<<

MARINE ECOLOGY PROGRESS SERIES Published January 12 Mar. Ecol. Prog. Ser.

Growth and mortality rates of and measured with a selective inhibitor techniaue

Hongbin Liu, Lisa Campbell, Michael R. Landry

Department of Oceanography, School of and Earth Science and Technology, University of Hawaii, Honolulu, Hawaii 96822, USA

ABSTRACT: A selective metabolic inhibitor method has been developed to estimate growth rates and mortalities due to protozoan grazing of the photoautotrophic prokaryotic Prochlorococcus and Synechococcus. Laboratory and field experiments show that 1 mg mll (final concentration) kana- mycin inhibits the growth of Prochlorococcus and Synechococcus effectively and does not significantly affect protozoan grazing At Station ALOHA (22' 45' N, 158' W) 100 km north of Oahu, Hawaii, USA, growth rates of Prochlorococcus ranged from 0.4 to 0.5 d1within the surface mixed layer to about 0.1 d1at the bottom of the euphotic zone. Synechococcus grew faster, with a daily growth rate of up to 1.0 d1in surface waters. Grazing mortalities varied for Prochlorococcus and Synechococcusfrom 20 to 116 % and 43 to 87 % of growth rates, respectively. Growth generally exceeded grazing. Because of its high abundance (up to 2 x lo5 cells mll in the upper 100 m), Prochlorococcus contributes significantly to and primary production in the subtropical North Pacific Ocean. At Station ALOHA in October 1993, integrated (0 to 175 m) carbon production due to Prochlorococcus was 382.2 mg C m2dl. In contrast, Synechococcus produced only 14.6 mg C m2dl.

KEY WORDS: Growth Mortality . Prochlorococcus . Synechococcus . Selective inhibitor FLB

INTRODUCTION ence of Prochlorococcus confounded these estimates. Prochlorococcus,a new group of oxyphototrophic ma- Within the last decade, photosynthetic picoplankton rme containing &vinyl chlorophylls a and (<2 pm cells) have been found to account for more b, were discovered in the mid-1980s by Chisholm et than half of the biomass and primary production in al. (1988) using . Prochlorococcus is the tropical and subtropical open ocean (Li et al. 1983, now recognized as a major component of the pico- Platt et al. 1983, Takahashi & Bienfang 1983, Herb- in warm oceanic waters of the Atlantic land et al. 1985). Among them, 2 groups of prokary- Ocean (Li & Wood 1988, Olson et al. 1990), the Pacific otic cells, Synechococcus and Prochlorococcus,were Ocean (Chavez et al. 1991, Campbell & Vaulot 1993), reported to be the predominant components in this as well as the Mediterranean (Vaulot et al. 1990) and size class. The -containing unicellular Red Seas (Veldhuis & Kraay 1993). The relative Synechococcus were the first to be importance of Prochlorococcus and Synechococcus studied in detail because they are easily recognized to the total photosynthetic biomass varies among by their distinctive orange fluorescence under the epi- oceanic regions. While Synechococcus contributes fluorescence microscope (Johnson & Sieburth 1979, only 2% of particulate organic carbon (POC) in the Waterbury et al. 1979, Olson et al. 1990a).Early stud- central North Pacific, Prochlorococcus contributes up ies reported that Synechococcus contributed up to to 30% of POC in the upper 200 m water column 95% of total primary production (Glover 1985, Itur- (Campbell et al. 1994). In the northern Indian Ocean, riaga & Mitchell 1986, Waterbury et al. 1986, Iturriaga Synechococcus accounts for up to 40% of the POC & Marra 1988). However, as we now realize, the pres- (Burkill et al. 1993). However, because of the overlap

0 Inter-Research 1995 Resale of full article not permitted Mar. Ecol. Prog. Ser. 116: 277-287, 1995 - in cell size between Prochlorococcus and Synechococ- mortality of Prochlorococcus and Synechococcus popu- cus,the contribution of either individual group to total lations. primary production cannot be obtained accurately by size-fractionated photosynthetic measurements. The growth rate of Synechococcus has been studied MATERIALS AND METHODS quite intensively (Landry et al. 1984, Campbell & Car- penter 1986a, b, Iturriaga & Mitchell 1986, Iturriaga & Culture. Prochlorococcus and Synechococcus strains Marra 1988, Burkill et al. 1993). In contrast, there are were isolated from Station ALOHA, located about only a few recent reports of the growth rates of Pro- 100 km north of Oahu, Hawaii, USA, in the central chlorococcus (Goericke & Welschmeyer 1993, Vaulot Pacific Ocean (22' 45' N, 158' W). Prochlorococcus was et al. 1994). There are many indirect and direct meth- cultured in PC media (modified K/2) as recommended ods to estimate a population's growth rate (see review by the Culture Collection of Marine Plankton at Bige- by Furnas 1990). Because of grazing, the change in cell low Laboratory (West Boothbay Harbor, ME, USA). abundance in the water column reveals the net growth Synechococcus was grown in f/2 media. Neither rate. The dilution method (Landry & Hassett 1982, Prochlorococcus nor Synechococcus cultures are Landry et al. 1984) and selective metabolic inhibition axenic. All culture containers were carefully cleaned (Newel1 et al. 1983, Fuhrman & McManus 1984) tech- by first soaking in 10% hydrochloric acid (HC1) for at niques are the 2 most frequently used methods to esti- least 24 h. The containers were then rinsed with dis- mate grazing mortalities and compute absolute growth tilled deionized water (DDW) several times, soaked in rates of phytoplankton. DDW for more than 24 h, rinsed several times, and The use of selective metabolic inhibitors in the study autoclaved at 12loCfor 20 min. of marine phytoplankton was first reported for eukary- Laboratory experiments. Kanamycin stock solutions otes (Thomas & Dodson 1974).Newel1 et al. (1983),for (10 to 100 mg ml)were made immediately prior to example, applied the eukaryotic inhibitors cyclohex- use by dissolving kanamycin (Sigma, K-4000) in DDW imide and thiram to estimate grazing rates on . and filter sterilizing (Acrodisc 0.2 pm syringe filters). Fuhrman & McManus (1984) modified this method by To determine the kanamycin concentration required using benzylpenicillin as a prokaryotic inhibitor for to inhibit growth without causing cell lysis or death, we estimating rates of bacterivory in Long Island Sound performed growth experiments with final kanamycin (New York, USA). Sherr et al. (1986) compared the concentrations of 0, 0.01, O.l'O, 0.50 and 1.0 mg mT1, eukaryotic inhibitors (cycloheximide + colchicine) and Triplicate cultures of Prochlorococcus and Synecho- prokaryotic inhibitors (vancomycin + penicillin) in the coccus were grown in 20 ml glass test tubes under con- study of trophic interactions between heterotrophic tinuous illumination of approximately 20 pEin m2s-', protozoa and in estuarine waters. Samples (0.5 ml) were taken at 6 h intervals from each Ampicillin, a prokayotic inhibitor, has also been used test tube using sterile technique, preserved with to estimate the growth rate of and grazing pressure on paraformaldehyde (0.2 % final concentration), frozen Synechococcus (Campbell & Carpenter 1986b). How- quickly in liquid nitrogen, and stored at -80°C ever, none of the above have proved satisfactory for Thawed samples were analyzed by flow cytometry fol- Prochlorococcus (M. R. Landry unpubl.). Through a lowing the method of Vaulot et al. (1989). series of laboratory and field tests, we found that To test whether kanamycin affected the grazing kanamycin is an effective growth inhibitor for both activity of protozoans, 2 enrichment cultures, a het- Prochlorococcus and Synechococcus. Kanamycin in- erotrophic nanoflagellate (HNAN) and a mixture of hibits protein synthesis by combining with the ribo- nanoflagellates and ciliates from Kaneohe Bay (KEW), somes to disrupt the genetic code and induce the were grown in rice culture media with 0, 0.1 and production of mutant forms (Pelczar et al. 1977). The 1.0 mg mlkanamycin. FLB were made from Vibrio formation of faulty membrane-associated proteins ren- damsela with 5-([4,6-dichlorotriazh-2-yl]amino)-fluo- ders the membrane leaky, causing the loss of essential rescein (DTAF) following the procedure of Sherr et al. small molecules and enhanced uptake of more kana- (1987), and were added to triplicate experimental mycin, which is then sufficient to prevent further pro- tubes at a final concentration of 20000 mll. Samples tein synthesis (Greenwood 1989). The key problem in (0.5 ml) were taken every 6 h for flow cytometric deter- using the selective metabolic inhibitor approach is the mination of FLB abundance. Grazing rates were com- specificity of the inhibitor. Therefore, we advocate the puted from the rate of decline of FLB concentration addition of fluorescently labeled bacteria (FLB) as an with time. internal control in all incubations to test this critical Field experiments. Field experiments were conducted assumption. We report preliminary results of the kana- in Kaneohe Bay, Hawaii, during February 1993 and at mycin method for estimating growth rates and grazing the Hawaii Ocean Time-series (HOT) Station ALOHA in Liu et a1 . Growth and mortality rates of Prochlorococcus and Synechococcus 279

May, September and October 1993. As described below, 5 W Argon lasers and MSDS automatic sampling. Filter some details varied among experiments. Similarities in- set-ups were as previously described by Campbell & cluded incubations of triplicate 250 ml polycarborate Vaulot (1993). For Prochlorococcus and Synechococ- bottles for each treatment and control. In addition, FLB cus,we used an excitation wavelength of 488 nm (1W) were added to all bottles at a final concentration of to collect green (525 nm  40 df), orange (575 nm & 10 000 cells mll to test whether the kanamycin inhib- 40 df) and red (680 run  40 df) fluorescence and right ited grazing activity and/or to correct for an inhibitor angle light scatter (RALS),which varies as a function of effect on grazing if it occurred. For individual experi- cell size. Sample volumes were 100 pl for most analy- ments, we used the difference between FLB dis- ses, however, 200 to 300 pl were analyzed for some appearance rates in the incubation treatments and samples to obtain a statistically significant cell number controls to compute inhibitor-corrected growth and for Synechococcus.Prochlorococcus and Synechococ- grazing rates for Prochlorococcus and Synechococcus. cus populations were distinguished from heterotrophic In most cases, we observed a small effect of kanamycin bacteria by their respective red (divmyl chlorophyll) on grazing activity based on the difference m FLB dis- and orange () autofluorescences under appearance rates between treatments. The difference 488 nm excitation. FLB were characterized by bright was most often statistically insignificant due to a large green fluorescence. In laboratory experiments, sub- standard deviation and small sample number (see samples for enumeration of heterotrophic bacteria Table 1). We did, however, correct our grazing and were stained with 1 pg mll Hoechst 33342 (Molecular growth rates because of the apparent effect. Probes, Eugene, OR, USA) and analyzed with dual Initial experiments in Kaneohe Bay were conducted UV/488 nm excitation (Monger & Landry 1993). Het- to verify the effective kanamycm concentration for nat- erotrophic bacteria have blue fluorescence under UV ural populations. Seawater samples were collected excitation (DNA content) but lack any autofluores- before dawn from the sea surface inside and from adja- cence. List mode files were analyzed on an IBM PC- cent oceanic waters outside the bay. Kanamycin was compatible microcomputer using CytoPC software added to the bottles to yield fmal concentrations of 0 (Vaulot 1989). (control), 0.01, 0.1 and 1.0 mg mi''. All bottles were Calculations. Instantaneous rates of population mcubated in situ in surface waters near Coconut Island growth and grazing mortality were calculated for in Kaneohe Bay for 24 h. A 1 ml subsample was taken Prochlorococcus and Synechococcus as described by from each bottle every 6 h using the same procedure Campbell & Carpenter (1986b). Protozoan grazing on described for laboratory experiments. FLB in all experiments were determined from the same For experiments at Station ALOHA, Go-Flo bottles exponential model. Integrated growth rates (d-l)in the were used to collect seawater before sunrise from 5,45 upper 175 m water column are calculated from: and 100 m in May and 75 and 125 m m September. Based on results from Kaneohe Bay, only 1 kanamycin concentration (1 mg mll) and a control were pre- pared. Triplicate bottles for the treatment and control where No is cell abundance at the beginning time of were incubated in a shipboard temperature- and light- incubation, NZ4is cell abundance after the 24 h incu- controlled incubator, and an additional set of triplicate bation without accounting for grazing (i.e.N24 = Nee^), samples was incubated in situ in order to compare the and z is the sampling depth. For the experiments at 2 incubation techniques. The in situ incubation lasted Station ALOHA, carbon production of Prochlorococcus between dawn and dusk, and subsamples for flow and Synechococcus were calculated from cell abun- cytometric analyses were taken at the beginning and dance, growth rates and a carbodcell conversion fac- end of the experiment. Subsamples were taken every tor and compared with total primary production mea- 6 h during the 24 h shipboard incubations. sured by 14C method (data provided by Joint Global For the October experiment, seawater was collected Ocean Flux Study). We assumed conversion factors of from 8 depths from 5 to 175 m. The incubation proce- 53 and 250 fg C cell1for Prochlorococcus and Syne- dure was also modified because previous results chococcus,respectively (Campbell et al. 1994).Carbon revealed that cell division of Prochlorococcus occurred production (P,mg C m3d) at each sampling depth more during nighttime and, therefore, that dawn-to- was calculated as: dusk incubations underestimated specific growth rates. Triplicate bottles were incubated in situ from dawn-to-dusk followed by shipboard mcubation after where N (cell mll dl)is the number of cells produced dark to complete a 24 h incubation. in 24 h which can be computed from Flow cytometry. All samples were analyzed using a Coulter EPICS 753 flow cytometer equipped with two 280 Mar Ecol. Prog. Ser 116: 277-287, 1995

RESULTS Kaneohe Bay field studies

Laboratory experiments The results of the Kaneohe Bay experiments showed a slight decrease of protozoan grazing on FLB in Laboratory incubations with different concentrations samples with 1 mg mlof kanamycin, but no signifi- of kanamycin indicated that concentrations of 1 and cant difference from controls (Table 1, Fig. 3b, c). 0.5 mg mllprevented cell division of Prochlorococcus The slower decrease of Prochlorococcus abundance and Synechococcus, respectively (Fig. 1). Kanamycin in the 1 mg mll kanamycin incubations than in had no effect on the growth of heterotrophic bacteria other treatments also suggests that protozoan grazing co-occurring in the cultures of laboratory experiments. activity was affected, in particular during the first Grazing experiments showed no statistically sig- few hours after adding kanamycin to incubations nificant difference in protozoan grazing on FLB be- (Fig. 3a). However, this grazing inhibition during the tween the control and kanamycin treatments (Fig. 2) first 6 h was not observed clearly for Synechococcus (ANOVA: F= 2.13, p = 0.234, df = 2 for HNAN; F= 0.86, (Fig. 3d, e). p = 0.469, df = 2 for KEW). We chose a kanamycin The growth rates of Synechococcus measured in concentration of 1 mg mll for all subsequent field Kaneohe Bay and in oceanic waters were similar, 0.70 experiments. and 0.58 d,respectively. Prochlorococcus growth rate was substantially lower, 0.33 d1for oceanic waters. Proch1orococcu.s were not observed inside Kaneohe Bay. The daily grazing mortality of Prochloro- coccus and Synechococcus calculated from 24 h in situ incubations revealed that only about 30% of the pro-

0 5 10 15 20 25 TIME (h)

Fig 1 Prochlorococcus (Pacific strain) and Synechococcus. Fig. 2. Results of laboratory experiments to test the effect of Laboratory experiments with different concentrations of kanamycin on protozoan grazing. Data show no difference of kanamycin under continuous light. For Synechococcus, protozoan grazing on fluorescently labeled bacteria (FLB) be- curves of control and 0 5 mg kanamycin mll (final concen- tween treatments. HNAN: a culture of a heterotrophic nano- tration) are significantly different (t-test, p < 0.005). For flagellate (ANOVA, F=2.13, p = 0.234, df = 2); KEW: a mix- Prochlorococcus, 1 mg kanamycin mll prevents cell division ture of nanoflagellates and ciliates (ANOVA, F = 0 86, p = (t-test, p < 0.0025) Error bars show 1 SD of the means of 0.469, df = 2). Error bars show 1 SD of the means of triplicate triplicate incubations incubations Liu et al.: Growth and mortality rates of Prochlorococcus and Synechococcus 281

Table 1. The difference in protozoan grazing on FLB in the 0.05 d at 75 m in September) because cell division control and inhibitor (1 mg mlkanamycin) incubations of Prochlorococcus occurs at a higher rate during the tested by t-test. Data are the means of triplicate measure- night (Fig. 5). In October, a higher Prochlorococcus ments  1 SD calculated from 24 h incubations. In the May 1993 experiment, rates from dawn-to-dusk incubation are growth rate of 0.59 d-I (SE = 0.09) was found at 25 m. given in parentheses, *p< 0.05; ns: p > 0.05 Again, the growth rates of Prochlorococcus were low- est at 75 m (0.09 dl) and increased slightly below Location Grazing rates on FLB (dl) t-test (Table 2, Fig. 7). Depth Control Inhibitor Synechococcus grew faster than Prochlorococcus, with growth rates ranging from 0.39 to 0.70 d-I KaneoheBay throughout the upper 100 m water column during May Bay, surface 0.47  0 07 Oceanic, surface 0.51  0.06 and September (Table 2). In contrast, during October Synechococcus growth rate was much higher (1.06 d-l, Station ALOHA SE = 0.09) near the sea surface but decreased to May 1993 0.17 d-I (SE = 0.04) at 75 m (Table 2, Fig. 7). The mte- 5 m grated growth rates in the upper 175 m water column are 0.26 dlfor Prochlorococcus and 0.44 d-l for Syne- chococcus. Mortality rates of Prochlorococcus measured in 24 h incubations at Station ALOHA ranged from 0.04 to October 1993 5m 0.50 d within the upper 175 m water column 25 m (Table 2). The ratio of grazing to growth estimates 45 m ranged from 0.20 to 1.16. Grazing on Synechococcus 75 m seemed less variable with grazing to growth ratios of 100 m 0.43 to 0.87. Grazing mortality (especially for Pro- 125 m 150 m chlorococcus),measured in in situ dawn-to-dusk incu- 175 m bations, was higher than in 24 h incubations and always resulted in negative net growth rates for Pro- Overall (n = 40) 0.32  0.18 chlorococcus. Grazing of FLB showed the same pattern duction of Prochlorococcus in the surface Table 2. Prochlorococcus and Synechococcus. Grazing rates (g)and growth water was lost to grazing. In comparison, coefficients (p)in Kaneohe Bay and at Station ALOHA measured by selective the grazing pressure on Synechococcus inhibitor method. Data are the means of triplicate 24 h incubations  1 SD spp. at both sites was 38 to 56% of its daily growth (Table 2). Location Prochlorococcus Synechococcus Depth P (d-l) fir (d-l) !J fir

Kaneohe Bay Station ALOHA field studies Inside, surface Outside, surface 0.33 Â 0.02 Results from 24 h deck incubations Station ALOHA during the first experiment (May 1993; May 1993 Figs. 4 to 6) indicated that Prochlorococ- 5m cus grows slowly under oligotrophic con- 45 m 0.20 Â 0.03 ditions. Calculated growth rates aver- 100 m 0.10 Â 0.09 aged 0.20 (k0.03) d-' at 45 m and 0.10 September 1993 (k0.09) d-I at 100 m. Growth rates of 75 m 0.09 Â 0.01 125 m 0.17 Â 0.06 Prochlorococcus in the September exper- iments were similar with higher rates October 1993 5 m 0.38 Â 0,10 measured at 125 m (0.17 Â 0.06 d-l) than 25 m 0.59 Â 0.19 at 75 m (0.09 Â 0.01 d-I). The growth 45 m 0.10 Â 0.03 rates of Prochlorococcus calculated from 75 m 0.09 Â 0.02 dawn-to-dusk (i.e. 12 to 14 h) in situ in- 100 m 0.15 Â 0.14 125 m 0.13 + 0.05 cubations in these experiments were 150 m 0.11 Â 0.10 lower (0.09 Â 0.06 d-I at 45 m and 0.04 Â 175 m 0.21 Â 0.11 0.01 d-I at 100 m in May, and 0.05 * 282 Mar. Ecol. Prog. Ser. 116: 277-287, 1995

5 10 15 20 25 TIME (h)

Fig. 3. Results of a field experiment in Kaneohe Bay, Hawaii (February 1993) using kanamycin as a prokaryotic inhibitor. (a) Prochlorococcus outside Kaneohe Bay; (b) Synechococcus in Kaneohe Bay; (c) Synechococcus outside Kaneohe Bay; (d) grazing on FLB in Kaneohe Bay; (e) grazing of FLB outside 0.81 5 10 15 20 25 Kaneohe Bay. Error bars show 1 SD of the means of TIME (h) triplicate incubations

which probably implies higher grazing activities of DISCUSSION protozoans during the daylight hours (Table 2). Over- all, the growth rates and grazing mortalities of both Selective metabolic inhibitors have several advan- Prochlorococcus and Synechococcus showed the simi- tages for estimating the growth rates of the marine lar trend with depth, but growth exceeded grazing in photosynthetic picoplankton Prochlorococcus and most cases (Table 2, Fig. 7). Synechococcus and their mortality due to grazing An alternative way of estimating the contribution of activities of heterotrophic nanoplankton. These in- both groups to total primary production is to determine clude minimal manipulation of seawater samples, no specific growth rates of Prochlorococcus and Syne- radioisotopes, and the ability to measure simultane- chococcus and convert to carbon production with car- ously growth and grazing for natural bacterial plank- bodcell conversion factors. The calculated carbon pro- ton populations (Campbell & Carpenter 1986b, Sherr duction of Prochlorococcus at Station ALOHA ranged et al. 1986). The application of the specific inhibitor from 10.05mg C m-3 dd"*in surface water to 0.04 mg C methods depends on both the effectiveness and speci- m3dl at the bottom of the euphotic zone (Table 3, ficity of the inhibitor. In our experiments, 1 mg mlY1 Fig. 7). It is interesting to note that Prochlorococcus kanamycin successfully inhibits cell division of Pro- contributed the most to production in the surface chlorococcus and Synechococcus. However, it also waters and at 125 m (Table 3).Synechococcus was only seems to exert a minor effect on the grazing activity of detected in the upper 100 m of the water column and protozoans (see Table 2). Although this effect was not accounted for less than 5% of the total primary pro- statistically significant in our experiments, it is essen- duction (Table 3, Fig. 7b). tial to include the FLB control in all grazing experi- Liu et al.: Growth and mortality rates of Prochloiococcus and Synechococcus 283

ments. One possible impact of the inhibition of proto- zoan activity is a decrease in growth rates of Prochloro- coccus and Synechococcus due to a decrease in the availability of remineralized nitrogen or phosphorus during the incubations. Another potential artifact is an alteration of the physiological status of prokaryotic picoplankton. We found that the chlorophyll fluores- cence of Prochlorococcus in the surface mixed layer samples in some cases decreased after incubation with the addition of kanamycin. This may have been due to an effect of kanamycin on chlorophyll synthesis or turnover rate. Our results suggest that a 24 h incubation period is critical for obtaining an unbiased growth rate because of die1 cycles of DNA synthesis and growth in Syne- chococcus (Campbell & Carpenter 1986a, Waterbury et

TIME (h)

Fig. 5. Prochlorococcus. Changes in density in 24 h shipboard incubations at Station ALOHA in May 1993. Error bars show 1 SE of the means of triplicate incubations. Control: natural seawater; inhibitor: 1 mg ml' kanamycin

al. 1986, Armbrust et al. 1989) and Prochlorococcus (Vaulot et al. 1994).The highest division rates of Syne- chococcus were realized in early evening, i.e. 12 to 15 h after dawn (Campbell & Carpenter 1986a),where- as Prochlorococcus started cell division after sunset (Vaulot et al. 1994). Therefore, dawn-to-dusk incuba- tions underestimate the 'true' daily growth rate, espe- cially for Prochlorococcus. Our results show that average daily growth rates for Prochlorococcus range from 0.4 to 0.6 d in near-sur- face waters and 0.1 to 0.2 d at the base of the euphotic zone. Our estimates are close to the observa- tions of Goericke & Welschmeyer (1993) for Prochloro- coccus in the Sargasso Sea near Bermuda using the I4C labeling technique. However, they are lower than the observation of Vaulot et al. (1994) for the equatorial Pacific (0.52 to 0.65 d-I integrated for upper 150 m) based on cell cycle analysis (Carpenter & Chang 1988). The growth rate of Synechococcus measured in Kaneohe Bay (0.58 to 0.70 d-l) and at Station ALOHA (0.17to 1.06 dl)are lower than those measured in pre- vious studies in the same bay (Landry et al. 1984) and in the oligotrophic North Pacific (Iturriaga & Mitchell Fig. 4. Changes in the density of FLB in 24 h shipboard mcu- bations at Station ALOHA m May 1993. Error bars show 1 SE 1986), but similar to the rates measured in the north- of the means of triplicate incubations. Control: natural sea- west Atlantic Ocean (Campbell & Carpenter 1986b) water; inhibitor: l mg mlkanamycin (final concentration) and the northwest Indian Ocean (Burkill et al. 1993). 284 Mar. Ecol. Prog. Ser. 116: 277-287, 1995

> 100 % of the estimated growth rates). The parity of growth and grazing rates for both Prochlorococcus and Synechococcus indicates that they are significant com- ponents of the marine pelagic food web and that most of their biomass is recycled fairly rapidly within the euphotic zone. Standard productivity experiments, such as post-incubation size-fractionated 14C experi- ments, may underestimate production of picoplankton because heterotrophic nanoplankton are not elimi- nated by pre-screening samples before incubation. Also, calculating carbon production from net growth rate (growth - grazing; Veldhuis et al. 1993) will greatly underestimate the production of Prochlorococ- cus and Synechococcus. One interesting phenomenon is the negative growth rate of Prochlorococcus observed in most dawn-to- dusk incubations. This trend could be explained by either a diel cycle in cell growth or heavier grazing during daytime. Our data show that grazing on FLB is more intense in dawn-to-dusk in situ incubations than in the 24 h shipboard incubation (Table 1).The disap- pearance of FLB in 24 h incubations (Fig. 4) was also faster during daytime. Whether this is attributable to the existence of a diel cycle in grazing activity or an increase in the abundance of heterotrophic nano- plankton needs further investigation. Based on the diel cycle of growth and grazing of Prochlorococcus and

Table 3. Prochlorococcus and Synechococcus. Total primary production and the production contributed in the upper 175 m water column at Station ALOHA. Total primary production as measured by the 14C uptake method; data courtesy of Tupas TIME (h) et al. (1994).Production values of Prochlorococcus and Syne- chococcus are calculated from growth rate measurements (see text for details). Integrated: from surface to 175 m (Octo- Fig 6. Synechococcus. Changes in density in 24 h shipboard ber 1993 only) (mg C m"2 d"') incubations at Station ALOHA in May 1993. Error bars show 1 SE of the means of triplicate incubations. Control: natural seawater; inhibitor: 1 mg mll kanamycm Depth Carbon production (mg C m3ddl) Total Prochlorococcus Synechococcut

May 1993 The trophic interactions between heterotrophic 5 m nanoplankton and bacterioplankton in the open ocean 45 m are one of the key links of the (Azam et 100 m al. 1983). Heterotrophic flagellates and ciliates are September 1993 thought to be the major grazers of picoplankton, 75 m including heterotrophic bacteria, Prochlorococcus, 125 m Synechococcus and picoeukaryotes (Sieburth 1984). October 1993 Our estimates of grazing pressure on Synechococcus in 5 m Kaneohe Bay (0.22 to 0.39 d-I) and at Station ALOHA 25 m 45 m (0.09 to 0.73 dl)are in close agreement with previous 75 m reports (Landry et al. 1984, Campbell & Carpenter 100 m 1986b, Iturriaga & Mitchell 1986). The estimates of 125 m grazing mortality on Synechococcus ranged from 43 to 150 m 175 m 87 % of growth rate. The grazing rate on Prochlorococ- cus observed outside Kaneohe Bay and in 24 h incuba- Integrated 274 382.2 14.6 tions at Station ALOHA are more variable (20 to Liu et al. Growth and mortality rates of Prochlorococcus and Synechococcus 285

ABUNDANCE (cells ml-') CARBON PRODUCTION (mgC m-3 d-') r 1 o3 1

- 50 -E T fc 100 w n

150 12- Prochlorococcus +Synechococcus -+- picoeukaryotes zoo

GROWTH AND GRAZING RATES (d-l) GROWTH AND GRAZING RATES (d-I )

- -Â¥ grazing - -w- grazing I +growth

Fig. 7.Depth profile of upper 200 m water column at Station ALOHA in October 1993 (a)Vertical distribution of Prochlorococcus, Synechococcus and picoeukaryotes Data are the means of samples at the start point (b)Carbon production at differentdepths; total production was measured by the 14C method (see Tupas et al. 1994) and production of Prochlorococcus and Synechococcus is calculated from growth rates (Table 2) (c)Growth rates and grazing mortalities of Prochlorococcus measured by selective in- hibitor method in dawn-to-dusk in situ incubations followed by dark deck incubations (means SE).(d) Growth and grazing rates of Synechococcus measured in the same experiment

Synechococcus, dawn-to-dusk in situ incubations are not comparable to the total primary production mea- inappropriate for estimating daily growth and grazing sured by ^C method. Our calculation is closer to the rates. Dawn-to-dusk in situ incubations followed by gross production, whereas the 14C method does not shipboard incubations during the dark period gave include carbon uptake that is subsequently ingested reasonable results in our October experiments at Sta- by protozoa and lost to respiration and DOC release. In tion ALOHA. October 1993, the integrated primary production due Prochlorococcus is extremely important at Station to Prochlorococcus in the upper 175 m at Station ALOHA in terms of cell abundance and carbon bio- ALOHA was 382.2 mg C m-2 d-I, which exceeds the mass in comparison to the Atlantic and Indian . integrated total primary production measured by the On the average, Prochlorococcus spp. represent 64 YO 14C method (274 mg C m"2 d-'). Nevertheless, this of the total photosynthetic carbon biomass in the upper result indicates that Prochlorococcus contribute signif- 200 m at Station ALOHA (Campbell et al. 1994). The icantly to primary production in the subtropical North percentages of Prochlorococcus to the total photosyn- Pacific Ocean. Given its relatively low contribution to thetic carbon biomass in the surface mixed layer, deep community biomass (Campbell et al. 1994), Syne- chlorophyll maximum, and < 0.05 % I. (generally 150 chococcus does not significantly contribute to the total to 170 m) are 733,47 % and 33 %, respectively. Our primary production at Station ALOHA (Table 3). data show that Prochlorococcus also contributes signif- The carbodcell conversion factor may also affectthe icantly to primary production at Station ALOHA, but accuracy of the estimation of primary production. The there is temporal variation between cruises (Table 3, conversion factors we used in this paper were based on Fig. 7).We cannot calculate the exact percentages of 0.47 pg C from Verity et al. (1992)which is about Prochlorococcus and Synechococcus contributions to 4 tunes higher than those used in earher studies total primary production because our estimations are (Glover et al. 1988). Also, the size of Prochlorococcus Mar. Ecol. Prog. Ser. 116: 277-287. 1995 - cells is smaller in near-surface waters and increases Campbell, L., Carpenter, E. J. (1986b).Estimating the grazing with depth (Campbell et al. 1994), as does Syne- pressure of heterotrophic nanoplankton on Synechococ- chococcus (Olson et al. 1990b, Burkill et al. 1993). cus spp. using the sea water dilution and selective inhibitor techniques. Mar. Ecol. Prog Ser, 33- 121-129 The contribution of Prochlorococcus to total primary Campbell, L,, Nolla, H. A., Vaulot, D. (1994).The importance production is higher in surface water. At the chloro- of Piochlo~ococcusto community structure in the central phyll maximum layer, which varies between 100 and North Pacific Ocean. Limnol. Oceanogr. 39: 954-961 120 m, the contribution from picoeukaryotes becomes Campbell, L., Vaulot, D (1993).Photosynthetic picoplankton community structure in the subtropical North Pacific more important as shown cell abundance (Fig. 7a) in Ocean near Hawan (Station ALOHA). Deep Sea Res. 40. and biomass (Campbell et al. 1994). Our method is 2043-2060 inappropriate for eukaryotic . Results from dilu- Carpenter, E. J., Chang, J. (1988). Species-specific phyto- tion experiments at Station ALOHA showed that the plankton growth rates via &el DNA synthesis cycles. I. algae grew faster than Prochlorococcus and Syne- Concept of the method. Mar. Ecol. Prog. Ser. 43: 105-111 Chavez, F. P., Buck, K., Coale, K., Martin, J. H , Ditulho, G R., chococcus (data not shown) and they should certainly Welschmeyer, N. A. (1991).Growth rates, grazing, sinking contribute significantly more to primary production and iron limitation of equatorial Pacific phytoplankton. than Synechococcus. Limnol. Oceanogr. 36: 1816-1833 Because Prochlorococcus is a dominant component Chisholm, S. W,, Olson, R. J., Zettler, E. R., Goericke, R., Waterbury, J. B., Welschmeyer, N. A. (1988).A novel free- of the phytoplankton community the tropical and in living prochlorophyte abundant in the oceanic euphotic subtropical open-ocean ecosystems, understanding its zone. 334: 340-343 contribution to photosynthetic biomass and primary Fuhrman, J. A., McManus, G. B. (1984). Do bacteria-sized production in the global oceans is relevant to elucidat- marine consume significant bacterial produc- ing the dynamics of pelagic marine food webs. The tion? Science 224: 1257-1260 Furnas, M. J. (1990). In situ growth rates of marine phyto- implications of these results are important with regard plankton: approaches to measurement, community and to issues such as global ocean flux. We have demon- species growth rates. J. Plankton Res. 12: 1117-1151 strated the effectiveness of kanamycin use in the selec- Gieskes, W. W. C., Kraay, G. W. (1989). Estimating the car- tive inhibitor technique. In using this approach bon-specific growth rate of the major algal species groups m eastern Indonesian waters by "C labeling of taxon- together with flow cytometric enumeration, we esti- specific carotenoid. Deep Sea Res. 36: 1127-1139 mate that the contribution of Prochlorococcus to pri- Glover, H. E. (1985). The physiology and ecology of the mary production in the North Pacific is approximately marine cyanobacterial Synechococcus. Adv. aquat. 1 order of magnitude greater than that of Synecho- Microbiol. 3: 49-107 coccus. Glover, H. E.,Pr6zelin, B. B., Campbell, L., Wyman, M., Gar- side, C. (1988).A nitrate-dependent Synechococcus bloom in surface Sargasso Sea water Nature 331: 161-163 Acknowledgements. We thank D. M. Karl, C. D. Winn, D, V. Goericke, R., Welschmeyer, N. A. (1993). The marine pro- Hebel, R. M Letelier and L. Tupas of JGOFS HOT program chlorophyte Prochlorococcus contributes significantly to for their continued support for our field experiments at Station phytoplankton biomass and primary production in the ALOHA, and H. A. Nolla for assistance in flow cytometric Sargasso Sea. Deep Sea Res. 40: 2283-2294 analyses. We also thank R. R. Bidigare and D. Vaulot for help- Greenwood, D. (1989).Inhibitors of bacterial protein synthe- ful comments on the manuscript. Support for this project is sis. In: Greenwood, D. (ed.) Antimicrobial chemotherapy. from National Science Foundation grants OCE 90-15883 and Oxford University Press, New York, p. 32-45 93-15432. This is contribution no. 3682 from the School of Herbland, A., Le Bouteiller, A,, Raimbault, P. (1985). Size Ocean and Earth Science and Technology, University of structure of phytoplankton biomass in the equatorial Hawaii. Atlantic Ocean. Deep Sea Res. 32: 819-836 Iturriaga, R., Marra, J. (1988).Temporal and spatial variability of chroococcoid cyanobacteria Synechococcus spp. spe- LITERATURE CITED cific growth rates and their contribution to primary pro- duction in the Sargasso Sea. Mar. Ecol. Prog. Ser. 44: Armbrust, E. V., Bowen, J. D., Olson, R. J., Chisholm, S. W. 175-181 (1989). Effect of light on the cell cycle of a marine Syne- Iturriaga, R., Mitchell, B. G. (1986).Chrococcoid cyanobacte- chococcus strain. Appl. environ. Microbiol. 55: 425-432 ria: a significant component in the food web dynamics of Azam, F., Fenchel, T., Field, J. G., Gray, J. S., Meyer-Reil, the open ocean. Mar. Ecol. Prog. Ser. 28: 291-297 L. A,, Thingstad, F. (1983). The ecological role of water- Johnson, P. W., Sieburth, J. M. (1979). Chroococcoid cyano- column microbes in the sea. Mar. Ecol. Prog. Ser. 10: bacteria in the sea: a ubiquitous and diverse phototrophic 257-263 biomass. Limnol. Oceanogr. 24: 928-935 Burkill, P. H., Leakey, R. J. G., Owens, N. J. P., Mantoura, Landry, M. R., Hassett, R. P. (1982). Estimating the grazing R F. C. (1993). Synechococcus and its importance to the impact of marine micro-zooplankton. Mar. Biol 67' microbial foodweb of the northwestern Indian Ocean. 283-288 Deep Sea Res. I1 40: 773-782 Landry, M. R., Haas, L. W., Fagerness, V. L. (1984).Dynamics Campbell, L., Carpenter, E. J. (1986a) Die1 patterns of cell of microbial plankton communities: experiments in Kane- division in marine Synechococcus spp. (Cyanobactena): ohe Bay, Hawaii. Mar. Ecol. Prog. Ser. 16: 127-133 use of the frequency of dividing cells technique to mea- Li, W. K. W,, Subba Rao, D. V., Harnson, W. G., Smith, J. C,, sure growth rate. Mar. Ecol. Prog. Ser. 32: 139-148 Cullen, J. J., Irwin, B., Platt, T. (1983). Autotrophic pico- Liu et al.: Growth and mortahty rates of Prochlorococcus and Synechococcus 287

plankton in the tropical ocean. Science 219 292-295 Thomas, W. H., Dodson, A N (1974) Inhibition of diatom Li, W. K., Wood, A M. (1988). Vertical distribution of North by germanic acid: separation of diatom Atlantic ultraphytoplankton: analysis by flow cytometry productivity from total marine primary productivity Mar. and epifluorescence microscopy Deep Sea Res. 35- Biol 27: 11-19 1615-1638 Tupas, L., Santiago-Mandujano.F, Hebel, D., Lukas, R., Karl, Monger, B. C., Landry, M R. (1993).Flow cytometric analysis D. M., Firing, E (1994) Hawaii Ocean Time-series data of marine bacteria with Hoechst 33342. Appl. environ. report 5. School of Ocean and Earth Science and Technol- Microbiol. 59: 905-911 ogy Tech. Rep., Univ. Hawaii, Honolulu (in press) Newell, S Y,, Sherr, B. F., Sherr, E. B., Fallon, R. D (1983). Vaulot, D (1989). CYTOPC: processing software for flow Bacterial response to the presence of eukaryotic inhibitors cytometric data Signal Noise 2: 8 in water from a coastal marine environment. Mar. environ. Vaulot, D., Courties, C., Partensky, F (1989).A simple method Res. 10: 147-157 to preserve oceamc phytoplankton for flow cytometric Olson, R J , Chisholm, S. W, Zettler, E. R., Altabet, M. A, analysis. Cytometry 10: 629-635 Dusenberry, J. A. (1990a). Spatial and temporal distnbu- Vaulot, D., Marie, D., Partensky, F., Olson, R. J., Chisholm, tions of prochlorophyte picoplankton in the North Atlantic S. W (1994). Prochlorococcus divides once a day at the Ocean Deep Sea Res. 37: 1033-1051 Equator. (Abstract). In. 1994 Ocean Science Meeting Olson, R. J , Chisholm, S. W., Zettler, E. R., Armbrust, E. V EOS, Trans. Am. Geophys. Union 75(3):28 (1990b) Pigments, size, and distnbution of Synechococcus Vaulot, D , Partensky, F., Neveux, J., Mantoura, R. F C., in the North Atlantic and Pacific Oceans. Limnol. Llewellyn, C. (1990). Winter presence of prochlorophytes Oceanogr. 35: 45-58 in surface waters of the northwestern Pelczar, M. J., Reid, D., Chan, E C. S. (1977) Microbiology Limnol Oceanogr. 35- 1156-1164 McGraw Hill, New York Veldhuis, M. J. W, Kraay, G, W., Gieskes, W W. C (1993). Platt, T,Subba Rao, D V., Irwin, B. (1983) Photosynthesis of Growth and fluorescence characteristics of ultraplankton picoplankton in the oligotrophic ocean. Nature 301: on a north-south transect in the eastern North Atlantic. 702-704 Deep Sea Res I1 40: 609-626 Sherr, B F., Sherr, E B , Andrew, T L., Fallon, R D , Newell, Veldhms, M. J. W., Kraay, G. W (1993). Cell abundance and S. Y. (1986) Trophic interactions between heterotrophic fluorescence of picoplankton m relation to growth irradi- protozoa and bacterioplankton in estuarine water ana- ance and nitrogen availability in the Red Sea Neth. J Sea lyzed with selective metabolic inhibitors Mar. Ecol. Prog. Res. 31: 135-145 Ser 32- 169-179 Verity, P G., Robertson, C. Y., Tronzo, C. R , Andrews, M G , Sherr, B F., Sherr, E. B , Fallon, R. D. (1987).Use of monodis- Nelson, J. R., Sieracki, M. E. (1992). Relationships be- persed, fluorescently labeled bacteria to estimate in situ tween cell volume and the carbon and nitrogen content of protozoan bacterivory Appl. environ. Microbiol 53: marine photosynthetic nanoplankton. Limnol. Oceanogr. 958-965 37: 1434-1446 Sieburth, J. McN. (1984). Protozoan bacterivory in pelagic Waterbury, J. B., Watson, S. W., Guillard, R. R. L., Brand, L. E. marine waters. In: Hobble, J. E., Williams, P. J. leB. (eds.) (1979). Widespread occurrence of a unicellular, marine, Heterotrophic activity m the sea. Plenum Press, New York, planktonic cyanobacterium. Nature 277: 293-294 p. 405-444 Waterbury, J B , Watson, S. W., Valois, F W., Franks, D. G. Takahashi, M., Bienfang, P. K. (1983).Size structure of phyto- (1986). Biological and ecological characterization of the plankton biomass and photosynthesis in subtropical marine unicellular cyanobactena Synechococcus. Can. Hawaiian waters. Mar Biol. 76. 203-211 Bull Fish. Aquat. Sci. 214: 71-120

This article was submitted to the editor Manuscript first received: April 27, 1994 Revised version accepted-September 29, 1994