
17. Marie D, Brussaard C, Thyrhaug R, Bratbak G, Vaulot D. 1999. Enu-
meration of marine viruses in culture and natural samples by flow cytom-
etry. Appl Environ Microbiol 65:45–52.
18. Parada V, Sintes E, van Aken HM, Weinbauer MG, Herndl GJ. 2007.
Viral abundance, decay, and diversity in the meso- and bathypelagic wa-
ters of the North Atlantic. Appl Environ Microbiol 73:44294438. http:
//dx.doi.org/10.1128/AEM.00029-07.
19. Wommack KE, Colwell RR. 2000. Virioplankton: viruses in aquatic eco-
systems. Microbiol Mol Biol Rev 64:69–114. http://dx.doi.org/10.1128
/MMBR.64.1.69-114.2000.
20. Holmfeldt K, Odic´ D, Sullivan MB, Middelboe M, Riemann L. 2012.
Cultivated single-stranded DNA phages that infect marine Bacteroidetes
prove difficult to detect with DNA-binding stains. Appl Environ Micro-
biol 78:892–894. http://dx.doi.org/10.1128/AEM.06580-11.
21. Brum JR, Schenck RO, Sullivan MB. 2013. Global morphological anal-
ysis of marine viruses shows minimal regional variation and dominance of
non-tailed viruses. ISME J 7:1738–1751. http://dx.doi.org/10.1038/ismej
.2013.67.
Cunningham et al.
3000 aem.asm.org May 2015 Volume 81 Number 9Applied and Environmental Microbiology
... Wet-mount EFM has been suggested as a cost-effective alternative to Whatman Anodisc filter membrane-based protocols (31). This EFM method relies on chemical flocculation concentration of the VLPs using iron chloride precipitation, followed by EDTA-ascorbate resuspension, SYBR Gold nucleic acid staining, and wet mounting on a standard glass slide (31). ...
... Wet-mount EFM has been suggested as a cost-effective alternative to Whatman Anodisc filter membrane-based protocols (31). This EFM method relies on chemical flocculation concentration of the VLPs using iron chloride precipitation, followed by EDTA-ascorbate resuspension, SYBR Gold nucleic acid staining, and wet mounting on a standard glass slide (31). With and without chemical flocculation-based wet mount, EFM showed similar concordance and precision with Anodisc-based methods at a much lower cost (31). ...
... This EFM method relies on chemical flocculation concentration of the VLPs using iron chloride precipitation, followed by EDTA-ascorbate resuspension, SYBR Gold nucleic acid staining, and wet mounting on a standard glass slide (31). With and without chemical flocculation-based wet mount, EFM showed similar concordance and precision with Anodisc-based methods at a much lower cost (31). It also was effective for natural ambient marine and freshwater ecosystems from the low end of 1 × 10 6 mL −1 to a high end of 1 × 10 8 mL −1 , similar to Anodisc methods (31). ...
Article
Full-text available
Enumeration is a fundamental measure of community ecology in which viruses represent the most numerous biological identities. Epifluorescence microscopy (EFM) has been the gold standard method for environmental viral enumeration for over 25 years. Currently, standard EFM methods using the Anodisc filters are no longer cost-effective (>$15 per slide) and have yet to be applied to modern microbialites. Microbialites are microbially driven benthic organosedimentary deposits that have been present for most of Earth’s history. We present a cost-effective method for environmental viral enumeration from aquatic samples, microbial mats, and exopolymeric substances (EPSs) within modern microbialites using EFM. Our integrated approach, which includes filtration, differential centrifugation, chloroform treatment, glutaraldehyde fixation, benzonase nuclease treatment, probe sonication (EPS and mat only), SYBR Gold staining, wet mounting, and imaging, provides a robust method for modern microbialites and aquatic samples. Viral abundances of modern microbialites and aquatic samples collected from Fayetteville Green Lake (FGL) and Great Salt Lake (GSL) did not differ across ecosystems by sample type. EPS and microbial mat samples had an order of magnitude higher viral-like particle abundance when compared to water regardless of the ecosystem (10 ⁷ vs 10 ⁶ ). Viral enumeration allows for estimates of total viral numbers and weights. The entire weight of all the viruses in FGL and GSL are ~598 g and ~2.2 kg, respectively. Further development of EFM methods and software is needed for viral enumeration. Our method provides a robust and cost-effective (~$0.75 per sample) viral enumeration within modern microbialites and aquatic ecosystems. IMPORTANCE Low-cost and robust viral enumeration is a critical first step toward understanding the global virome. Our method is a deep drive integration providing a window into viral dark matter within aquatic ecosystems. We enumerated the viruses within Green Lake and Great Salt Lake microbialites, EPS, and water column. The entire weight of all the viruses in Green Lake and Great Salt Lake are ~598 g and ~2.2 kg, respectively.
... Viruses were concentrated from 0.22-μm-pore-sized filtrate, which excluded intracellular viruses including temperate viruses [61], and then treated with DNase to remove free DNA. Counts of VLPs in the two samples were below the detection limit using a wet-mount method (<10 6 VLPs ml −1 [62];). Thus, we applied the low-input quantitative viral metagenomic sequencing that was previously established to study seawater viral communities [46,47,63,64], to the viral concentrates in our low-biomass glacier ice samples. ...
... Pseudoalteromonas phages strain PSA-HP1 (NCBI: txid134839) were harvested from 95% lysed plaque assays (agar overlay technique). The concentration of PSA-HP1 was counted by a wet-mount method using SYBR Gold (Cat No. S11494, Life Technologies) staining and glass beads as described previously [62]. The lambda phage DNA (100 μg/ml; 1.88×10 9 copies/μl; genome size 4.8 kb) was purchased from Life Technologies (Cat. ...
... No. A7181 and A7211, respectively; Promega, USA) [45]. Viral abundance, calculated prior to DNA extraction, was obtained by enumerating and comparing the counts of VLPs and beads (with a known concentration) using the wetmount method [62]. ...
Article
Full-text available
Background Glacier ice archives information, including microbiology, that helps reveal paleoclimate histories and predict future climate change. Though glacier-ice microbes are studied using culture or amplicon approaches, more challenging metagenomic approaches, which provide access to functional, genome-resolved information and viruses, are under-utilized, partly due to low biomass and potential contamination. Results We expand existing clean sampling procedures using controlled artificial ice-core experiments and adapted previously established low-biomass metagenomic approaches to study glacier-ice viruses. Controlled sampling experiments drastically reduced mock contaminants including bacteria, viruses, and free DNA to background levels. Amplicon sequencing from eight depths of two Tibetan Plateau ice cores revealed common glacier-ice lineages including Janthinobacterium, Polaromonas, Herminiimonas, Flavobacterium, Sphingomonas, and Methylobacterium as the dominant genera, while microbial communities were significantly different between two ice cores, associating with different climate conditions during deposition. Separately, ~355- and ~14,400-year-old ice were subject to viral enrichment and low-input quantitative sequencing, yielding genomic sequences for 33 vOTUs. These were virtually all unique to this study, representing 28 novel genera and not a single species shared with 225 environmentally diverse viromes. Further, 42.4% of the vOTUs were identifiable temperate, which is significantly higher than that in gut, soil, and marine viromes, and indicates that temperate phages are possibly favored in glacier-ice environments before being frozen. In silico host predictions linked 18 vOTUs to co-occurring abundant bacteria (Methylobacterium, Sphingomonas, and Janthinobacterium), indicating that these phages infected ice-abundant bacterial groups before being archived. Functional genome annotation revealed four virus-encoded auxiliary metabolic genes, particularly two motility genes suggest viruses potentially facilitate nutrient acquisition for their hosts. Finally, given their possible importance to methane cycling in ice, we focused on Methylobacterium viruses by contextualizing our ice-observed viruses against 123 viromes and prophages extracted from 131 Methylobacterium genomes, revealing that the archived viruses might originate from soil or plants. Conclusions Together, these efforts further microbial and viral sampling procedures for glacier ice and provide a first window into viral communities and functions in ancient glacier environments. Such methods and datasets can potentially enable researchers to contextualize new discoveries and begin to incorporate glacier-ice microbes and their viruses relative to past and present climate change in geographically diverse regions globally. 8eoz6b1Gq7e8M2dM_SBXdyVideo Abstract
... This study focused on extracellular viruses in the filtrate. The virus-like particles were counted using the wet-mount method [68]. For the samples SB and SW, viruses were concentrated using an iron chloride flocculation method [69] and stored at 4°C at the BARC. ...
Article
Full-text available
Background Climate change threatens Earth’s ice-based ecosystems which currently offer archives and eco-evolutionary experiments in the extreme. Arctic cryopeg brine (marine-derived, within permafrost) and sea ice brine, similar in subzero temperature and high salinity but different in temporal stability, are inhabited by microbes adapted to these extreme conditions. However, little is known about their viruses (community composition, diversity, interaction with hosts, or evolution) or how they might respond to geologically stable cryopeg versus fluctuating sea ice conditions. Results We used long- and short-read viromics and metatranscriptomics to study viruses in Arctic cryopeg brine, sea ice brine, and underlying seawater, recovering 11,088 vOTUs (~species-level taxonomic unit), a 4.4-fold increase of known viruses in these brines. More specifically, the long-read-powered viromes doubled the number of longer (≥25 kb) vOTUs generated and recovered more hypervariable regions by >5-fold compared to short-read viromes. Distribution assessment, by comparing to known viruses in public databases, supported that cryopeg brine viruses were of marine origin yet distinct from either sea ice brine or seawater viruses, while 94% of sea ice brine viruses were also present in seawater. A virus-encoded, ecologically important exopolysaccharide biosynthesis gene was identified, and many viruses (~half of metatranscriptome-inferred “active” vOTUs) were predicted as actively infecting the dominant microbial genera Marinobacter and Polaribacter in cryopeg and sea ice brines, respectively. Evolutionarily, microdiversity (intra-species genetic variations) analyses suggested that viruses within the stable cryopeg brine were under significantly lower evolutionary pressures than those in the fluctuating sea ice environment, while many sea ice brine virus-tail genes were under positive selection, indicating virus-host co-evolutionary arms races. Conclusions Our results confirmed the benefits of long-read-powered viromics in understanding the environmental virosphere through significantly improved genomic recovery, expanding viral discovery and the potential for biological inference. Evidence of viruses actively infecting the dominant microbes in subzero brines and modulating host metabolism underscored the potential impact of viruses on these remote and underexplored extreme ecosystems. Microdiversity results shed light on different strategies viruses use to evolve and adapt when extreme conditions are stable versus fluctuating. Together, these findings verify the value of long-read-powered viromics and provide foundational data on viral evolution and virus-microbe interactions in Earth’s destabilized and rapidly disappearing cryosphere.
... This study focused on extracellular viruses in the ltrate. The virus-like particles were counted using the wet-mount method [70]. For the samples SB and SW, viruses were concentrated using an iron chloride occulation method [71] and stored at 4°C at the BARC. ...
Preprint
Full-text available
Background: Climate change threatens Earth’s ice-based ecosystems which currently offer archives and eco-evolutionary experiments in the extreme. Arctic cryopeg brine (marine-derived, within permafrost) and sea-ice brine, similar in subzero temperature and high salinity but different in temporal stability, are inhabited by microbes adapted to these extreme conditions. However, little is known about their viruses (community composition, diversity, interaction with hosts, or evolution) or how they might respond to geologically stable cryopeg versus fluctuating sea-ice conditions. Results: We used long- and short-read viromics and metatranscriptomics to study viruses in Arctic cryopeg brine, sea-ice brine, and underlying seawater, recovering 11,088 vOTUs (~species-level taxonomic unit), a 4.4-fold increase of known viruses in these brines. More specifically, the long-read-powered viromes doubled the number of longer (≥25 kb) vOTUs generated and recovered more hypervariable regions by >5-fold compared to short-read viromes. Distribution assessment, by comparing to known viruses in public databases, supported that cryopeg-brine viruses were of marine origin yet distinct from either sea-ice-brine or seawater viruses, while 94% of sea-ice-brine viruses also presented in seawater. A virus-encoded, ecologically important exopolysaccharide biosynthesis gene was identified, and many viruses (~half of metatranscriptome-inferred ‘active’ vOTUs) were predicted as actively infecting the dominant microbial genera Marinobacter and Polaribacterin cryopeg and sea-ice brines, respectively. Evolutionarily, microdiversity (intra-species genetic variations) analyses suggested that viruses within the stable cryopeg brine were under significantly lower evolutionary pressures than those in the fluctuating sea-ice environment, while many sea-ice-brine virus-tail genes were under positive selection, indicating virus-host co-evolutionary arms races. Conclusions: Our results confirmed the benefits of long-read-powered viromics in understanding the environmental virosphere through significantly improved genomic recovery, expanding viral discovery and the potential for biological inference. Evidence of viruses actively infecting the dominant microbes in subzero brines and modulating host metabolism underscored the potential impact of viruses on these remote and underexplored extreme ecosystems. Microdiversity results shed light on the different strategies viruses use to evolve and adapt when extreme conditions are stable versus fluctuating. Together, these findings verify the value of long-read-powered viromics and provide foundational data on viral evolution and virus-microbe interactions in Earth’s destabilized and rapidly disappearing cryosphere.
... Bacterial cells were then mixed separately with fluorescently-labeled viruses HM1 and HS8 to final MOIs = 1, 2, and 4 and 2, 5, and 10, respectively (six samples in total). After incubation for 10 min followed by three repetitions of centrifugation (16,000 × g for 1 min at room temperature) and resuspension in MSM (450 mM NaCl, 50 mM MgSO 4 · 7H 2 O, 50 mM Tris-HCl, pH 7.5), each sample was immediately fixed with glutaraldehyde (0.25% final concentration) and ascorbic acid antifade solution (1% final concentration, [83]). In parallel, to compare the removal of free viruses, we additionally prepared the sample (phage HS8 and H71 cells) without centrifugation and resuspension (Fig. S7). ...
Article
Full-text available
Viral metagenomics (viromics) has reshaped our understanding of DNA viral diversity, ecology, and evolution across Earth’s ecosystems. However, viromics now needs approaches to link newly discovered viruses to their host cells and characterize them at scale. This study adapts one such method, sequencing-enabled viral tagging (VT), to establish “Viral Tag and Grow” (VT + Grow) to rapidly capture and characterize viruses that infect a cultivated target bacterium, Pseudoalteromonas. First, baseline cytometric and microscopy data improved understanding of how infection conditions and host physiology impact populations in VT flow cytograms. Next, we extensively evaluated “and grow” capability to assess where VT signals reflect adsorption alone or wholly successful infections that lead to lysis. Third, we applied VT + Grow to a clonal virus stock, which, coupled to traditional plaque assays, revealed significant variability in burst size—findings that hint at a viral “individuality” parallel to the microbial phenotypic heterogeneity literature. Finally, we established a live protocol for public comment and improvement via protocols.io to maximally empower the research community. Together these efforts provide a robust foundation for VT researchers, and establish VT + Grow as a promising scalable technology to capture and characterize viruses from mixed community source samples that infect cultivable bacteria.
... We acknowledge that this was not an ideal set-up to detect Vibrio phages, and we cannot completely rule out phage activity at this time. Ideally, future experiments should analyze the viral abundance throughout the experimental timeframe via the methods discussed in John et al. (2011) and Cunningham et al. (2015). ...
Article
Full-text available
The potential spread of infectious diseases in response to climate change and rising sea surface temperatures in temperate regions has been a growing concern for the past several decades. Extreme heat waves in the North Atlantic and North Sea regions have been correlated with an increase in human Vibrio infections; of particular concern to human health are Vibrio cholerae, Vibrio parahaemolyticus, and Vibrio vulnificus. While these species are well-known to cause disease in humans, most environmental strains are not pathogenic. Studying not only the behavior of the pathogenic strains, but that of non-pathogenic environmental isolates, may better elucidate their ecological relationship in their native microbiome and the dispersal of these species in coastal regions. Using red fluorescent protein-tagged and gentamycin-resistant V. cholerae, V. parahaemolyticus, and V. vulnificus strains, we investigated whether increasing temperatures confer greater competitive fitness to these species when incubated within a natural North Sea water sample still containing its microbiome in a small-scale niche investigation. Increased incubation temperatures alone did not confer a competitive advantage to V. cholerae, V. parahaemolyticus, and V. vulnificus. The microbial community could limit Vibrio growth at all temperatures. To the best of our knowledge, we also demonstrate the first (albeit unintentional) genetic modification of multiple species of marine bacteria through the introduction of a genetically modified V. vulnificus strain into a natural water sample in a contained system.
... Gradient fractions were collected by sipping from the top of the tube using a piston fractionator (Labconco Auto Densi-Flow). The viruscontaining fractions were identified by a combination of epifluorescence microscopy(Cunningham et al., 2015) and fluorometric DNA measurements using the Quant-iT dsDNA high sensitivity assay kit (ThermoFisher Scientific; P/N: Q33120). CsCl was removed from the virus-containing fractions by gradually diluting the sample with MSM buffer (50% volume MSM was added to the sample, mixed, incubated for 5 min, and repeated until reaching reach a final dilution ratio of approx. ...
Thesis
Full-text available
Death by viral infection rivals predation as a source of mortality for all types of microscopic plankton in the ocean, including phytoplankton that are the foundation of the marine food web. This has profound consequences for plankton ecology and nutrient cycling in the sea. Viruses tend to be quite specific in the cells they infect, so the known extraordinary diversity among the marine phytoplankton implies that there is a similar high diversity of viruses in the sea. Our knowledge of viral diversity in the ocean has dramatically improved in recent years using metagenomic techniques (random sequencing of genome fragments from mixed communities). However, basic information about the viruses being detected, such as which organisms they infect and the details of their infection cycle cannot be reliably determined from sequence data alone. Having more model virus-host systems in culture that can be experimentally manipulated and studied in the lab would provide valuable new insights into the functional roles of the viruses in the marine food web. In this dissertation, cultivation-based techniques were used to characterize novel virus-host systems for eukaryotic phytoplankton from the tropical North Pacific Ocean, thereby identifying new virus-host linkages and establishing model systems for further study. Over 300 phytoplankton strains were cultivated and used in the isolation of over 60 virus strains. Described herein is a summary of these isolation efforts, including preliminary characterizations for 19 virus isolates using electron microscopy and genome sequencing. This is followed by more in-depth analyses of the genome and virion proteome of the giant virus Tetraselmis virus 1 (TetV-1), which infects the cosmopolitan green alga Tetraselmis. This work establishes new virus-host linkages and highlights previously uncharacterized viral diversity, including the first protist-infecting isolates from viral families previously only known to infect plants or animals. Furthermore, the genomic analyses have revealed a high number of viral encoded metabolic genes not previously seen in viruses, and the proteomic analyses have identified novel virion-associated enzymes. Marine viruses continue to represent an enormous amount of unknown or uncharacterized taxonomic and metabolic diversity, and this work demonstrates the utility of cultivation-based approaches in illuminating some of these mysteries.
... Cells captured on the filters were subjected to DNA extraction and microbial metagenomic sequencing. Virus-like particles (VLP) in the filtrate were counted using the wet-mount method (39). For the sea-ice brine sample, viruses in the filtrate were concentrated using an iron chloride flocculation method (95) and stored at 4°C at BARC. ...
Article
Full-text available
This study explores viral community structure and function in remote and extreme Arctic environments, including subzero brines within marine layers of permafrost and sea ice, using a modern viral ecogenomics toolkit for the first time. In addition to providing foundational data sets for these climate-threatened habitats, we found evidence that the viruses had habitat specificity, infected dominant microbial hosts, encoded host-derived metabolic genes, and mediated horizontal gene transfer among hosts. These results advance our understanding of the virosphere and how viruses influence extreme ecosystems. More broadly, the evidence that virally mediated gene transfers may be limited by host range in these extreme habitats contributes to a mechanistic understanding of genetic exchange among microbes under stressful conditions in other systems.
Preprint
Full-text available
Enumeration is a fundamental measure of community ecology in which viruses represent the most numerous biological identities. Epifluorescence microscopy (EFM) has been the gold standard method for environmental viral enumeration for over 25 years. Currently, standard EFM methods using the Anodisc filters are no longer cost-effective (>$15 per slide) and have yet to be applied to modern microbialites. Microbialites are microbially driven benthic organosedimentary deposits that have been present for most of Earth’s history. We present a cost-effective method for environmental viral enumeration from aquatic samples, microbial mats, and exopolymeric substances (EPS) within modern microbialites using EFM. Our integrated approach, which includes filtration, differential centrifugation, chloroform treatment, glutaraldehyde fixation, benzonase nuclease treatment, probe sonication (EPS and mat only), SYBR Gold staining, wet-mounting, and imaging, provides a robust method for modern microbialites and aquatic samples. Viral abundances of modern microbialites and aquatic samples collected from Fayetteville Green Lake (FGL) and Great Salt Lake (GSL) did not differ across ecosystems by sample type. EPS and microbial mat samples had an order of magnitude higher viral-like particle (VLP) abundance when compared to water regardless of the ecosystem (10 ⁷ vs. 10 ⁶ ). Viral enumeration allows for estimates of total viral numbers and weights. The entire weight of all the viruses in FGL and GSL are ∼598 g and ∼2.2 kg, respectively; this is equivalent to a loaf of bread for FGL and standard brick for GSL. Further development of EFM methods and software is needed for viral enumeration. Our method provides a robust and cost-effective (∼$0.75 per sample) viral enumeration within modern microbialites and aquatic ecosystems.
Article
The rhizosphere is a vital soil compartment providing key plant‐beneficial functions. However, little is known about the mechanisms driving viral diversity in the rhizosphere. Viruses can establish lytic or lysogenic interactions with their bacterial hosts. In the latter, they assume a dormant state integrated in the host genome and can be awakened by different perturbations that impact host cell physiology, triggering a viral bloom, which is potentially a fundamental mechanism driving soil viral diversity, as 22‐68% of soil bacteria are predicted to harbor dormant viruses. Here we assessed the viral bloom response in rhizospheric viromes by exposing them to three contrasting soil perturbation agents: earthworms, herbicide and antibiotic pollutant. The viromes were next screened for rhizosphere‐relevant genes and also used as inoculant on microcosms incubations to test their impacts on pristine microbiomes. Our results show that while post‐perturbation viromes diverged from control conditions, viral communities exposed to both herbicide and antibiotic pollutant were more similar to each other than those influenced by earthworms. The latter also favored an increase in viral populations harboring genes involved in plant‐beneficial functions. Post‐perturbation viromes inoculated in soil microcosms changed the diversity of pristine microbiomes, suggesting that viromes are important components of the soil ecological memory driving eco‐evolutionary processes that determines future microbiome trajectories according to past events. Our findings demonstrate that viromes are active players in the rhizosphere and need to be considered in efforts to understand and control the microbial processes towards sustainable crop production.
Article
Full-text available
A new nucleic acid stain, SYBR Green I, can be used for the rapid and accurate determi-nation of viral and bacterial abundances in diverse marine samples. We tested this stain with formalin-preserved samples of coastal water and also from depth profiles (to 800 m) from sites 19 and 190 km off-shore, by filtering a few m1 onto 0.02 pm pore-size filters and staining for 15 min. Comparison of bacterial counts to those made with acridine orange (AO) and virus counts with those made by trans-mission electron microscopy (TEM) showed very strong correlations. Bacterial counts with A 0 and SYBR Green 1 were indistinguishable and almost perfectly correlated (r2 = 0.99). Virus counts ranged widely, from 0.03 to 15 X 10' virus ml-l. Virus counts by SYBR Green 1 were on the average higher than those made by TEM, and a SYBR Green 1 versus TEM plot yielded a regression slope of 1.28. The cor-relation between the two was very high with an value of 0.98. The precision of the SYBR Green I method was the same as that for TEM, with coefficients of variation of 2.9%. SYBR Green I stained viruses and bacteria are intensely stained and easy to distinguish from other particles with both older and newer generation epifluorescence microscopes. Detritus is generally not stained, unlike when the alternative dye YoPro I is used, so this approach may be suitable for sediments. SYBR Green I stained samples need no desalting or heating, can be fixed with formalin prior to filtration, the optimal staining time is 15 min (resulting in a total preparation time of less than 25 min), and counts can be easily per-formed at sea immediately after sampling. This method may facilitate incorporation of viral research into most aquatic microbiology laboratories.
Article
Full-text available
Epifluorescent microscopy was used to determine the abundance of viruses in samples from marine and freshwater environments and in laboratory cultures that were filtered onto 0.02-pm pore-size filters and stained with a cyanine-based dye (Yo-Pro-l). Estimates of viral abundance based on Yo-Pro-stained samples were 1.2-7.1 times greater than estimates obtained with transmission electron microscopy (TEM). Moreover, the precision ofthe Yo-Pro-based method was much greater than that for TEM (C.V. 7% vs. 20%, respectively). DNase treatment of samples did not result in lower numbers of particles that could be stained by Yo-Pro, suggesting that the fluorescence was not the result of nucleic acids associated with the surface of particles. These results indicate that the concentration of viruses in natural waters may be higher than previously recognized and imply that the TEM-based method significantly underestimates virus abundance. Virus abundances ranged from 1 07-> 1 OS ml-l in surface waters along a transect in the western Gulf of Mexico to 1 O9 ml-l in water overlying a submerged cyanobacterial mat. High counting efficiency, ease of preparation, modest equipment requirements, and the possibility of preparing specimens for long-term storage, make the Yo-Pro-based method ideal for routine environmental analysis.
Article
Full-text available
Transmission electron microscopy (TEM) and epifluorescence microscopy of DAPI and Yo-Pro-1 stained samples were used to estimate viral abundance in natural communities along a transect from the oligotrophic central Gulf of Mexico to the productive near-shore waters at Post Aransas, Texas (USA). Estimates of viral abundance based on TEM averaged only 66% (range 26 to 108%) of those made using epifluorescence microscopy and the cyanine-based dye, Yo-Pro-1. DAPI staining provided estimates that were much closer and averaged 86% (range 72 to 109%) of those made using Yo-Pro. However, all 3 methods provided similar estimates at viral abundances <10(6) ml(-1). The precision of the Yo-Pro and DAPI methods (coefficient of variation 8 and 11%, respectively) was much greater than for the TEM method (25%). Experiments with cultures indicated that grazing by flagellates was unlikely to be a significant source of viral-size particles that could interfere with the DAPI or Yo-Pro method. Estimates of viral abundance made using the Yo-Pro method ranged from 0.3 x 10(6) to 39 x 10(6) ml(-1) in surface water along the transect. Across the investigated environments viral and bacterial abundances were well correlated (r = 0.929), although the slope of the relationship was significantly greater than 1, indicating that viral abundance increased more rapidly than that of bacteria. These results extend previous observations by showing that epifluorescence microscopy is suitable for counting viruses in very oligotrophic waters, that DAPI and Yo-Pro stained samples provide similar estimates of viral abundance and that grazing by flagellates is not a significant source of particles that could interfere with the epifluorescence method. The study supports the use of epifluorescence microscopy over TEM for obtaining accurate estimates of viral abundances in natural waters.
Article
Full-text available
Viruses influence oceanic ecosystems by causing mortality of microorganisms, altering nutrient and organic matter flux via lysis and auxiliary metabolic gene expression and changing the trajectory of microbial evolution through horizontal gene transfer. Limited host range and differing genetic potential of individual virus types mean that investigations into the types of viruses that exist in the ocean and their spatial distribution throughout the world's oceans are critical to understanding the global impacts of marine viruses. Here we evaluate viral morphological characteristics (morphotype, capsid diameter and tail length) using a quantitative transmission electron microscopy (qTEM) method across six of the world's oceans and seas sampled through the Tara Oceans Expedition. Extensive experimental validation of the qTEM method shows that neither sample preservation nor preparation significantly alters natural viral morphological characteristics. The global sampling analysis demonstrated that morphological characteristics did not vary consistently with depth (surface versus deep chlorophyll maximum waters) or oceanic region. Instead, temperature, salinity and oxygen concentration, but not chlorophyll a concentration, were more explanatory in evaluating differences in viral assemblage morphological characteristics. Surprisingly, given that the majority of cultivated bacterial viruses are tailed, non-tailed viruses appear to numerically dominate the upper oceans as they comprised 51-92% of the viral particles observed. Together, these results document global marine viral morphological characteristics, show that their minimal variability is more explained by environmental conditions than geography and suggest that non-tailed viruses might represent the most ecologically important targets for future research.
Article
Full-text available
Ocean viruses alter ecosystems through host mortality, horizontal gene transfer and by facilitating remineralization of limiting nutrients. However, the study of wild viral populations is limited by inefficient and unreliable concentration techniques. Here, we develop a new technique to recover viruses from natural waters using iron-based flocculation and large-pore-size filtration, followed by resuspension of virus-containing precipitates in a pH 6 buffer. Recovered viruses are amenable to gene sequencing, and a variable proportion of phages, depending upon the phage, retain their infectivity when recovered. This Fe-based virus flocculation, filtration and resuspension method (FFR) is efficient (> 90% recovery), reliable, inexpensive and adaptable to many aspects of marine viral ecology and genomics research.
Article
A new nucleic acid stain, SYBR Green I, can be used for the rapid and accurate determination of viral and bacterial abundances in diverse marine samples. We tested this stain with formalin-preserved samples of coastal water and also from depth profiles (to 800 m) from sites 19 and 190 km offshore, by filtering a few mi onto 0.02 mu m pore-size filters and staining for 15 min. Comparison of bacterial counts to those made with acridine orange (AO) and virus counts with those made by transmission electron microscopy (TEM) showed very strong correlations. Bacterial counts with AO and SYBR Green I were indistinguishable and almost perfectly correlated (r(2) = 0.99). Virus counts ranged widely, from 0.03 to 15 x 10(7) virus ml(-1). Virus counts by SYBR Green I were on the average higher than those made by TEM, and a SYBR Green I versus TEM plot yielded a regression slope of 1.28. The correlation between the two was very high with an r(2) value of 0.98. The precision of the SYBR Green I method was the same as that for TEM, with coefficients of variation of 2.9%. SYBR Green I stained viruses and bacteria are intensely stained and easy to distinguish from other particles with both older and newer generation epifluorescence microscopes. Detritus is generally not stained, unlike when the alternative dye YoPro I is used, so this approach may be suitable for sediments. SYBR Green I stained samples need no desalting or heating, can be fixed with formalin prior to filtration, the optimal staining time is 15 min (resulting in a total preparation time of less than 25 min), and counts can be easily performed at sea immediately after sampling. This method may facilitate incorporation of viral research into most aquatic microbiology laboratories.
Chapter
For many laboratories, flow cytometry is becoming the routine method for quantifying viruses in aquatic systems because of its high reproducibility, high sample throughput, and ability to distinguish several subpopulations of viruses. Comparison of viral counts between flow cytometry and epifluorescence microscopy typically shows slopes that are statistically not distinguishable from 1, thus confirming the usefulness of flow cytometry. Here we describe in detail all steps in the procedure, discuss potential problems, and offer solutions.
Chapter
To circumvent the limitations of cultivation-based studies of complex microbial communities, molecular fingerprinting techniques such as pulsed field gel electrophoresis (PFGE) and denaturing gradient gel electrophoresis (DGGE) have been used to examine their richness, diversity, and dynamics. PFGE is based on the electrophoretic separation of extremely large DNA, raising the upper size limit from 50 kb (standard agarose separation) to well over 10 Mb. This technique has been used to separate aquatic virus genomes ranging in size from tens to hundreds of kilo base pairs (kb); aquatic virus genomes range from 15 to 630 kb, with the majority between 20 and 80 kb. DGGE, on the other hand, is based on the electrophoretic separation of PCR- amplified gene fragments of similar sizes, but differing in base composition or sequence. In this chapter, we provide a brief overview of each of these methods and their application to the study of aquatic viruses. We describe some of the common equipment, reagents, and procedures involved, and conclude by briefly considering some of the strengths and weaknesses of each method.