Global Change Biology (2006) 12, 1595–1607, doi: 10.1111/j.1365-2486.2006.01181.x

Variation in groenlandicus () growth in a Norwegian high-Arctic fjord: evidence for local- and large-scale climatic forcing

WILLIAM G. AMBROSE Jr.*w , MICHAEL L. CARROLLw , MICHAEL GREENACREz, SIMON R. THORROLD§ andKELTON W. McMAHON*§ *Bates College, Department of Biology, Lewiston, ME 04240, USA, wAkvaplan-niva, Polar Environmental Centre, N9296 Troms, Norway, zPompeu Fabra University, 08005 Barcelona, Spain, §Woods Hole Oceanographic Institution, Biology Department MS 50, Woods Hole, MA 02543, USA

Abstract We examined the growth rate of the circumpolar Greenland (Serripes groenlandi- cus) over a period of 20 years (1983–2002) from Rijpfjord, a high-Arctic fjord in northeast Svalbard (801100N, 221150E). This period encompassed different phases of large-scale climatic oscillations with accompanying variations in local physical variables (tempera- ture, atmospheric pressure, precipitation, sea ice cover), allowing us to analyze the linkage between growth rate, climatic oscillations, and their local physical and biological manifestations. Standard growth index (SGI), an ontogenetically adjusted measure of annual growth, ranged from a low of 0.27 in 2002 up to 2.46 in 1996. Interannual variation in growth corresponded to the Arctic climate regime index (ACRI), with high growth rates during the positive ACRI phase characterized by cyclonic ocean circulation and a warmer and wetter climate. Growth rates were influenced by local manifestations of the ACRI: positively correlated with precipitation and to a lesser extent negatively correlated with atmospheric pressure. A multiple regression model explains 65% of the variability in growth rate by the ACRI and precipitation at the nearest meteorological station. There were, however, complexities in the relationship between growth and physical variables, including an apparent 1 year lag between physical forcing changes and biological response. Also, when the last 4 years of poor growth are excluded, there is a very strong negative correlation with ice cover on a pan-arctic scale. Our results suggest that bivalves, as sentinels of climate change on multi-decadal scales, are sensitive to environmental variations associated with large-scale changes in climate, but that the effects will be determined by changes in environmental parameters regulating marine production and food availability on a local scale. Key words: Arctic, arctic climate regime index, benthic community, benthos, bivalve growth, climate forcing, , Svalbard

Received 6 July 2005; revised version received 21 December 2005; accepted 3 March 2006

panied by changes in terrestrial and marine ecosystems Introduction (Oechel & Vourlitis, 1997; Morison et al., 2000; Serreze The Arctic climate has changed dramatically in the last et al., 2000). Effects of persistent climate change on several decades (Maxwell, 1997; Overpeck et al., 1997; Arctic marine ecosystems are largely undetermined, Johannessen et al., 2003, 2004; AICA, 2004). The average but changes that occur in response to decadal-scale annual air temperature has increased by 11 to 4 1C in the climate oscillations may provide insight into longer- last half century (AICA, 2004), and this has been accom- term effects of more persistent climate change. Several large-scale climate oscillations have been Correspondence: W. G. Ambrose, Jr., Bates College, Department of shown to influence marine systems (see Allan et al., Biology, Lewiston, ME 04240, USA, tel. 1 1 207 786 6114, 1996; Ottersen et al., 2001; Walther et al., 2002; Stenseth fax 1 1 207 786 8334, e-mail: [email protected] et al., 2003 for reviews). Linkages between the two r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd 1595 1596 W. G. AMBROSE et al. climate oscillations with nodes centered in the Arctic, latitude bivalves have life spans of decades (Tallqvist & the Arctic oscillation (AO) and the Arctic climate regime Sundet,2000;Mu¨ller-Lupp et al., 2003; Sejr & Christensen, index (ACRI), and the marine ecosystem, however, have 2006) to well over 100 years (Turekian et al., 1975; not been demonstrated. Both indices reflect differences Thompson et al., 1980; Zolotarev, 1980; Peck & Bullough, in wind-driven motion in the central Arctic alternating 1993; Witbaard et al., 1999; Sejr et al., 2004). Bivalves can, between two phases, an anticyclonic circulation regime thus, serve as bioproxies by providing uninterrupted (ACCR) and a cyclonic circulation regime (CCR). The records of environmental conditions over decades to climate regimes manifest as physical variables in the centuries, which is critical in the Arctic given the paucity Arctic; ACCR is characterized by a cold and dry high- of long-term data on community structure and dynamics. Arctic atmosphere and a colder and saltier polar ocean We examined interannual variation in growth of the (low AO and negative ACRI), whereas the cyclonic circumpolar Greenland Cockle (Serripes groenlandicus) regime is characterized by a warm and wet atmosphere from 1983 to 2002 in a high-Arctic fjord in northeast and a warm and fresh polar ocean (high AO, positive Svalbard (Norwegian Arctic) to explore the relationship ACRI). Climatic conditions associated with both the AO between benthic communities and environmental var- and ACRI may affect marine ecosystems as has been iations associated with decadal climate oscillations in demonstrated for the North Atlantic oscillation (NAO) the Arctic. Variation in bivalve growth associated with (Ottersen et al., 2001). changes in environmental conditions that occur over the Seafloor communities may be the best location to course of a decadal-scale oscillation cycle provides in- examine the impact of Arctic climate oscillations, and sight into the response of a dominant member of the by extension the potential effects of climate change on Arctic benthos to predicted long-term climate change. the Arctic ecosystem. There is often a close relationship between water column and benthic processes (Grebme- Materials and methods ier et al., 1988; Ambrose & Renaud, 1995; Piepenburg et al., 1997; Wollenburg & Kuhnt, 2000; Dunton et al., Study site 2005), therefore, long lived, sessile benthic organisms, may be more appropriate monitors of climate change Rijpfjord (801100N, 221150E) is located on the north- (e.g. Kro¨ncke et al., 1998, 2001; Dunton et al., 2005) than central shore of Nordaustlandet (Fig. 1), north and east the more transient pelagic system. Additionally, benthic of Spitsbergen in the Svalbard Archipelago. Rijpfjord is communities are key components in the carbon cycle on oriented south–north and opens to a broad shallow Arctic shelves (Grebmeier et al., 1989; Grant et al., 2002; shelf of approximately 200 m depth extending to the Stein & Macdonald, 2004; Clough et al., 2005) and food shelf-break of the Polar Basin at roughly 811N. The for higher trophic levels (e.g. bottom feeding fish, bottom depth averages 200–250 m, but an irregular sill mammals, and birds; Dayton, 1990). Consequently, crosses the width of the fjord midway through its changes to the benthos may have profound effects on length. Shallower depths are dominated by bedrock carbon cycling, trophic structure, and food web dy- and stones covered with a thin layer of mud, while soft namics on Arctic shelves. sediments predominate deeper sections. Bivalves comprise a significant proportion of the Rijpfjord is a true polar fjord. It is predominately ice- benthic biomass of Arctic shelves (Zenkevich, 1963; covered for at least 9 months a year (October–June), McDonald et al., 1981; Gulliksen et al., 1985; Grebmeier with breakup occurring between mid-July and mid- et al., 1988; Dayton, 1990; Feder et al., 1994). The shells of August. Even during the summer period, winds often most bivalves exhibit periodic banding, or growth lines force drifting pack ice into the fjord. The shallow outlet (Rhoads & Pannella, 1970; Clark, 1974; Rhoads & Lutz, of the Rijpfjord shelf to the Polar Ocean results in little 1980), that have proved valuable in developing a history warm Atlantic subsurface water entering the fjord from of environmental change in marine systems (Andrews, the north (A. Sundfjord, unpublished data). The upper 1972; Hudson et al., 1976; Jones, 1981; Jones et al., 1989; 10 m is a mixed layer with surface water of lower Witbaard, 1996; Witbaard et al., 1997, 1999; Tallqvist & salinity and higher temperature associated with sea Sundet, 2000; Scho¨ne et al., 2003; Mu¨ller-Lupp & Bauch, ice melt and in situ radiative warming; below which is 2005). Temperature and food are the two main factors a water mass with a temperature consistently o1 1C influencing bivalve growth (Beukema et al., 1985; Jones and a salinity 433% (Fig. 2). et al., 1989; Beukema & Cade´e, 1991; Lewis & Cerrato, 1997; Witbaard et al., 1997, 1999; Dekker & Beukema, 1999; Bivalve collection Scho¨ne et al., 2005), and both are likely to be influenced by climate change in Arctic marine systems (Carroll & Divers collected four Serripes groenlandicus individuals Carroll, 2003). Furthermore, many deep water and high from 17 m depth on 28 August 2003. Individuals were

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 BIVALVE GROWTH AND ARCTIC CLIMATE 1597

0

50 Temperature

100 Salinity Depth (m)

150

200 –2–10123 Temperature (°C) 28 30 32 34 36 Salinity (‰)

Fig. 2 Temperature and salinity profile from Inner Rijpfjord, near the location of bivalve collection, on 28 August 2003.

Fig. 1 Map of the study region showing the Spitsbergen Archi- pelago, general current patterns, the locations of the meteorolo- gical stations (BI, Bear Island, HO, Hopen, LB, Longyearbyen, a Thermo Finnigan Element2 single collector sector field NA˚ ,NyA˚ lesund), and the collection site of Serripes groenlandicus ICP-MS. The valves were mounted in resin and a 1.0 mm (Rijpfjord, RF). The polar front is the average location of max- cross-section was cut along the axis of maximum imum ice cover in late winter. growth using a low-speed saw equipped with a dia- mond blade. The section was mounted on a petro- graphic slide with cyanoacrylic glue, polished using identified by their pale, white siphon, and were dug 30 and 3 mmAl2O3 lapping film, and then decontami- from the sediment where they were buried to a depth of nated in a clean room. Each valve was scrubbed with a

5 cm. All were alive at the time of collection. nylon brush, triple rinsing with 2% HNO3, sonicated for Soft tissues were immediately separated from the shells, 5 min in ultra-pure H2O, triple-rinsed in water, and then and the shells air dried. dried for at least 24 h under a laminar flow clean bench. Instrument set-up was similar to that outlined by Gu¨ nther & Heinrich (1999) as modified by Thorrold Strontium/calcium ratios et al. (2001). Linear spatial resolution was 100 mm along S. groenlandicus deposit distinctive lines during growth the growth axis which corresponded to weekly resolu- which appear as alternating thin dark and thicker light tion during early years of life, declining to monthly bands on the external shell surface. We used changes in resolution as shell growth rates decreased at older ages. the ratio of Strontium (Sr) to Calcium (Ca) in different Quantification of Sr/Ca ratios followed the approach areas of the shell to identify whether external lines were outlined by Rosenthal et al. (1999) for precise element/ deposited annually. The Sr/Ca ratio in carbonate varies Ca ratios using sector field ICP-MS. Average external with the water temperature at the time of deposition precision (RSD) of Sr/Ca ratios in the laboratory stan- (Zacherl et al., 2003; Bath et al., 2004), so systematic dard was 0.2%. changes between presumptive growth lines, reflecting seasonal changes in water temperature, would suggest Growth rates the lines are deposited annually. We measured Sr/Ca ratios along the growth axis within the prismatic layer The external growth lines on S. groenlandicus are annual of two S. groenlandicus shells (N1 and N3) using a New growth checks (Khim et al., 2003; this study – see Sr/Ca Wave Research UP213 laser ablation system coupled to ratio results), thus, can be used to determine growth r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 1598 W. G. AMBROSE et al. rates. Since we collected all clams live, each growth 1997; Johnson et al., 1999; A. Proshutinsky, personal com- increment can be assigned to a specific calendar year. munication for updated index). The sea surface height The distance between the ventral edges of successive anomaly at the North Pole is indicative of the predo- growth lines along the line of maximum growth (shell minant high-Arctic wind pattern. Data for the NAO, height) was measured with a digital caliper to the using the sea level pressure difference between Gibral- nearest 0.01 cm. We do not include growth beyond the tar and Southwest Iceland (Jones et al., 1997; Osborn last growth line in our analyses because we do not et al., 1999), were obtained from (http://www.cru.uea. know what portion of the present year this growth ac.uk/cru/data/nao.html) and for the AO from (http:// represents. www.cpc.ncep.noaa.gov/products/precip/CWlink/ Shell growth of S. groenlandicus was modeled by daily_ao_index/ao_index.html). fitting the von Bertalanffy growth function to age and The Barents Sea temperature data (PINRO, Mur- shell height data for each clam using prism (Ver. 3.0, mansk) is a time series of integrated ocean temperature 1993). from 0 to 200 m depth along the Kola transect, which Bivalve growth declines with age, so growth incre- runs from the Kola Peninsula northward to the ice ments within an individual and among individuals of edge along the 331300E meridian (Bochkov, 1982; different ages must be standardized before growth Tereshchenko, 1997). We have used yearly means of among years can be compared. We use the methods of Barents Sea temperature in our analysis. Jones et al. (1989), employing the first derivative of the Meteorological data were obtained from the four von Bertalanffy function with respect to time, to derive official weather stations around Svalbard (Longyear- an ontogenetically adjusted measure of annual growth: byen, Ny A˚ lesund, Bear Island, and Hopen; Fig. 1) kt maintained by the Norwegian Meteorological Institute dSHt=dt ¼ kSH1e ; (http://eklima.met.no). Daily data of precipitation, where SHt the shell height at age t, dSHt/dt the mod- pressure, and temperature were used to calculate sea- eled yearly change in shell height, t the age in years, sonal and yearly averages. SH1 the maximum, asymptotic shell height, k the Local ice conditions were estimated from imagery by growth constant. the Nimbus-7 SMMR and DMSP SSM/I passive micro- After determining the average yearly changes in shell wave satellite (Cavalieri et al., 1997). The spatial resolu- height based on growth data from all clams, we calcu- tion of the satellite imagery is 25 km 25 km, and the lated the expected yearly increase in shell height for cell used for the ice analysis is immediately northward each clam for each year. We then divided the measured of Rijpfjord proper. The temporal resolution is daily or observed shell growth for each year by the expected from 1988 to 2003 and every second day from 1978 to growth for that year to generate a standardized growth 1988. We calculated the ice-free days per year as the index (SGI). This removes the ontogenetic changes in number of days with ice cover o25% from 1 July to 30 growth and equalizes the variance for the entire series October of a given year (the estimated period of active (Fritts, 1976). Once annual changes in shell growth were growth of S. groenlandicus). This measure relates closely standardized, we calculated the mean SGI for each (490%) to ice free days in a calendar year and average calendar year (because growth lines are deposited yearly ice concentration. Data on total Arctic-wide in winter, growth year is virtually synonymous with spatial extent (km2) of pack ice were obtained from the calendar year). The result is a record of year-by-year US National Snow and Ice Data Center (http://www. growth for the S. groenlandicus individual, with an SGI nsidc.com/data/seaice_index/) (Fetterer & Knowles, greater than one indicating a better than average year 2002). for growth, while a value less than one reflects a worse than average growth year. Statistical analyses We calculated Pearson’ correlation coefficients in order Climatic and meteorological data to determine basic pairwise relationships between SGI We examined relationships between clam growth and and the environmental and physical variables. We com- climate indices with potential influence on the region: pared annual means and also investigated the time the ACRI, NAO and AO. The AO is the first principal dependence between data in consecutive years, leading component of the sea level pressure field at latitudes us to incorporate two data transformations: 2 year- 4201N (Thompson & Wallace, 1998; Stenseth et al., running means of environmental data were used to 2003), while the ACRI measures variations in Arctic reduce the magnitude of interannual variability of these Ocean and ice circulation based on the sea level height data, and a 1-year lag was used to account for the time anomaly at the North Pole (Proshutinsky & Johnson, for physical processes to be reflected in shell growth.

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 BIVALVE GROWTH AND ARCTIC CLIMATE 1599

We used a general linear model to investigate the data 5 Clam N1 at an individual clam level, firstly to identify predictors of S. groenlandicus growth, and secondly to test whether there were significant differences between the clams, 4 included in the model as a random effect. To further justify pooling the growth rates of the sample of clams, we computed the Cronbach a measure of reliability (e.g. 3 Bland & Altman, 1997), for the available growth data on a common set of years available for all clams (1992– )

1 2

2002) to examine the homogeneity of the four clam − growth rates. Once we established that the clams were a homogenous sample, we combined the clam data into an annual mean value as a more reliable indicator of Clam N3 annual clam growth rate. This mean was then used as a response variable in a multiple linear regression model, Sr/Ca (mmol mol 4 investigating relationships with all available annual predictor variables. In all analyses, we checked for serial correlation in the residuals, which should be 3 absent to satisfy the assumptions of the regression model. Statistical analyses were done using Statistica (ver. 6) or SPSS (ver. 11.5.1). 2

Results 1 The clams collected were 20, 19, 18, and 11 years old. We 0 100 200 300 400 assayed Sr/Ca ratios from 440 samples spanning 16 Sample number (consecutive from umbo growth increments, and 350 samples spanning 12 incre- toward shell margin) ments, in the 20-year-old (N1) and 19-year-old (N3) Fig. 3 Ratio of strontium to calcium in two Serripes groenlandi- shells, respectively. The two Sr/Ca profiles were similar, cus shells (N1 and N3) from Rijpfjord. Samples were taken along with maximum values of between 3.5 and 4.5 mmol each clam’s height and are numbered from the umbo (sample mol1 in aragonite deposited during early life, and a number 0) to the ventral margin. Triangles correspond to the minimum value of approximately 1.5 mmol mol1 re- location of dark growth lines visible on a shell’s exterior. corded as the individuals grew older (Fig. 3). There was clear periodicity in Sr/Ca ratios that corresponded to growth increments visible in the shells. The amplitude appear to follow a cyclic pattern with better growth in of Sr/Ca variation within an increment ranged from 2 the mid-1990s bracketed by two poorer growth periods to 2.5 mmol mol1 in wide increments during the first in the late-1980s and the early part of this century. 2–3 years of life to 1–1.5 mmol mol1 as growth lines Growth has declined steadily for the last 4 years of narrowed with age. Sharp declines in Sr/Ca ratios were the series, with the worst growth in the last 2 years invariably associated with the presence of a dark (2001, 2002). growth line, confirming the growth lines are winter To justify combining the growth rates for the indivi- growth checks. dual clams into an average Cronbach’s reliability coeffi- The von Bertalanffy growth equation (SHt ¼ 112:8 cient was calculated for the years 1992 until 2002, for ð1 e0:032ðtt0ÞÞ had an R2 of 0.956 (P 0), indicating which we had data from all four clams. The value that it was an excellent descriptor of these clams’ a 5 0.663 indicates a reasonably high level of reliability. growth. This allowed us to correct for changes in Unless mentioned otherwise, the following results are growth with age and confidently generate the expected for the average annual SGI values. growth for each calendar year SGI. SGI is most highly correlated with precipitation on SGI varied considerably over the 20 years of the data Hopen and the ACRI (Table 1). These positive correla- set (Fig. 4), with a range for individual clams between tions generally improve when running means are used 0.27 in the poorest growth year (N1 in 2002) to 2.46 in instead of annual means for the environmental para- the best growth year (N2 in 1996). The total population meter and when the environmental data are lagged SGI ranged from 0.45 (2002) to 1.64 (1995). The differ- 1 year with respect to growth, growth corresponding ences in SGIs among years are clearly not random, but better to the previous year’s environmental data than r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 1600 W. G. AMBROSE et al. the current year’s. The ACRI was the only large-scale Considering all explanatory variables and their lags climate index related to clam growth (Fig. 5) and this as well as their running means as possible predictors, correlation was very significant with a 1-year lag and the best subset of predictor variables for the SGI which using the running mean (r 5 0.671, Po0.001). Atmo- we identified are the ACRI averaged over the 2 years spheric pressure at Hopen and Longyearbyen are also before SGI measurement (running mean with a 1-year significantly correlated, in these cases negatively, with lag) and the running mean (present and previous year) the SGI when a 1-year lag and the running mean of the of precipitation at Ny A˚ lesund. The model is:

Average SGI ¼ 1:715 þ 0:319 ACRI 0:00183 Ny Alesund precipitation ð2 year mean; ð2 year meanÞ 1 -year lagÞ ðP < 0:00001ÞðP < 0:007Þ environmental data are applied. Removal of the varia- This model explains nearly 65% of the variability in bility explained by the ACRI and a correlation analysis the mean SGI (R2 5 0.649). Although the individual between the residuals and the SGI revealed significant variables are serially correlated, the residuals in the negative correlations between the SGI and the Kola Sea model are not autocorrelated (Durbin-Watson test sta- temperature, precipitation at Ny A˚ lesund and air tem- tistic 5 1.67), thus satisfying the assumption of the perature in Longyearbyen. The marginally significant regression model. Figure 6 shows the observed mean correlations in Table 1 should be interpreted with some SGI plotted against the value predicted by our model, caution, because of multiple testing. In order to avoid where the 451 line indicates perfect prediction. This the problem of higher Type I error incurred by testing shows that the model-generated SGI’s closely track several correlations, we should lower the significance the measured values. The only significant residual is level below the customary value of 0.05 (Bland & Alt- for the year 1991, which is over-predicted because of a man, 1995). However, the exact threshold necessary for 1-year drop in average SGI in an otherwise multi-year each individual correlation for an overall significance uptrend in SGI (Fig. 5). In other words, the serial level of 0.05 is difficult to determine because the corre- autocorrelation breaks down in 1991. lations in each row are clearly related. Discussion

2.2 Growth of S. groenlandicus from northeast Svalbard is clearly related to the ACRI, a large-scale arctic climate 2.0 oscillation dependent on high-Arctic atmospheric 1.8 circulation patterns (Proshutinsky & Johnson, 1997; 1.6 Johnson et al., 1999; Polyakov et al., 1999), through 1.4 1.2 2 1.0 1.6

0.8 1 1.4

Standard growth index 0.6 1.2 0.4 n=4 0 1.0 rowth index rowth

0.2 g 1985 1990 1995 2000 0.8 Year –1 0.6 (2 yr running mean) Climate regime index regimeClimate index Fig. 4 Standard Growth Index (SGI) (1 standard error of the Standard mean) of the Serripes groenlandicus from Rijpfjord, Svalbard (Nor- 0.4 –2 way) collected live on 28 August 2003. Growth increments were 1980 1985 1990 1995 2000 calculated using external measurements and the von Bertalanffy Year equation to remove ontogenetic changes in growth. An SGI greater than 1.0 indicates greater than average growth while Fig. 5 Temporal patterns of the Arctic climate regime index one less than 1.0 poorer than average. Growth was measured (ACRI) (gray bars), lagged 1 year, and Serripes groenlandicus SGI along the line indicated on the shell. The dark bands represent- (line). The ACRI data are 2-year running means of raw data, ing winter cessation of growth are visible on the shell. while the SGI data are untransformed.

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 BIVALVE GROWTH AND ARCTIC CLIMATE 1601

Table 1 Pearson Correlations relating Serripes groenlandicus standard growth index (SGI) to various environmental variables, 1983–2002

Correlations with average SGI Correlations with Present Year 1-year lag residuals: 1-year lag

Running Running Running Environmental variable Annual mean Annual mean Annual mean

Kola Transect sea temperature 0.032 0.008 0.045 0.033 0.345 0.462* Summer ice free days in Rijpfjord 0.166 0.107 0.005 0.050 0.132 0.145 Pack ice extent March 0.315 0.232 0.319 0.137 0.416 0.422 Ny A˚ lesund air temperature 0.020 0.141 0.204 0.263 0.409 0.318 Ny A˚ lesund pressure 0.169 0.349 0.381 0.428 0.044 0.202 Ny A˚ lesund precipitation 0.212 0.257 0.158 0.134 0.509* 0.250 Longyearbyen air temperature 0.002 0.112 0.181 0.112 0.434* 0.341 Longyearbyen pressure 0.198 0.383 0.408 0.455* 0.023 0.177 Longyearbyen precipitation 0.047 0.181 0.199 0.389 0.095 0.047 Hopen air temperature 0.015 0.114 0.175 0.199 0.427 0.362 Hopen pressure 0.216 0.417 0.432 0.451* 0.055 0.150 Hopen precipitation 0.476* 0.592** 0.592** 0.515* 0.341 0.244 Bear island air temperature 0.124 0.167 0.152 0.196 0.386 0.354 Bear island pressure 0.221 0.374 0.366 0.364 0.038 0.175 Bear island precipitation 0.323 0.408 0.319 0.213 0.255 0.037 NAO index (Annual) 0.025 0.106 0.140 0.209 0.162 0.148 NAO index (Winter) 0.178 0.297 0.241 0.219 0.081 0.229 AO index (Annual) 0.021 0.129 0.233 0.332 0.175 0.219 AO index (Winter) 0.242 0.358 0.322 0.368 0.124 0.223 Arctic Climate Regime Index 0.297 0.522* 0.492* 0.671*** 0.018 0.000

The first set of two columns of coefficients are from annual data (temperature and pressure data used are annual means, precipitation is annual total sum) and 2-year running means of the annual data. The second set is from the environmental data being set back (lagged) by 1 year with respect to the growth data. The third set contains the correlations (1-year lag only) after the effects of the Arctic Climate Regime Index (2-year running mean, 1-year lag) have been removed. Significant correlations are shown in bold. The levels of the significant correlations are: *Po0.05, **Po0.01, ***Po0.001.

local environmental conditions. Fortunately, the S. groen- 2003), or more rarely in polar environments, mark and landicus growth we examined spanned large changes in recapture studies (Sejr et al., 2002), they have proved to the ACRI (from positive to negative phase and vice be annual and are associated with a winter cessation of versa), offering us the opportunity to examine the growth. In S. groenlandicus from the Chukchi Sea, d18O response of this important member of the benthic com- values vary systematically between growth lines with munity to large changes in climate. The response of S. the highest values, reflecting the coldest temperatures, groenlandicus to the ACRI and local conditions, however, coincident with the dark lines on the shell, strongly was complex, including an apparent 1-year lag between suggesting the lines are deposited annually during physical forcing changes and biological response. winter (Khim et al., 2003). The use of bivalves as bioproxies for climate changes We found that Sr/Ca profiles also varied system- hinges on the ability to recognize annual markers in atically between individual growth lines, with mini- their shell. While it is generally assumed that lines mum values always associated with dark growth visible on the external shell or internally in cross section bands. Recent culturing studies have found that the are deposited annually, particularly at high latitudes, Sr/Ca ratio is positively correlated with temperature in this assumption is not always tested (Andrews, 1972; low-Sr aragonite from mollusk shells and fish otoliths Tallqvist & Sundet, 2000). When the periodicity of (Zacherl et al., 2003; Bath et al., 2004), meaning that increments has been examined using changes in oxygen growth lines in S. groenlandicus from Rijpfjord are likely and carbon isotopes between presumptive annual lines deposited during the winter. The within-year ampli- (Witbaard et al., 1994; Heilmayer et al., 2003; Khim et al., tudes of Sr/Ca variations we found are on the order of r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 1602 W. G. AMBROSE et al.

2.0 1.8

1.6 95 1.4 1.5 96 92 93 1.2 98 1.0 97 9484 83 1.0 89 90 0.8 85 99 Standard growth index 00 91 y = –0.858x + 14.586 88 0.6 SGI (observed) 86 P < 0.0001, R2=0.53 0.4 0.5 8701 02 15.0 15.2 15.4 15.6 15.8 16.0 16.2 Total Arctic ice pack extent in March (106 km2)

Fig. 7 Relationship between total Arctic-wide winter (March) 0.0 sea ice extent (2-year running mean) and Serripes groenlandicus 0.0 0.5 1.0 1.5 2.0 SGI from 1983 to 1998. (Ice data courtesy of National Snow and SGI (modeled) Ice Data Center, Boulder, CO, USA).

Fig. 6 Scatterplot of observed vs. modeled SGI from 1983 to 2002 with the growth years shown as data points. Line repre- sents a perfect prediction by the model of actual growth rates of with Arctic climate. Once the variability in the SGI the Serripes groenlandicus population. attributed to the ACRI is removed, either in the multiple regression model or in the correlation analysis, precipi- tation at Ny A˚ lesund is related to the SGI. 50–100%, and are of a similar magnitude to those The polar-scale effect of climate on S. groenlandicus reported by Stecher et al. (1996) for the bivalve Merce- growth, however, is not immediate or direct because naria mercenaria. Thus, the results are clearly consistent variation in S. groenlandicus growth is best explained by with an annual deposition rate of growth lines in the climate index one year earlier than the current S. groenlandicus shells and justify use of external lines years’ growth (Table 1, regression model). Lagged re- on the shells as annual markers. sponse to climate oscillations are common in marine Considerable effort has recently been directed to- systems and can typically span many trophic levels wards understanding the relationships between climate (Post, 2004) from benthic infauna (Tunberg & Nelson, variation and ecosystem structure and function. Cli- 1998), including the bivalve Arctica islandica (Witbaard mate oscillations and the associated changes in physical et al., 2003), to zooplankton (Pershing et al., 2004), fish parameters such as temperature, water circulation, and (Ottersen et al., 2004) and birds (Thompson & Ollason, ice cover have been shown to measurably influence 2001). In the case of S. groenlandicus, the lag is probably marine ecosystems in the North Pacific, North Atlantic, because the local manifestations of the climate oscilla- and Southern Ocean by regulating the abundances of tion (i.e. precipitation) take a period of time to develop. organisms at the base and upper levels of food webs Furthermore, there is likely an additional time lag (Fomentin & Planque, 1996; Stabeno & Overland, 2001; associated with the biophysical coupling via physical Hunt & Stabeno, 2002). S. groenlandicus growth is not constraints on food production, biological processes strongly related to the NAO, the AO, which is itself of consumption and assimilation as tissue and shell related to the NAO (Deser, 2000; Hurrell et al., 2003), or growth, and storage of energy as tissue from previous the Barents Sea temperature (until after the effects of the years. ACRI are removed) (Table 1), which is influenced by Surprisingly, given the importance of ice in mediating incursions of the North Atlantic Current into the environmental conditions in the Arctic, no measure of Barents Sea (Loeng, 1991), and only correlated with ice condition (summer ice free days, total arctic ice pack one of the environmental variables from locations out- extent) was related to S. groenlandicus growth (Table 1). side the polar front (e.g. all locations except Hopen), When the last 4 years of poor growth are removed from atmospheric pressure at Longyearbyen. This pattern of the analysis, however, there is a very strong, significant relationships indicates that we are measuring the negative relationship between SGI and Arctic-wide response of the marine ecosystem to an exclusively extent of the pack ice (Fig. 7). It is remarkable that such Arctic phenomenon, rather than a temperate interaction a large-scale measure of ice conditions as the extent of

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 BIVALVE GROWTH AND ARCTIC CLIMATE 1603

Arctic climatic Regional jord is also food-rather than temperature-limited, pro- regime index atmospheric cesses regulating the quantity and quality of food pressure reaching the bottom of the fjord will have a strong effect on S. groenlandicus growth. In the absence of a + 1 year strong temperature signal, variation in food quantity, Local precipitation and possibly quality, is the most probable explanation Biophysical Bivalve for interannual differences in S. groenlandicus growth coupling growth in Rijpfjord. This is the same conclusion reached by Sea ice concentration Witbaard et al. (2003) for interannual variations in growth of Arctica islandica in the North Sea, and Sejr et al. (2004) for Hiatella arctica in east Greenland. Ecosystem ˚ response Precipitation at Ny Alesund is the only environmen- tal variable that enters the multiple linear regression model and it is negatively related to the SGI. Precipita- Fig. 8 Schematic representation of the Arctic Climate Cascade, tion at Hopen is positively correlated with growth in where bivalve growth, and by inference, other ecosystem pro- current and previous years (Table 1) (precipitation at Ny cesses are linked to climatic forcing factors through variation in ˚ physical variables. Alesund and Hopen are negatively correlated, though not significantly, hence their opposite relationships with the SGI). The effect of precipitation on S. groenlandicus total pack ice across the Arctic explains over 50% of the growth could be due to direct or indirect effects and we interannual variability in the growth of S. groenlandicus. have little data from Rijpfjord to help us interpret these The lack of relationship between growth and Arctic- relationships. The land around Rijpfjord is heavily wide ice conditions when the last 4 years are included glaciated and the un-glaciated areas are sparsely vege- in the analysis suggests a decoupling of the previously tated with nutrient-poor soil, so spring runoff is un- strong relationship in the last 4 years. likely to be laden with nutrients necessary to stimulate a The growth of S. groenlandicus is linked to the ACRI phytoplankton bloom. Precipitation might also stabilize though the impact of the climate oscillation on the local the water column, which has been shown to initiate a physical conditions of precipitation and ice cover (Fig. spring bloom in the Bering Sea (Stabeno & Overland, 8). The integration of water column processes by long- 2001) and west Greenland (Nielsen & Hansen, 1995), lived Arctic benthos means that the effect of climate on but is not necessary for a spring bloom to develop in the S. groenlandicus growth may be direct or indirect, and Rijpfjord system (Hegseth et al., 1995). Precipitation the specific mechanisms of coupling between physical could affect the salinity of surface water and the abun- conditions and bivalve growth in Rijpfjord (Fig. 8) are dance of herbivorous zooplankton, which are capable in not well understood. Variation in bivalve growth is Arctic fjords of consuming over 80% of annual primary typically best explained by variation in temperature production (Nielsen & Hansen, 1995) and indirectly and food (Beukema et al., 1985; Jones et al., 1989; influencing bivalve growth (Witbaard et al., 2003); Beukema & Cade´e, 1991; Lewis & Cerrato, 1997; Wit- a top-down rather than a bottom-up effect. It is also baard et al., 1997, 1999; Dekker & Beukema, 1999). Some possible that precipitation is a reflection of storms, of the best documented effects of climate oscillations on which may cause resuspension of settled phytodetritus, individuals, populations and benthic community struc- and in shallow enough water, benthic microalgae, both ture are mediated through temperature (Kro¨ncke et al., of which could be consumed by S. groenlandicus,a 1998; Ottersen et al., 2001; Hagberg et al., 2004). In late positive relationship, or excessive wind might suspend August when the clams in Rijpfjord were collected, the bottom sediment clogging the gills of S. groenlandicus thermocline roughly coincided with collection depth and resulting in lowered growth, a negative relation- (Fig. 2), indicating that the clams lived deep enough ship. Without more information on environmental to experience relatively small differences in tempera- conditions in Rijpfjord and how they affect primary ture over the year, meaning the temperature effects on productivity and delivery of food to the benthos, we can growth will be limited and likely minor compared to do no more than speculate on the relationships between other factors controlling food availability. growth of S. groenlandicus and precipitation. The growth of many polar benthic organisms appears Perhaps the most intriguing aspect of the relation- to be food limited (Brockington & Clarke, 2001) and ships between S. groenlandicus growth and environmen- differences in growth among sites have been related to tal conditions is the decoupling that occurred during differences in food supply (Brey et al., 1995; Norkko the last 4 years between ice cover and growth. Annual et al., 2005). If the S. groenlandicus population in Rijpf- phytoplankton production in the Arctic is directly pro- r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 1604 W. G. AMBROSE et al. portional to the length of open water (Rysgaard et al., S. groenlandicus and other bottom dwelling organisms to 1999) and benthic biomass on Arctic shelves is inversely climate fluctuations, we need to better understand related to ice cover (Ambrose & Renaud, 1995). In variation in the quality and quantity of food reaching Rijpfjord, however, 3 weeks after the break-up of ice the sea floor. in 2003 during the period when the ‘spring’ bloom would be expected, there was a very shallow mixed- layer depth (Fig. 2) and little suspended chlorophyll Acknowledgements biomass in the water column (maximum 1 mgL1,E.N. Hegseth, unpublished data). Locally produced phyto- This research was supported in part by the Norwegian Research Council, NORDKLIMA Program (150356-S30 and 151815-S30 to plankton alone would appear to be insufficient to M. L. C.), the US National Science Foundation Offices of Polar sustain the rich benthic communities observed in Rijpf- Programs (OPP-0138596, OPP-0222423 to W. G. A.) and Ocean jord, with infaunal benthic biomass of 130 g WW m2 Sciences (OCE-0215905 to S. R. T.), the BBVA Foundation in (M. L. Carroll, unpublished data) at 200 m depth adja- Madrid (to M. G.), and with funds from the Howard Hughes cent to the bivalve collection site. Thus, we assume a Medical Institute through Bates College. We thank the captain and crew of the R/V Lance. Divers H. Hop and S. Nordang significant amount of food must reach the bottom of collected the bivalve samples and J. Edgerly measured growth Rijpfjord from nonlocal sources, with ice algae from ice lines. Sea ice data were provided by the National Snow and Ice flows episodically driven into the fjord by wind and Data Center (USA), and meteorological data were provided by advected phytoplankton being the most likely explana- the Norwegian Meteorological Institute. O. Pavlova collated the  tion. To the extent that the benthos in Rijpfjord is satellite data of ice concentration. Discussions with E. N st- Hegseth clarified our interpretations. J. Carroll, L. Clough, B. dependent on the advection of ice and associated ice Johnson, M. Retelle, P. Renaud, and two anonymous reviewers algae for food, the recent reduction in summer ice north provided valuable comments on earlier versions of the manu- of Svalbard (Serreze et al., 2003; http://nsidc.org/news/ script. press/20050928_trendscontinue.html) or any change in wind or current patterns resulting in less ice entering the fjord would cause a reduction in this food source. References Ice algae is a potentially important food source for the AICA (2004) Impacts of a Warming Arctic: Arctic Climate Impact and Arctic benthos (Ambrose et al., 2001 and references Assessment. Cambridge University Press, Cambridge. therein) including S. groenlandicus (McMahon et al., Allan R, Lindesay J, Parker D (1996) El Nin˜o Southern Oscillation 2006). There may well be a balance between enough and climate variability. CSIRO, Collingswood, Australia. ice for ice algae-laden flows to be advected into Rijpf- Ambrose WG Jr, Clough LM, Tilney PR et al. (2001) Role of jord and too much ice reducing phytoplankton produc- echinoderms in benthic remineralization in the Chukchi Sea. tion in surrounding waters (the source for any advected Marine Biology, 139, 937–949. phytoplankton) and limiting the movement of ice flows. Ambrose WG Jr, Renaud PE (1995) Benthic response to water There are, as yet, insufficient data to determine if there column productivity: evidence for benthic–pelagic coupling in has been a fundamental shift in the biophysical cou- the Northeast Water Polynya. Journal of Geophysical Research, pling in Rijpfjord, but it is possible that the quality or 100, 4411–4421. Andrews JT (1972) Recent and fossil growth rates of marine quantity of food reaching the benthos has recently bivalves, Canadian Arctic, and Late–Quaternary arctic Marine changed. This change could be a reflection of long-term environments. Paleogeography, Paleoclimatology, Paleoecology, 11, climate change in the Arctic that is now exerting an 157–176. overriding effect on the benthos compared to local Bath GE, Thorrold SR, Jones CM (2004) The influence of tem- climatic conditions. perature and salinity on strontium uptake in the otoliths Bivalves have been used to reconstruct environmen- of juvenile spot (Leiostomes xanthurus). Canadian Journal of tal conditions in ancient and modern environments for Fisheries Aquatic Science, 60, 34–42. many decades. The long life of polar bivalves and the Beukema JJ, Cade´e GC (1991) Growth of the bivalve Macoma tight coupling between pelagic and benthic processes balthica in the Wadden Sea during a period of eutrophication: make them ideal candidates for monitoring climate relationships with concentrations of pelagic diatoms and change and for predicting the impact of global warming flagellates. Marine Ecology Progress Series, 68, 249–256. Beukema JJ, Knoll E, Cade´e GC (1985) Effects of temperature on on Arctic marine ecosystems. Our results confirm the the length of the annual growing season of the Tellinid bivalve importance of S. groenlandicus as a biological proxy of Macoma balthica (L.) living on tidal flats in the Dutch Wadden climatic forcing on the marine ecosystem and are the Sea. Journal of Experimental Marine Biology and Ecology, 90, first evidence of an Arctic-centered climate oscillation 129–144. influencing marine benthic processes. Our results, how- Bland JM, Altman DG (1995) Statistical notes. Multiple signifi- ever, do not reveal the proximate causes of variability cance tests: the Bonferroni method. British Medical Journal, in bivalve growth. To fully understand the response of 310, 170.

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 BIVALVE GROWTH AND ARCTIC CLIMATE 1605

Bland JM, Altman DG (1997) Statistics notes: Cronbach’s alpha. Grebmeier JM, McRoy CP, Feder HM (1988) Pelagic–benthic British Medical Journal, 314, 572. coupling on the shelf of the northern Bering and Chukchi Seas Bochkov YA (1982) Water temperature in the 0–200 m layer in the I: food supply and carbon cycling. Marine Ecology Progress Kola Meridian Section in the Barents Sea, 1900–1981 (in Russian), Series, 48, 57–67. Sb. Nauchn. Trud., Polar Research Institute of Marine Fisheries Gulliksen B, Holte B, Jakola K-J (1985) The soft bottom fauna in and Oceanography (PINRO). Murmansk, 46, 113–122. Van Mijenfjord and Raudfjord, Svalbard. In: Marine Biology of Brey T, Pearse J, Basch L et al. (1995) Growth and production of Polar Regions and Effects of Stress on Marine Organisms. Proceed- Sterechinus neumayeri (Echinoida: Echinodermata) in McMurdo ings of the XVIIIth European Marine Biology Symposium Sound, Antarctica. Marine Biology, 124, 279–292. (eds Gray JS, Christiansen ME), pp. 199–215. J. Wiley & Sons, Brockington S, Clarke A (2001) The relative influence of tem- New York. perature and food on the metabolism of a marine invertebrate. Gu¨nther D, Heinrich CA (1999) Enhanced sensitivity in laser Journal of Experimental Marine Biology and Ecology, 258, 87–99. ablation-ICP mass spectrometry using helium-argon mixtures Carroll ML, Carroll J (2003) The Arctic seas. In: Biogeochemistry of as aerosol carrier. Journal of Analytical Atomic Spectrometry, 14, Marine Systems (eds Black KD, Shimmield GB), pp. 127–156. 1363–1368. Blackwell Publishing Ltd., Oxford. Hagberg J, Tunberg G, Wicking G et al. (2004) Effects of climate Cavalieri D, Parkinson C, Gloerson P et al. (1997) Sea ice con- variability on benthic communities. In: Marine Ecosystems and centrations from Nimbus-7 SMMR and DMSP SSM/I passive Climate Variation, The North Atlantic A Comparative Perspective microwave data, January 1979 to December 2003. National (eds Stenseth N Chr., Ottersen G), pp. 115–121. Oxford Snow and Ice Data Center, Boulder, Colorado, USA. Digital University Press, Oxford. media and CD-ROM (http://nsidc.org/data/nsidc-0051.html) Hegseth EN, Svendsen H, von Quillfeldt CH (1995) Phytoplank- (updated 2004). ton in fjords and coastal waters of northern Norway: environ- Clark GR (1974) Growth lines in invertebrate skeletons. Annual mental conditions and dynamics of the spring bloom. In: Review of Earth and Planetary Science, 2, 77–99. Ecology of Fjords and Coastal Waters (eds Skjoldal HR, Hopkins Clough LM, Renaud PE, Ambrose WG Jr (2005) Sediment oxy- C, Erikstad KE, Leinaas HP), pp. 45–72. Elsevier Science B.V., gen demand, infaunal biomass, and sediment pigment con- Amsterdam. centration in the western Arctic Ocean. Canadian Journal of Heilmayer O, Brey T, Chanter M et al. (2003) Age and produc- Fisheries and Aquatic Science, 62, 1756–1765. tivity of the Antarctic scallop, Adamussium colbecki, in Terra Dayton PK (1990) Polar benthos. In: Polar Oceanography, Part B Nova Bay (Ross Sea, Antarctica). Journal of Experimental Marine Chemistry, Biology, and Geology (ed. Smith WO), pp. 631–685. Biology and Ecology, 288, 239–256. Academic Press, San Diego. Hudson I, Shinn E, Halley R et al. (1976) Sclerchronology: a new Dekker R, Beukema J (1999) Relations of summer and winter tool for interpreting past environments. Geology, 4, 361–364. temperatures with dynamics and growth of two bivalves, Hunt GL Jr, Stabeno PJ (2002) Climate change and the control Tellina tenuis and Abra tenuis, on the northern edge of their of energy flow in the southeastern Bering Sea. Progress in intertidal distribution. Journal of Sea Research, 42, 207–220. Oceanography, 55, 5–22. Deser C (2000) On the teleconnectivity of the ‘Arctic Oscillation’. Hurrell JW, Kushnir Y, Ottersee G et al. (2003) The North Atlantic Geophysical Research Letters, 27, 779–782. Oscillation: Climate Significance and Environmental Impact. Dunton KW, Goodall JL, Schonberg SV et al. (2005) Multi-decadal Geophysical Monograph Series. American Geophysical Union, synthesis of benthic-pelagic coupling in the western arctic: role Washington, DC. of cross-shelf advective processes. Deep-Sea Research II, 52, Johannessen OM, Bengtsson L, Miles MW et al. (2004) Arctic 3462–3477. climate change: observed and modeled temperature and Feder HM, Naidu AS, Jewett SW et al. (1994) The northeastern sea-ice variability. Tellus, 4, 328–341. Chukchi Sea: benthos-environmental interactions. Marine Ecol- Johannessen OM, Miles MW, Bengtsson L et al. (2003) Arctic ogy Progress Series, 111, 171–190. climate change. In: Arctic Environment Variability in the Context Fetterer F, Knowles K (2002, updated 2004) Sea Ice Index. National of Global Change (eds Bobylerv PK, Kondratyev KY, Johannes- Snow and Ice Data Center, Boulder, CO, USA Digital media sen OM), pp. 1–15. Praxis Publishing Ltd., Chichester. (updated 2004). Johnson MA, Proshutinsky AY, Polyakov IV (1999) Atmospheric Fomentin JM, Planque B (1996) Calanus and environment in the patterns forcing two regimes of Arctic circulation: a return eastern North Atlantic. II Influence of the North Atlantic to anticyclonic conditions? Geophysical Research Letters, 26, Oscillation on C. finmarchicus and C. helgolandicus. Marine 1621–1624. Ecology Progress Series, 134, 111–118. Jones DS (1981) Annual growth increments in shells of Spisula Fritts HC (1976) Tree Rings and Climate. Academic Press, New York. solidissima recorded marine temperature variability. Science, Grant J, Hargrave G, MacPherson P (2002) Seasonal and spatial 211, 165–165. patterns in mass and organic matter sedimentation in the Jones DS, Arthus MA, Allard DJ (1989) Sclerochronological North Water. Deep-Sea Research II, 49, 5227–5244. records of temperature and growth from shells of Mercenaria Grebmeier JM, Feder HM, McRoy CP (1989) Pelagic–benthic mercenaria from Narragansett Bay, Rhode Island. Marine coupling on the shelf of the northern Bering and Chukchi Seas Biology, 102, 225–234. II: benthic community structure. Marine Ecology Progress Series, Jones PD, Jonsson T, Wheeler D (1997) Extension to the North 51, 253–268. Atlantic Oscillation using early instrumental pressure obser- r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 1606 W. G. AMBROSE et al.

vations from Gibraltar and South-West Iceland. International Ottersen G, Alheit A, Drinkwater K et al. (2004) In: Marine Journal Climatology, 17, 1433–1450. Ecosystems and Climate Variation The North Atlantic A Compara- Khim B-K, Kranz DE, Cooper LW et al. (2003) Seasonal discharge tive Perspective (eds Stenseth NC, Ottersen G), pp. 73–94. to the western Chukchi Sea shelf identified in stable isotope Oxford University Press, Oxford. profiles of mollusk shells. Journal of Geophysical Research, Ottersen G, Planque B, Belgrano A et al. (2001) Ecological effects 108(ca), 16-1–16-10, 3300, doi:10.1029/2003JC001816. of the North Atlantic Oscillation. Oecologia, 128, 1–14. Kro¨ncke I, Dippner JW, Heyen H et al. (1998) Long-term changes Overpeck JK, Hughen K, Hardy D et al. (1997) Arctic environ- in macrofaunal communities off Norderney (East Frisia, Ger- mental change of the last four centuries. Science, 278, many) in relation to climate variability. Marine Ecology Progress 1251–1256. Series, 167, 25–36. Peck LS, Bullough LW (1993) Growth and population structure Kro¨ncke I, Zeiss B, Rensing C (2001) Long-term variability in in the infaunal bivalve Yoldia eightsi in relation to iceberg macrofauna species composition off the Island of Norderney activity at Signy Island, Antarctica. Marine Biology, 117, (East Frisia, Germany) in relation to changes in climate and 235–241. environmental conditions. Senckenbergiana Maritima, 31, 65–82. Pershing AJ, Greene CH, Planque B et al. (2004) The influence of Lewis DE, Cerrato RM (1997) Growth uncoupling and the climate variability on North Atlantic zooplankton populations. relationship between shell growth and metabolism in the soft In: Marine Ecosystems and Climate Variation The North Atlantic shell clam Mya arenaria. Marine Ecology Progress Series, 158, A Comparative Perspective (eds Stenseth NC, Ottersen G), 177–189. pp. 59–94. Oxford University Press, Oxford. Loeng H (1991) Features of the physical oceanographic condi- Piepenburg D, Ambrose WG Jr, Brandt A et al. (1997) Benthic tions of the Barents Sea. Polar Research, 10, 5–18. community patterns reflect water column processes in the Maxwell B (1997) Recent climate patterns in the Arctic. In: Global Northeast Water Polynya (Greenland). Journal of Marine Change and Arctic Terrestrial Ecosystems (eds Oechel WC, Call- Systems, 10, 476–482. aghan T, Gilmarov T, Holten JI, Maxwell B, Molau U, Sveinb- Polyakov IV, Proshutinsky AY, Johnson MA (1999) Seasonal jo¨rnsson B), pp. 21–46. Springer Verlag, New York. cycles in two regimes of Arctic climate. Journal of Geophysical McDonald J, Feder HM, Hoberg M (1981) Bivalve mollusks of the Research, 104, 25761–25788. Southeastern Bering Sea. In: The Eastern Bering Sea Shelf: Post E (2004) Time lags in terrestrial and marine environments. Oceanography and Resources (eds Hood DW, Calder JA), pp. In: Marine Ecosystems and Climate Variation The North Atlantic 1155–1204. University of Washington Press, Seattle. A Comparative Perspective (eds Stenseth NC, Ottersen G), McMahon KW, Ambrose WG Jr, Johnson BJ et al. (2006) Benthic pp. 165–167. Oxford University Press, Oxford. community response to ice algae and phytoplankton in Ny Proshutinsky A, Johnson M (1997) Two circulation regimes of the A˚ lesund, Svalbard. Marine Ecology Progress Series, 310, 1–14. wind-driven Arctic Ocean. Journal of Geophysical Research, 102, Morison J, Aagaard K, Steele M (2000) Recent environmental 12493–12512. changes in the Arctic: a review. Arctic, 53, 359–371. Rhoads DC, Lutz RA (eds) (1980) Skeletal Growth of Aquatic Mu¨ller-Lupp T, Bauch HA (2005) Linkage of Arctic shelf atmo- Organisms: Biological Records of Environmental Change. Plenum spheric circulation and Siberian shelf hydrography: a proxy Press, New York. validation using d18O records of bivalve shells. Global Planetary Rhoads DC, Pannella G (1970) The use of molluscan shell growth Change, 48, 175–186. patterns in ecology and paleoecology. Lethaia, 3, 143–161. Mu¨ller-Lupp T, Erlenkeuser H, Bauch HA (2003) Seasonal and Rosenthal Y, Field MP, Sherrell RM (1999) Precise determination interannual variability of Siberian river discharge in the of element/calcium ratios in calcareous samples using sector Laptev Sea inferred from stable isotopes in modern bivalves. field inductively coupled plasma mass spectrometry. Analyti- Boreas, 32, 292–303. cal Chemistry, 71, 3248–3253. Nielsen TG, Hansen B (1995) Plankton community structure and Rysgaard S, Nielsen T, Hansen BW (1999) Seasonal variation in carbon cycling on the western coast of Greenland during and nutrients, pelagic primary production and grazing in a high- after the sedimentation of a diatom bloom. Marine Ecology Arctic marine ecosystem, Young Sound, Northeast Greenland. Progress Series, 125, 239–257. Marine Ecology Progress Series, 179, 13–25. Norkko J, Norkko A, Thursh SF et al. (2005) Detecting growth Scho¨ne BR, Houk SD, Freyre Castro AD et al. (2005) Daily growth under environmental extremes: spatial and temporal patterns rates in shells of Arctica islandica: assessing sub-seasonal in nucleic acid rations in two Antarctic bivalves. Journal of environmental controls on a long-lived bivalve mollusk. Experimental Marine Biology and Ecology, 326, 144–156. Palaios, 20, 78–92. Oechel WC, Vourlitis GL (1997) Climate change in northern Scho¨ne BR, Oschmann W, Ro¨ssler J et al. (2003) North Atlantic latitudes: alterations in ecosystem structure and function and Oscillation dynamics recorded in shells of a long-lived bivalve effects on carbon sequestration. In: Global Change and Arctic mollusk. Geology, 31, 1037–1042. Terrestrial Ecosystems (eds Oechel WC, Callaghan T, Gilmarov Sejr MK, Christensen PB (2006) Growth, production and carbon T, Holten JI, Maxwell B, Molau U, Sveinbjo¨rnsson B), pp. 381– demand of macrofauna in Young Sound, with special empha- 401. Springer, New York. sis on the bivalves Hiatella arctica and Mya truncate. In: Carbon Osborn TJ, Briffa KR, Tett SFB et al. (1999) Evaluation of the Cycling in Arctic Marine Ecosystems: Case Study Young Sound. North Atlantic Oscillation as simulated by a coupled climate Meddr Greenland, Bioscience Special Issue (eds Rysgaard S, model. Climate Dynamics, 15, 685–702. Glud RN) in press.

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607 BIVALVE GROWTH AND ARCTIC CLIMATE 1607

Sejr M, Jensen KT, Rysgaard S (2002) Annual growth bands in the Thorrold SR, Latkoczy C, Swart PK et al. (2001) Natal homing bivalve Hiatella arctica validated by a mark-recapture study in a marine fish. Science, 291, 297–299. in NE Greenland. Polar Biology, 25, 794–796. Tunberg BG, Nelson WG (1998) Do climate oscillations influence Sejr M, Petersen JK, Jensen TK et al. (2004) Effects of food cyclic patterns of soft bottom macrobenthic communities concentration on clearance rate and energy budget of the on the Swedish west coast? Marine Ecology Progress Series, Arctic bivalve Hiatella arctica (L) at subzero temperature. 170, 85–94. Journal of Experimental Marine Biology and Ecology, 311, 171–183. Turekian KK, Cochran JK, Kharkar D et al. (1975) Slow growth Serreze MC, Maslanik JA, Scambos TA et al. (2003) A record rate of a deep-sea clam determined by 228 Ra chronology. minimum arctic sea ice extent and area in 2002. Geophysical Proceedings of the National Academy of Science USA, 72, 2829– Research Letters, 30, 111, doi:10.1029/2002GL016406. 2832. Serreze MC, Walsh JE, Chapin FS III et al. (2000) Observational Walther GR, Post E, Convery P et al. (2002) Ecological responses evidence of recent change in the northern high-latitude to recent climate change. Nature, 414, 389–395. environment. Climate Change, 46, 159–207. Witbaard R (1996) Growth variations in Arctica islandica L. Stabeno PJ, Overland JE (2001) Bering Sea shifts toward an (): a reflection of hydrography-related food supply. earlier spring transition. EOS, 82, 317–321. ICES Journal of Marine Science, 53, 981–987. Stecher HA III, Krantz DE, Lord CJ III et al. (1996) Profiles of Witbaard R, Duineveld GCA, de Wilde PAWJ (1999) Geographi- strontium and barium in Mercenaria mercenaria and Spisula soli- cal differences in growth rates of Arctica islandica (Mollusca: dissima shells. Geochimica et Cosmochimica Acta, 60, 3445–3456. Bivalvia) from the North Sea and adjacent waters. Journal of the Stein R, Macdonald RW (eds) (2004) The Organic Carbon Cycle in Marine Biological Association of the United Kingdom, 79, 907–915. the Arctic Ocean. Springer, Berlin. Witbaard R, Franken R, Visser B (1997) Growth of juvenile Arctica Stenseth NC, Ottersen G, Hurrell JW et al. (2003) Studying islandica under experimental conditions. Helgolander Meeresun- climate effects on ecology through the use of climate indices: tersuchungen, 51, 417–431. the North Atlantic Oscillation, El Nin˜o Southern Oscillation Witbaard R, Jansma E, Sass Klaassen U (2003) Copepods link and beyond. Proceedings of the Royal Society of London, 270, quahog growth to climate. Journal of Sea Research, 50, 77–83. 2087–2096. Witbaard R, Jenness MI, Van der Brog K et al. (1994) Verification Tallqvist MI, Sundet JH (2000) Annual growth of the cockle of annual growth increments in Arctica islandica L. from the Clinocardium ciliatum in the Norwegian Arctic (Svalbard area). North Sea by means of oxygen and carbon isotopes. Nether- Hydrobiologia, 440, 331–338. lands Journal of Sea Research, 333, 91–101. Tereshchenko VV (1997) Seasonal and interannual temperature and Wollenburg JE, Kuhnt W (2000) The response of benthic forami- salinity variations of main flows at the Kola section of the Barents nifera to carbon flux and primary production in the Arctic Sea. Polar Research Institute of Marine Fisheries and Oceano- Ocean. Marine Micropaleontology, 40, 189–231. graphy (PINRO), Murmansk, 71 pp (with updates). Zacherl DC, Paradis G, Lea DW (2003) Barium and strontium Thompson I, Jones DS, Dreibelbis D (1980) Annual internal uptake into larval protoconchs and statoliths of the marine growth banding and life history of the ocean quahog Arctica neogastropod Kelletia kelletii. Geochimica et Cosmochimica Acta, islandica (Mollusca: Bivalvia). Marine Biology, 57, 25–34. 21, 4091–4099. Thompson PM, Ollason J (2001) Lagged effects of ocean climate Zenkevich L (1963) Biology of the Seas of the USSR. George Allen & change on fulmar population dynamics. Nature, 413, 417–420. Unwin, London. Thompson D, Wallace J (1998) The Arctic Oscillation signature in Zolotarev VN (1980) The life span of bivalves from the Sea of the wintertime geopotential height and temperature fields. Japan and Sea of Okhotsk. Soviet Journal of Marine Biology, Geophysical Research Letters, 25, 1297–1300. 6, 301–308.

r 2006 The Authors Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 12, 1595–1607