You are on page 1of 26

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, C10011, doi:10.

1029/2006JC003706, 2007

Seasonal rhythms of net primary production and particulate organic


carbon flux to depth describe the efficiency of biological pump in the
global ocean
Michael J. Lutz,1 Ken Caldeira,2 Robert B. Dunbar,1 and Michael J. Behrenfeld3
Received 10 June 2005; revised 8 January 2007; accepted 2 February 2007; published 10 October 2007.

[1] We investigate the functioning of the ocean’s biological pump by analyzing the
vertical transfer efficiency of particulate organic carbon (POC). Data evaluated include
globally distributed time series of sediment trap POC flux, and remotely sensed estimates
of net primary production (NPP) and sea surface temperature (SST). Mathematical
techniques are developed to compare these temporally discordant time series using NPP
and POC flux climatologies. The seasonal variation of NPP is mapped and shows
regional- and basin-scale biogeographic patterns reflecting solar, climatic, and
oceanographic controls. Patterns of flux are similar, with more high-frequency variability
and a subtropical-subpolar pattern of maximum flux delayed by about 5 days per degree
latitude increase, coherent across multiple sediment trap time series. Seasonal
production-to-flux analyses indicate during intervals of bloom production, the sinking
fraction of NPP is typically half that of other seasons. This globally synchronous pattern
may result from seasonally varying biodegradability or multiseasonal retention of POC.
The relationship between NPP variability and flux variability reverses with latitude, and
may reflect dominance by the large-amplitude seasonal NPP signal at higher latitudes.
We construct algorithms describing labile and refractory flux components as a function of
remotely sensed NPP rates, NPP variability, and SST, which predict POC flux with
accuracies greater than equations typically employed by global climate models. Globally
mapped predictions of POC export, flux to depth, and sedimentation are supplied. Results
indicate improved ocean carbon cycle forecasts may be obtained by combining
satellite-based observations and more mechanistic representations taking into account
factors such as mineral ballasting and ecosystem structure.
Citation: Lutz, M. J., K. Caldeira, R. B. Dunbar, and M. J. Behrenfeld (2007), Seasonal rhythms of net primary production and
particulate organic carbon flux to depth describe the efficiency of biological pump in the global ocean, J. Geophys. Res., 112, C10011,
doi:10.1029/2006JC003706.

1. Introduction vertical flux of particulate organic carbon between oceano-


graphic regions [Lutz et al., 2002; Neuer et al., 2002]
[2] The marine biosphere is a major component of the implies the potential for substantial biogeochemical con-
global carbon cycle, responsible for roughly half of the sequences resulting from alterations to upper ocean dyna-
annual photosynthetic absorption of CO2 from the atmo- mics. Changes in the efficiency of the marine organic
sphere [Field et al., 1998]. However, the influence of carbon cycle may contribute to variability of atmospheric
marine carbon cycle variability on atmospheric CO2 con- CO2 on glacial-interglacial timescales [Sigman and Boyle,
centrations is poorly understood [Sarmiento and Le Quéré, 2000]. Understanding of the fate of biogenic carbon within
1996; Le Quéré et al., 2005]. In particular, there is little the ocean is essential for improved forecasts of future
basis to predict how potential alterations in upper ocean atmospheric CO 2 concentrations and rates of global
ecosystems may influence the capacity of the ocean to store warming, especially in the context of changing ocean
carbon [Sarmiento et al., 1998]. Significant variability in the circulation [Sarmiento et al., 1998; Schmittner, 2005].
[3] The transport of biogenic elements from surface
1
waters to the deep ocean and sediments occurs through a
Department of Geological and Environmental Sciences, Stanford
University, Stanford, California, USA.
variety of processes collectively known as the ‘‘biological
2
Department of Global Ecology, Carnegie Institution of Washington, pump’’ [Volk and Hoffert, 1985]. This biogeochemical and
Stanford, California, USA. physical system may be conceptually divided into four
3
Department of Botany and Plant Pathology, Oregon State University, interrelated components: production, export, flux to depth,
Corvallis, Oregon, USA. and sedimentation. During primary production, phytoplank-
Copyright 2007 by the American Geophysical Union.
ton incorporate dissolved nutrients and inorganic carbon
0148-0227/07/2006JC003706 into particulate organic matter (POM), lowering the partial

C10011 1 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

pressure of CO2 in surface waters, thus enhancing the upper ocean food web [Berger et al., 1989; Berger and
ocean’s ability to absorb CO2 from the atmosphere. Much Wefer, 1990; Lampitt and Antia, 1997; Rühlemann et al.,
POM is recycled within surface waters where associated 1999; Fischer et al., 2000]. This hypothesis suggests that
nutrients support regenerated primary production [Eppley organic matter recycling is influenced by controls related to
and Peterson, 1979]. A fraction this POM is exported variability of upper ocean ecosystem structure. In particular,
beneath surface waters and fluxes to depth within the the fraction and biodegradability of production exported
ocean’s interior where it is predominantly consumed and from more variable ecosystems may be greater than from
regenerated by subsurface heterotrophs. The minor portion stable ones. Where the supply of solar irradiation and
of POM that reaches the seafloor is mostly remineralized in nutrients are relatively constant, environments are stable
support of benthic biological activity. Flux to the seafloor throughout the year and activities of autotrophs and hetero-
escaping benthic regeneration is buried in the sediments. trophs are balanced. Trophic transfer is efficient and recy-
[4] In general, the greater the depth at which sinking cling dominates. The fraction of production exported is
organic carbon is regenerated, the longer time it takes to minimal. Conversely, seasonal environments reflect time-
return to the photic zone as dissolved CO2, where it may varying physical conditions, solar irradiation, and nutrients,
reenter atmospheric carbon cycle. Organic carbon reaching and are characterized by ‘‘leaky’’ food webs with tempo-
the deep ocean is entrained in water masses with longer rally imbalanced components. New production dominates
flow pathways back to the surface and smaller advective and rapid autotrophic growth outpaces consumer feeding
velocities than in the upper ocean. Ventilation of subsurface and decompositional activity. Particulate matter has a grea-
waters occurs on timescales ranging from annual to ter opportunity to avoid surface heterotrophy and be
hundreds of years in the upper ocean (100– 1000 m) to exported in a more labile or ‘‘fresh’’ state. Thus in variable
thousands of years in the deep ocean (>1500 m). Carbon environments particulate matter has a greater opportunity to
entering sediments may be geologically stored for hundreds escape the euphotic zone and be more biodegradable than in
of millions of years. Thus to describe the storage of stable environments.
biogenic carbon in the ocean, the depth of remineralization
must be known. 3. Previous Flux Algorithms
[5] This paper explores how global-scale environmental
parameters, sea surface temperature (SST) and the seaso- [8] The empirical algorithms of Suess [1980], equation
nality of net primary production (NPP), relate to the (1) [Six and Maier-Reimer, 1996; Dymond et al., 1997;
efficiency of the biological pump. The ability to infer SST Tyrell, 1999], and Martin et al. [1987], equation (2)
and NPP from remote observations allows for their synoptic [Sarmiento et al., 1998; Hedges et al., 1999; Buesseler et
comparison to sediment trap data. This investigation intro- al., 2000; Fischer et al., 2000] are commonly used to
duces methodology to facilitate the synthesis of these time describe oceanic particulate organic carbon flux:
series. Our metadata analysis is used to describe the
seasonal efficiency of production-to-flux relationships, eva- Cnpp
CfluxðzÞ ¼ ð1Þ
luate hypotheses regarding the functioning of the biological ð0:0283z þ 0:212Þ
pump, assemble new particle flux algorithms, and test
parameterization accuracy.
 0:858
z
CfluxðzÞ ¼ Cexport ð2Þ
2. Surface Temperature and Seasonality of NPP z0
[6] Recent evidence indicates water temperature may be a Particulate organic carbon flux (Cflux(z); mg Corg m2 d1) is
key environmental parameter influencing the proportion of described as a function of the production of organic carbon
NPP exported from surface waters to the deep ocean [Laws in surface waters (Cnpp) or the export of organic carbon
et al., 2000; Rivkin and Legendre, 2001]. This hypothesis (Cexport) from the base of the euphotic zone (z0), scaled to
suggests the rate of organic matter recycling is influenced depth below the sea-surface (z). The Suess [1980] equation
by SST. Specifically, lower-water temperatures limit the was determined from field-based production and sediment
activity of microbial heterotrophic decomposers more than trap flux measurements, collected from the subtropical
the activity of autotrophic producers. This scenario implies eastern Pacific and northwestern Atlantic at depths between
that in colder regions of the ocean, a greater fraction of 50 and 5400 m. The Martin et al. [1987] normalized power
production is not remineralized and is therefore available for function is a ‘‘best fit’’ based on sediment trap data
export. Furthermore, due to the lack of surface recycling collected in the low to midlatitude east Pacific from depths
where SST is colder, exported detritus may be more labile between 100 and 2000 m. Global carbon models including
than where SST is warmer. Laws et al. [2000] propose a marine biology commonly describe flux to depth using the
euphotic zone food web model describing biogenic export fixed parameters shown in equations (1) and (2).
relying heavily on this premise. In support of this rationale,
the authors demonstrate a strong near-linear correlation
between observed f ratios, the ratio of new to total produc- 4. Methods
tion [Eppley and Peterson, 1979], and those predicted by [9] Our approach builds on the relationships of Suess
their model. Water temperature alone explains most of the [1980] and Martin et al. [1987] [equations (1) and (2)] by
f ratio variability found in the model by Laws et al. [2000]. using the greater amount of recently available data to
[7] Another proposed control on biogenic carbon export constrain more of the variability between production and
is variability in the efficiency of trophic transfer within the flux to depth. In particular, we compare globally distributed

2 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 1. Geographic locations of sediment trap particulate organic carbon flux to depth measurements.
Observations include annual estimates (6) and subannual time series (.). Arrows indicate locations of
the meridional transects of net primary production time series shown in Figure 5.

annual and time series sediment trap flux measurements, to the seafloor [Honjo, 1978]. Settling detritus intercepted
and remotely sensed estimates of NPP and SST. Sediment by these funnels collects in sample bottles opened and
traps and satellites allow for the characterization of closed at the funnel base. Each sediment trap measurement
marine biogeochemical variability on timescales longer reflects the total mass of flux collected during each bottle
than the data sets collected through traditional ship-based open-close interval. Sampling frequency and interval dura-
oceanography. tion typically varies within sediment trap experiments.
[10] The global sediment trap data set assembled for this Experiments often include interruptions produced during
analysis contains estimates of flux collected during the collection interval malfunctions and between trap deploy-
past 25 years. Satellite-based estimates of production are ments. The result of a sediment trap experiment is typically
employed because field-based observations of production a discontinuous time series of average flux rates represen-
are not available at the same locations and durations as the ting bottle open-close intervals of variable durations.
sediment trap experiments. Annual climatologies are used to [12] The following criteria are used in assembling sedi-
compare the seasonality of these temporally discordant time ment trap data that best reflect surface production (Table 1
series. Interannual variability is not focused upon in this and Figures 1 and 2):
study. [13] (1) Values are from depths greater than the local
export depth (ze). The depth of export herein describes
4.1. Sediment Traps maximum of the euphotic zone and mixed layer depths
[11] Moored sediment traps are funnel-shaped devices during flux formation. Sediment traps located above this
attached at varying depths to line that is fixed in location depth may not reflect the flux of particulate matter to the

Figure 2. Depth versus latitude of sediment trap particulate organic carbon flux to depth measurements.
Observations include annual estimates (6) and subannual time series (.).

3 of 26
Table 1. Sediment Trap Particulate Organic Carbon Flux to Depth (Corg) Observations
C10011

Water Trap Collection interval


depth depth Corg
Region Trap ID Lat. Lon. (m) (m) Start End (mg m2 d1) Labela Reference
Pacific Ocean
Central Bering Sea 58 179 3783 3137 18 Jul 1991 12 Jun 1992 5.53 Honjo [1996]b
Bering Sea, Aluetian Basin AB 53.5 177 3788 3198 7 Aug 1990 25 Jul 1995 6.96 1-D Takahashi et al. [2000]c
Central Okhotsk Sea OS 53 149 1170 258 12 Aug 1990 12 Aug 1991 8.58 Honjo [1996]b
Central Okhotsk Sea OS 53 149 1170 1061 12 Aug 1990 12 Aug 1991 4.08 1-I Honjo [1996]b
S of Kamchatka Peninsula GD 51.5 165 4960 4500 Jul 1991 Jun 1992 4.38 2-D Wong et al. [1994]d,e
W Pacific Subarctic Gyre 50N 50 165 5500 1227 1 Dec 1997 18 May 2000 4.6 2-I Honda [2001]; Honda et al. [2002]b,c
W Pacific Subarctic Gyre 50N 50 165 5500 3260 1 Dec 1997 18 May 2000 3.1 3-D Honda [2001]; Honda et al. [2002]b,c
W Pacific Subarctic Gyre 50N 50 165 5500 5090 1 Dec 1997 18 May 2000 2.3 4-D Honda [2001]; Honda et al. [2002]b,c
Subarctic Pacifc OSP 50 145 4240 200 8 May 1989 16 May 1994 18.2 1-S Wong et al. [1999]c,d,e
Subarctic Pacifc OSP 50 145 4240 1000 27 Mar 1983 16 May 1994 7.42 2-S Wong et al. [1999]c,d,e
Subarctic Pacifc OSP 50 145 4240 3800 23 Sep 1982 16 May 1994 3.10 5-D Wong et al. [1999]c,d,e
Subarctic Pacifc SA 49 173.9 5406 4806 9 Aug 1990 2 Aug 1995 3.93 6-D Takahashi et al. [2000]c
Juan de Fuca Ridge JDF 48.0 128.1 na 2200 9 Sep 1984 10 Aug 1985 2.99 Dymond and Lyle [1994]
N Central Pacific NP-B 46.8 162.1 5670 700 Aug 1985 1 Jun 1986 33.7 Tsunogai and Noriki [1991]d
N Central Pacific NP-B 46.8 162.1 5670 5200 Aug 1985 1 Jun 1986 5.1 Tsunogai and Noriki [1991]d
Subarctic Front Site 8 46.1 175.0 5435 1412 15 Jun 1993 16 Apr 1994 8.11 3-I Kawahata [2002]b
E of Kurile Is. GA 45 165 5830 5300 Jul 1991 Jun 1992 3.01 7-D Wong et al. [1994]d,e
S of Aleutian Is. GB 45 177 6100 5600 Jul 1991 Jun 1992 1.37 8-D Wong et al. [1994]d,e
W Pacific Subarctic Gyre KNOT 44 155 5500 924 1 Dec 1997 13 May 2000 9.1 3-S Honda et al. [2002]c,d
W Pacific Subarctic Gyre KNOT 44 155 5500 2960 1 Dec 1997 13 May 2000 6.1 Honda et al. [2002]c
W Pacific Subarctic Gyre KNOT 44 155 5500 4989 1 Dec 1997 5 Mar 2000 4.2 Honda et al. [2002]c
California Current, Midway MW 42.2 127.6 na 2830 22 Sep 1987 16 Sep 1988 6.02 Dymond and Lyle [1994]

4 of 26
California Current, Nearshore NS 42.1 125.8 na 2829 22 Sep 1987 16 Sep 1988 13.4 Dymond and Lyle [1994]
California Current, Gyre G 41.6 132 na 3664 28 Sep 1987 14 Sep 1988 2.46 Dymond and Lyle [1994]
NW Pacific EM 40.9 142.0 5370 5000 1 Sep 1988 1 May 1989 12.8 Tsunogai and Noriki [1991]d
Subarctic boundary 40N 40 165 5500 953 1 Dec 1997 20 Jan 2000 4.9 4-S Honda [2001]; Honda et al. [2002]b,c
Subarctic boundary 40N 40 165 5500 2986 1 Dec 1997 20 Jan 2000 3.9 9-D Honda [2001]; Honda et al. [2002]b,c
Subarctic boundary 40N 40 165 5500 5016 1 Dec 1997 20 Jan 2000 3.4 10-D Honda [2001]; Honda et al. [2002]b,c
E Japan Sea Sta. T 39.7 132.4 3300 2800 19 Jul 1994 18 Jul 1995 8.1 11-D Hong et al. [1996]
California Current MFZ 39.5 127.7 na 4230 9 Sep 1983 1 Sep 1984 3.48 Dymond and Lyle [1994]
Subarctic Front Site 7 37.4 174.9 5105 1482 1 Jun 1993 9 Apr 1994 6.44 4-I Kawahata [2002]b
Subarctic Front Site 7 37.4 174.9 5105 4588 1 Jun 1993 9 Apr 1994 3.06 12-D Kawahata [2002]b
Monterey Bay S1 36.8 122.0 700 450 Aug 1989 1 Nov 1992 39.5 Pilskaln et al. [1996]c
Kuroshio Extension WCT-7-s 36.7 154.9 5578 1191 19 Aug 1999 29 Aug 2000 4.44 5-I Mohiuddin et al. [2004]
LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP

Kuroshio Extension WCT-7-d 36.7 154.9 5578 5034 19 Aug 1999 29 Aug 2000 3.83 13-D Mohiuddin et al. [2004]
Kuroshio Extension WCT-3-s 36.0 147.0 5615 1108 30 Aug 1998 10 Aug 1999 2.50 6-I Mohiuddin et al. [2004]
Kuroshio Extension WCT-3-d 36.0 147.0 5615 5081 30 Aug 1998 10 Aug 1999 4.06 14-D Mohiuddin et al. [2004]
Subtropical Front Site 5 34.4 177.7 3365 1342 15 Jun 1993 16 Apr 1994 2.89 7-I Kawahata [2002]b
Subtropical Front Site 5 34.4 177.7 3365 2848 15 Jun 1993 1 Jun 1994 4.28 15-D Kawahata [2002]b
Santa Barbara Basin 34.2 120.0 590 540 12 Aug 1993 10 Sep 1996 12.3 Thunell [1998a]c
Subtropical Front Site 6 30.0 175.0 5390 3873 15 Jun 1993 1 Jun 1994 2.67 16-D Kawahata [2002]b
Gulf of California Guaymas Basin 27.9 111.7 700 500 8 Jul 1990 5 Jan 1997 18.9 5-S Thunell [1998b]c
Okinawa Trough JAST01 27.2 126.4 1650 1010 19 Jan 1993 10 Aug 1995 6.23 8-I Honda [2001]b,c
Okinawa Trough JAST01 27.2 126.4 1650 1547 19 Jan 2003 30 Dec 2003 4.11 9-I Honda [2001]b
Ryukyu Trench JAST03 25.1 127.4 3771 3157 8 Oct 1994 10 Aug 1995 1.23 17-D Honda [2001]b,c
South China Sea SCS-N 18.5 116 3750 1000 10 Sep 1987 21 Oct 1988 3.92 Wiesner et al. [1996]; Jianfang et al. [1998]
South China Sea SCS-N 18.5 116 3750 3350 10 Sep 1987 21 Oct 1988 2.02 Wiesner et al. [1996]; Jianfang et al. [1998]
C10011
Table 1. (continued)
Water Trap Collection interval
C10011

depth depth Corg


Region Trap ID Lat. Lon. (m) (m) Start End (mg m2 d1) Labela Reference
South China Sea SCS-C 14.6 115.1 4310 1200 1 Dec 1990 29 May 1995 4.20 Wiesner et al. [1996]; Jianfang et al. [1998]c
South China Sea SCS-C 14.6 115.1 4310 2240 1 Dec 1990 29 May 1995 3.51 Wiesner et al. [1996]; Jianfang et al. [1998]c
South China Sea SCS-C 14.6 115.1 4310 3770 1 Dec 1990 29 May 1995 2.52 Wiesner et al. [1996]; Jianfang et al. [1998]c
N Equatorial Current NEC 12.0 134.3 5300 1200 21 Nov 1988 16 Dec 1989 0.38 10-I Kemp and Knaack [1996]
N Equatorial Current NEC 12.0 134.3 5300 4300 21 Nov 1988 16 Dec 1989 0.43 18-D Kemp and Knaack [1996]
N Equatorial Counter Current MANOP-S 11 140 na 700 29 Dec 1982 14 Feb 1984 3.12 6-S Dymond and Collier [1988]c
N Equatorial Counter Current MANOP-S 11 140 na 1600 29 Dec 1982 14 Feb 1984 2.33 11-I Dymond and Collier [1988]c
N Equatorial Counter Current MANOP-S 11 140 na 3400 29 Dec 1982 14 Feb 1984 1.61 19-D Dymond and Collier [1988]c
N Equatorial Counter Current MANOP-S 11.1 140.1 na 4620 29 Dec 1982 14 Feb 1984 0.82 Dymond and Lyle [1994]
Equatorial Pacific JGOFS-EqPac-9N 9 140 5100 2150 2 Feb 1992 24 Jan 1993 1.51 12-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-9N 9 140 5100 2250 2 Feb 1992 24 Jan 1993 1.52 13-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-9N 9 140 5100 4400 2 Feb 1992 4 Dec 1992 1.09 20-D Honjo et al. [1995]
E Tropical Pacific MANOP-M 8.8 104.0 na 3150 12 Sep 1980 23 Oct 1981 3.78 Dymond and Lyle [1985]; Dymond and Lyle [1994]
Equatorial Counter Current Site 4 7.9 175.0 5260 4743 25 Sep 1992 13 Apr 1993 0.78 Kawahata et al. [2000]f
E Tropical Pacific MANOP-H 6.6 92.8 na 3565 20 Sep 1980 17 Oct 1981 2.44 Dymond and Lyle [1985]; Dymond and Lyle [1994]
Panama Basin PB2 5.4 85.6 3860 890 3 Dec 1979 2 Dec 1980 9 7-S Honjo [1982]d
Panama Basin PB2 5.4 85.6 3860 2590 3 Dec 1979 2 Dec 1980 11 21-D Honjo [1982]d
Panama Basin PB2 5.4 85.6 3860 3560 3 Dec 1979 2 Dec 1980 14 22-D Honjo [1982]d
Equatorial Pacific JGOFS-EqPac-5N 5 140 4493 1191 2 Feb 1992 24 Jan 1993 6.02 14-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-5N 5 140 4493 2091 2 Feb 1992 24 Jan 1993 4.50 15-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-5N 5 140 4493 3793 2 Feb 1992 7 Jan 1993 3.69 23-D Honjo et al. [1995]
Equatorial Counter Current ECC 5 138.8 4130 1130 21 Nov 1988 16 Dec 1989 1.78 16-I Kemp and Knaack [1996]
Equatorial Counter Current ECC 5 138.8 4130 3130 21 Nov 1988 16 Dec 1989 0.67 24-D Kemp and Knaack [1996]

5 of 26
Equatorial Counter Current Site 2 4.1 136.3 4888 1769 4 Jun 1991 15 Apr 1992 6.72 17-I Kawahata et al. [2000]; Kawahata et al. [2002]f
Equatorial Counter Current Site 2 4.1 136.3 4888 4574 4 Jun 1991 15 Apr 1992 6.06 25-D Kawahata et al. [2000]f
Equatorial Counter Current Site 1 3.0 135.0 4402 1592 4 Jun 1991 27 Apr 1992 9.33 18-I Kawahata et al. [1998]; Kawahata et al. [2002]f
Equatorial Counter Current Site 1 3.0 135.0 4402 3902 4 Jun 1991 27 Apr 1992 7.83 26-D Kawahata et al. [1998]f
Equatorial Pacific JGOFS-EqPAC-2N 2 140 4397 2203 2 Feb 1992 24 Jan 1993 4.02 19-I Honjo et al. [1995]
S Equatorial Current Site 10 1.22 160.6 3181 1164 1 Oct 1994 16 Apr 1995 1.94 20-I Kawahata et al. [2000]
Equatorial Pacific C 1.04 138.9 na 4445 23 Dec 1982 3 May 1985 3.56 Dymond and Lyle [1994]c
S Equatorial Current MANOP-C 1 139 na 1089 12 Dec 1982 1 May 1985 4.17 21-I Dymond and Collier [1988]c
S Equatorial Current MANOP-C 1 139 na 1889 12 Dec 1982 1 May 1985 5.52 22-I Dymond and Collier [1988]c
S Equatorial Current MANOP-C 1 139 na 2908 12 Dec 1982 25 Feb 1984 5.18 27-D Dymond and Collier [1988]
S Equatorial Current MANOP-C 1 139 na 3495 25 Feb 1984 1 May 1985 3.26 28-D Dymond and Collier [1988]
Equatorial Pacific JGOFS-EqPac-Eq 0.0 140 4358 880 2 Feb 1992 24 Jan 1993 4.65 8-S Honjo et al. [1995]
LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP

Equatorial Pacific JGOFS-EqPac-Eq 0.0 140 4358 2284 2 Feb 1992 24 Jan 1993 4.49 23-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-Eq 0.0 140 4358 3618 2 Feb 1992 24 Jan 1993 4.37 29-D Honjo et al. [1995]
S Equatorial Current Site 3 0.0 175.2 4880 1357 1 Jun 1992 16 Apr 1993 2.44 24-I Gupta and Kawahata [2000]; Kawahata et al. [2000]f
S Equatorial Current Site 3 0.0 175.2 4880 4363 1 Jun 1992 16 Apr 1993 1.50 30-D Gupta and Kawahata [2000]; Kawahata et al. [2000]f
Equatorial Pacific JGOFS-EqPac-2S 2 140 4293 3593 2 Feb 1992 24 Jan 1993 3.61 31-D Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-5S 5 140 4198 2099 2 Feb 1992 24 Jan 1993 2.73 25-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-5S 5 140 4198 2209 2 Feb 1992 24 Jan 1993 2.72 26-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-5S 5 140 4198 2316 2 Feb 1992 24 Jan 1993 2.79 27-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-12S 12 135 4294 1292 2 Feb 1992 24 Jan 1993 1.52 28-I Honjo et al. [1995]
Equatorial Pacific JGOFS-EqPac-12S 12 135 4294 3594 2 Feb 1992 24 Jan 1993 0.71 32-D Honjo et al. [1995]
Tropical Converg., Coral Sea Site 11 13.0 156.0 1832 1315 16 May 1995 1 Apr 1996 1.50 29-I Kawahata and Ohta [2000]f
Tropical Converg., Coral Sea Site 12 17.8 154.8 2821 2304 16 May 1995 16 Mar 1996 0.56 30-I Kawahata and Ohta [2000]f
Peru-Chile current CH3, CH4 29.5 73.2 4360 2323 21 Jul 1993 23 Sep 1994 7.69 31-I Hebbeln et al. [2000]c
Peru-Chile current CH1, CH3 29.5 73.2 4360 3687 2 Nov 1992 23 Jan 1994 7.56 33-D Hebbeln et al. [2000]c
C10011
Table 1. (continued)
Water Trap Collection interval
C10011

depth depth Corg


Region Trap ID Lat. Lon. (m) (m) Start End (mg m2 d1) Labela Reference
Tasman Front Site 13 35.5 161 3174 1161 1 Jun 1995 1 Mar 1996 2.50 32-I Kawahata and Ohta [2000]f
Subtrop. Front, N Chatham Rise NCR 42.7 178.6 1500 300 9 Jun 1996 15 May 1997 10.1 Nodder and Northcote [2001]
Subtrop. Front, NChatham Rise NCR 42.7 178.6 1500 1000 9 Jun 1996 15 May 1997 20.6 Nodder and Northcote [2001]
Subtrop. Front, S Chatham Rise SCR 44.6 178.6 1500 300 9 Jun 1996 15 May 1997 4.11 9-S Nodder and Northcote [2001]
Subtrop. Front, S Chatham Rise SCR 44.6 178.6 1500 1000 9 Jun 1996 15 May 1997 4.93 10-S Nodder and Northcote [2001]
Subantarctic Zone JGOFS-AESOPS-MS1 53.0 174.7 5441 986 28 Nov 1996 24 Dec 1997 1.25 11-S Honjo et al. [2000]b
Polar Frontal Zone JGOFS-AESOPS-MS2 56.9 170.2 4924 982 28 Nov 1996 27 Jan 1998 4.92 12-S Honjo et al. [2000]b
Polar Frontal Zone JGOFS-AESOPS-MS2 56.9 170.2 4924 4224 28 Nov 1996 27 Jan 1998 1.71 34-D Honjo et al. [2000]b
Antarctic Circumpolar Current JGOFS-AESOPS-MS3 60.3 170.1 3958 1003 28 Nov 1996 27 Jan 1998 6.40 33-I Honjo et al. [2000]b
Antarctic Circumpolar Current JGOFS-AESOPS-MS4 63.1 169.9 2886 1031 28 Nov 1996 27 Jan 1998 6.03 34-I Honjo et al. [2000]b
Antartic Zone JGOFS-AESOPS-MS5 66.2 168.7 3016 1842 28 Nov 1996 3 Nov 1997 2.82 35-I Honjo et al. [2000]b
Antartic Zone JGOFS-AESOPS-MS5 66.2 168.7 3016 937 28 Nov 1996 27 Jan 1998 5.31 13-S Honjo et al. [2000]b
Ross Sea polynya C top 72.5 172.5 493 230 22 Jan 1990 22 Feb 1992 10.8 Dunbar et al. [1998]c
Ross Sea polynya C bot 72.5 172.5 493 443 22 Jan 1990 22 Feb 1992 2.78 Dunbar et al. [1998]c
Ross Sea polynya JGOFS-AESOPS-MS6 73.5 176.9 565 200 28 Nov 1996 27 Jan 1998 5.23 14-S Collier et al. [2000]
Ross Sea polynya JGOFS-AESOPS-MS7b 76.5 178.0 581 206 28 Nov 1996 27 Jan 1998 13.6 15-S Collier et al. [2000]g
Ross Sea polynya JGOFS-AESOPS-MS7a 76.5 178.0 581 465 28 Nov 1996 27 Jan 1998 28.9 16-S Collier et al. [2000]g
Ross Sea polynya JGOFS-AESOPS-MS7b 76.5 178.0 581 481 28 Nov 1996 27 Jan 1998 33.4 17-S Collier et al. [2000]g,h
Ross Sea polynya A top 76.5 167.5 719 230 14 Jan 1990 24 Apr 1991 13.7 Dunbar et al. [1998]
Ross Sea polynya A bot 76.5 167.5 719 669 13 Jan 1990 7 Feb 1992 11.1 18-S Dunbar et al. [1998]c
Ross Sea polynya B top 76.5 175.0 519 230 17 Jan 1990 1 Jan 1992 19.2 19-S Dunbar et al. [1998]c,g,i
Ross Sea polynya B bot 76.5 175.0 519 469 17 Jan 1990 9 Feb 1992 11.2 20-S Dunbar et al. [1998]c,g,i

6 of 26
Atlantic Ocean
Central Fram Strait FS-1 78.9 1.4 2823 2442 20 Aug 1984 15 Aug 1985 0.65 36-I Honjo et al. [1987]
E Fram Strait SP-2, SP-3 78.9 6.7 1669 1118 1 Jul 1988 5 Jul 1990 19.7 37-I Hebbeln [2000]c
N Water polynya, Baffin Bay S5A 76 78 365 258 Aug 1997 Jun 1999 26.9 Hargrave et al. [2002]c
N Water polynya, Baffin Bay S2B 76 73 561 511 Sep 1997 Jul 1999 30 Hargrave et al. [2002]c
N Norwegian Sea BI-1 75.9 11.5 2123 1700 12 Aug 1984 10 Aug 1985 4.60 38-I Honjo et al. [1987]
Greenland Basin GB-23 75.6 6.7 3445 2823 4 Aug 1985 20 Jul 1986 0.88 35-D Honjo et al. [1987]
E Greenland Sea OG7 75N 75.0 10.6 3073 500 23 Jul 1994 15 Aug 1995 10.7 Ramseier et al. [1999]
Greenland Sea OG 72.5 7.7 2700 500 Sep 1988 Jul 1991 7.83 21-S Bodungen et al. [1995]c,d
Greenland Sea OG 72.5 7.7 2700 1000 Sep 1988 Jul 1991 3.16 Bodungen et al. [1995]c
Greenland Sea OG 72.5 7.7 2700 2200 Sep 1988 Jul 1991 0.95 Bodungen et al. [1995]c
Norwegian Basin NB 70 0.4 3350 500 19 Feb 1992 19 Feb 1993 9.81 Bodungen et al. [1995]
LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP

Norwegian Basin NB 70 0.4 3350 1000 19 Feb 1992 19 Feb 1993 6.03 Bodungen et al. [1995]
E Lofoten Basin LB-1 69.5 10 3161 2760 15 Aug 1983 1 Aug 1984 2.21 36-D Honjo et al. [1987]
Voering Plateau VP 67.8 5.5 1300 525 Jun 1986 Oct 1987 6.66 22-S Bathmann et al. [1990]d
Aegir Ridge NA-1 65.5 1 3058 2630 21 Aug 1985 16 Jun 1986 1.62 37-D Honjo et al. [1987]
E Gotland Sea 57.3 20 na 140 1 Sep 1996 30 Jul 1997 16.8 23-S Schneider et al. [2000]d
NE Atlantic L3 54.6 21.2 3027 2200 10 Jun 1992 24 May 1994 2.15 Kuss and Kremling [1999]c
NE Atlantic L3 54.5 21.1 2979 2880 10 Jun 1992 12 May 1993 2.03 Kuss and Kremling [1999]
Mid-European cont. mar. OMEX-2 49.2 12.8 1450 600 1 Jul 1993 3 Sep 1994 5.75 Antia et al. [1999]
Mid-European cont. mar. OMEX-2 49.2 12.8 1450 1050 1 Jul 1993 3 Sep 1994 6.30 Antia et al. [1999]
Mid-European cont. mar. OMEX-3 49.1 13.4 3660 580 1 Jul 1993 12 May 1994 6.03 Antia et al. [1999]
Mid-European cont. mar. OMEX-3 49.1 13.4 3660 1450 1 Jul 1993 3 Sep 1994 10.1 Antia et al. [1999]
Mid-European cont. mar. OMEX-3 49.1 13.4 3660 3260 1 Jul 1993 3 Sep 1994 5.75 Antia et al. [1999]
Madeira Abyssal Plain PAP 48.8 16.5 4850 1000 12 Oct 1995 8 Mar 1998 2.60 24-S Lampitt et al. [2001]c
Madeira Abyssal Plain PAP 48.8 16.5 4850 3000 18 Apr 1989 26 Sep 1999 3.32 38-D Lampitt et al. [2001]c
C10011
Table 1. (continued)
Water Trap Collection interval
C10011

depth depth Corg


Region Trap ID Lat. Lon. (m) (m) Start End (mg m2 d1) Labela Reference
Madeira Abyssal Plain PAP 48.8 16.5 4850 4700 18 Apr 1989 29 Aug 1999 3.18 39-D Lampitt et al. [2001]c
NE Atlantic L2 47.8 19.7 4553 500 27 Mar 1992 12 Jun 1994 2.96 Scholten et al. [2001]
NE Atlantic L2 47.8 19.8 4551 1033 4 Apr 1993 28 Jun 1995 3.56 Kuss and Kremling [1999]; Scholten et al. [2001]c
NE Atlantic L2 47.8 19.8 4551 2017 10 Jun 1992 28 Jun 1995 5.11 Kuss and Kremling [1999]; Scholten et al. [2001]c
NE Atlantic L2 47.8 19.8 4553 3515 10 Jun 1992 28 Jun 1995 2.35 Kuss and Kremling [1999]; Scholten et al. [2001]c
North Atlantic Drift JGOFS-NABE-48N 47.7 20.9 4435 1110 3 Apr 1989 16 Apr 1990 4.05 39-I Honjo and Manganini [1993]c
North Atlantic Drift JGOFS-NABE-48N 47.7 20.9 4435 2109 3 Apr 1989 16 Apr 1990 3.78 40-I Honjo and Manganini [1993]c
North Atlantic Drift JGOFS-NABE-48N 47.7 20.9 4435 3734 3 Apr 1989 16 Apr 1990 2.74 Honjo and Manganini [1993]
North Atlantic Drift BOFS-48N 47.3 19.5 4555 3100 18 Apr 1989 16 Sep 1990 5.19 Newton et al. [1994]; Jickells et al. [1996]
NE Atlantic POMME-NE 43.5 17.3 3760 400 14 Feb 2001 25 Jun 2002 3.29 Guieu et al. [2005]
NE Atlantic POMME-NE 43.5 17.3 3760 1000 14 Feb 2001 25 Jun 2002 4.93 Guieu et al. [2005]
NE Atlantic POMME-SE 39.5 17.3 4940 400 14 Feb 2001 25 Jun 2002 6.30 Guieu et al. [2005]
NE Atlantic POMME-SE 39.5 17.3 4940 1000 14 Feb 2001 25 Jun 2002 3.29 Guieu et al. [2005]
E Alboan Gyre ALB-4-S 36.3 1.5 2240 645 1 Jul 1997 22 May 1998 26.4 Sanchez-Vidal et al. [2004]
E Alboan Gyre ALB-4-I 36.3 1.5 2240 1170 1 Jul 1997 22 May 1998 16.6 Sanchez-Vidal et al. [2004]
W Alboan Gyre ALB-1-S 36.2 4.26 1004 471 7 Jan 1997 22 May 1998 26.9 Fabres et al. [2002]
W Alboan Gyre ALB-2-S 36.0 4.29 1337 396 7 Jan 1997 22 May 1998 46.6 Fabres et al. [2002]
E Subtropical Gyre JGOFS-NABE-34N 33.8 21.0 5172 1159 3 Apr 1989 16 Apr 1990 2.74 41-I Honjo and Manganini [1993]c
E Subtropical Gyre JGOFS-NABE-34N 33.8 21.0 5172 1981 3 Apr 1989 16 Apr 1990 2.82 42-I Honjo and Manganini [1993]c
E Subtropical Gyre JGOFS-NABE-34N 33.8 21.0 5172 4478 3 Apr 1989 16 Apr 1990 2.03 40-D Honjo and Manganini [1993]c
NE Atlantic L1 33.1 22 5269 2010 20 Sep 1993 25 Sep 1995 1.82 Kuss and Kremling [1999]; Scholten et al. [2001]c
NE Atlantic L1 33.1 22 5269 4075 20 Sep 1993 14 Oct 1995 1.59 Kuss and Kremling [1999]; Scholten et al. [2001]c
Hatteras Abyssal Plain HAP 32.7 70.8 na 5400 7 Jun 1982 28 Apr 1983 2.09 Dymond and Lyle [1994]

7 of 26
W Sargasso Sea OFP 31.8 64.2 4500 500 29 Mar 1984 25 Jun 1998 4.1 25-S Conte et al. [2001]c,j
W Sargasso Sea OFP 31.8 64.2 4500 1500 29 Mar 1984 25 Jun 1998 2.4 43-I Conte et al. [2001]c,j
W Sargasso Sea OFP 31.8 64.2 4500 3200 6 Apr 1978 25 Jun 1998 1.7 41-D Deuser et al. [1981]; Conte et al. [2001]c,j
Madeira Abyssal Plain BOFS-31N 31.6 24.7 5440 4440 10 Dec 1985 27 Oct 1986 1.44 Lampitt [1992]; Jickells et al. [1996]
Canary Current ESTOC-CI 29.1 15.5 3600 900 25 Nov 1991 2 Sep 1994 1.37 26-S Fischer et al. [1996a]; Neuer et al. [1997]c,d
Canary Current ESTOC-CI 29.1 15.5 3600 3077 25 Nov 1991 2 Sep 1994 2.30 42-D Neuer et al. [1997]c,d
Subtropical Gyre E BOFS-28N 28 22.0 4820 3600 7 Oct 1990 28 Jul 1991 1.34 Jickells et al. [1996]
Subtropical Gyre E BOFS-25N 24.6 22.8 4860 3870 14 Oct 1990 27 Sep 1991 1.99 Jickells et al. [1996]
Nares Abyssal Plain NAP 23.2 64.0 na 5847 21 Aug 1983 27 Sep 1984 1.34 Dymond and Lyle [1994]
Cabe Blanc CB3, CB4 21.2 20.7 4100 732 8 Apr 1990 19 Nov 1991 7.1 27-S Fischer et al. [1996b]c
Cabe Blanc CB1 21.2 20.7 3646 2195 22 Mar 1988 8 Mar 1989 3.3 44-I Fischer et al. [1996b]; Fischer et al. [2000]
Cabe Blanc CB2, CB3, CB4 21.2 20.7 4100 3540 15 Mar 1989 19 Nov 1991 4.8 43-D Fischer et al. [1996b]; Fischer et al. [2000]c
LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP

Mauritanian upwelling zone EUMELI-O 21 31 4600 1000 19 Feb 1991 2 Dec 1992 1.8 28-S Bory et al. [2001]
Mauritanian upwelling zone EUMELI-O 21 31 4600 2500 23 Feb 1991 2 Dec 1992 1.5 45-I Bory et al. [2001]
E Atlantic Coastal-Boundary BOFS-19N 19 20.2 3295 2190 21 Oct 1990 2 Jun 1991 21.3 Jickells et al. [1996]
Mauritanian upwelling zone EUMELI-M 18.5 21 3100 1000 12 Feb 1991 2 Dec 1992 20.2 29-S Bory et al. [2001]
Mauritanian upwelling zone EUMELI-M 18.5 21 3100 2500 12 Feb 1991 2 Dec 1992 16.2 46-I Bory et al. [2001]
Mauritanian upwelling zone EUMELI-M 18.5 21 3100 3000 9 Oct 1991 2 Dec 1992 5.3 44-D Bory et al. [2001]
Cabe Verde CV1, CV2 11.5 21 4971 989 5 Oct 1992 5 Apr 1994 4.52 Fischer et al. [2000]c
Cariaco Basin A 10.5 64.7 1400 275 8 Nov 1995 9 May 2005 72.1 30-S Thunell et al. [2000]; Müller-Karger et al. [2004]c,k
Cariaco Basin B 10.5 64.7 1400 440 8 Nov 1995 9 May 2005 53.6 31-S Thunell et al. [2000]; Müller-Karger et al. [2004]c,k
Cariaco Basin C 10.5 64.7 1400 840 8 Nov 1995 9 May 2005 37.3 32-S Thunell et al. [2000]; Müller-Karger et al. [2004]c,k
Cariaco Basin D 10.5 64.7 1400 1255 8 Nov 1995 9 May 2005 30.6 47-I Thunell et al. [2000]; Müller-Karger et al. [2004]c,k
E Equatorial Atlantic GB2, GBN3, GBN6, EA2 1.8 11.2 4314 853 15 Mar 1988 13 Apr 1991 5.97 33-S Fischer et al. [2000]c
Guinea Basin GBN3-l, GBN6-l 1.8 11.1 4502 3921 1 Mar 1989 7 Apr 1991 5.48 45-D Fischer and Wefer [1996]c
W Equatorial Atlantic WA8 0.0 23.5 3744 718 25 Aug 1994 17 Aug 1995 6.82 Fischer et al. [2000]
C10011
Table 1. (continued)
Water Trap Collection interval
C10011

depth depth Corg


Region Trap ID Lat. Lon. (m) (m) Start End (mg m2 d1) Labela Reference
E Equatorial Atlantic EA3c, EA7, EA9, EA10 0.0 10.8 4385 1138 13 Apr 1991 27 May 1994 4.35 Fischer et al. [2000]c
E Equatorial Atlantic GBS4, GBS5 2.2 9.9 3916 696 1 Mar 1989 30 Mar 1991 5.62 34-S Fischer et al. [2000]c
Guinea Basin GBZ5-l 2.2 9.9 3920 3382 1 Apr 1990 30 Mar 1991 5.75 Fischer and Wefer [1996]
W Equatorial Atlantic WA1, WA4, WA7 4 25.7 5552 771 17 Oct 1992 17 Aug 1995 4.61 Fischer et al. [2000]
E Equatorial Atlantic EA5 4.3 10.3 3490 947 13 Apr 1991 29 Nov 1991 2.16 Fischer et al. [2000]
E Equatorial Atlantic EA8 5.8 9.4 3450 598 15 Dec 1991 6 Oct 1992 6.05 Fischer et al. [2000]
W Equatorial Atlantic WA2, WA3, WA6 7.5 28 5363 602 19 Oct 1992 17 Aug 1995 2.57 Fischer et al. [2000]c
Walvis Ridge WR1, WR2, WR3, WR4-l 20.0 9.2 2221 1663 4 Mar 1988 17 Dec 1991 10.1 Fischer et al. [2000]c
Walvis Ridge WR2-u 20.0 9.2 2196 599 18 Mar 1989 13 Mar 1990 14.0 35-S Wefer and Fischer [1993]
Namibia Upwelling NU2 28.9 14.6 3055 2516 20 Jan 1992 3 Feb 1993 4.41 Fischer et al. [2000]
Polar Front PF3, PF5, PF7, PF8 50.1 5.87 3795 3145 10 Nov 1989 15 Jan 1996 4.93 Fischer et al. [2002]c
Polar Front PF1, PF3, PF5, PF7, PF8 50.1 5.85 3787 658 15 Jan 1987 15 Jan 1996 3.84 Fischer et al. [2002]c
Bovert Island BO1, BO2, BO3, BO4, BO5 54.3 3.34 2728 487 28 Dec 1990 15 Jan 1996 2.19 Fischer et al. [2002]c
Bovert Island BO1, BO2, BO5 54.3 3.34 2729 2209 28 Dec 1990 15 Jan 1996 1.10 Fischer et al. [2002]c
N Weddell Sea Gyre VIII-u 62.1 40.6 3280 2453 1990 1992 0.61 Pudsey and King [1997]
Bransfield St., King George Is. KG1 62.3 57.5 1952 494 31 Dec 1983 25 Nov 1984 21.1 36-S Wefer et al. [1988]; Wefer and Fischer, [1991]
Bransfield St., King George Is. KG1 62.3 57.5 1952 1588 1 Dec 1983 25 Nov 1984 8.20 48-I Wefer et al. [1988]; Wefer and Fischer, [1991]
Bransfield St., King George Is. KG2, KG3 62.4 57.7 1992 687 4 Dec 1984 7 May 1986 3.01 37-S Wefer et al. [1988]; Wefer and Fischer, [1991]c,l
N. Weddell Sea Gyre I-u 63.2 42.7 3798 2971 1990 1992 0.24 Pudsey and King [1997]
Weddell Sea WS3 64.9 2.6 5053 360 16 Jan 1988 4 Feb 1989 6.25 38-S Wefer and Fischer [1991]
Weddell Sea WS2 64.9 2.5 5000 4456 20 Jan 1987 20 Nov 1987 0.47 46-D Wefer et al. [1990]; Wefer and Fischer [1991]

Indian Ocean

8 of 26
Arabian Sea MS-1 17.7 57.9 1447 808 11 Nov 1994 24 Dec 1995 10.5 39-S Honjo et al. [1999]b,c
Arabian Sea MS-1 17.7 57.9 1447 999 11 Nov 1994 24 Dec 1995 10.5 40-S Honjo et al. [1999]b,c
Bay of Bengal North-s 17.4 89.6 2263 809 1 Oct 1987 1 Oct 1988 9.84 Ittekkot et al. [1991]
Bay of Bengal North-d 17.4 89.6 2263 1727 1 Oct 1987 1 Oct 1988 7.26 Ittekkot et al. [1991]
Arabian Sea MS-2 17.4 58.8 1447 839 11 Nov 1994 24 Dec 1995 13.5 41-S Honjo et al. [1999]b,c
Arabian Sea MS-2 17.4 58.8 1447 914 11 Nov 1994 13 Sep 1995 17.2 42-S Honjo et al. [1999]b,c
Arabian Sea MS-2 17.4 58.8 3642 1985 11 Nov 1994 24 Dec 1995 17.4 49-I Honjo et al. [1999]b,c
Arabian Sea MS-2 17.4 58.8 3642 3150 11 Nov 1994 24 Dec 1995 13.2 47-D Honjo et al. [1999]b,c
Arabian Sea MS-3 17.2 59.6 3465 778 11 Nov 1994 24 Dec 1995 13.2 43-S Honjo et al. [1999]b,c
Arabian Sea MS-3 17.2 59.6 3465 873 11 Nov 1994 24 Dec 1995 17.5 44-S Honjo et al. [1999]b,c
Arabian Sea MS-3 17.2 59.6 3465 1870 11 Nov 1994 24 Dec 1995 16.3 50-I Honjo et al. [1999]b,c
Arabian Sea MS-3 17.2 59.6 3465 2979 11 Nov 1994 24 Dec 1995 12.8 48-D Honjo et al. [1999]b,c
LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP

Arabian Sea WAST 16.5 60.5 4016 3027 10 May 1986 15 Oct 1990 8.77 Haake et al. [1993]
Arabian Sea EAST 15.5 68.7 3785 2830 10 May 1986 23 Oct 1990 5.75 Haake et al. [1993]
Arabian Sea MS-4 15.3 61.5 3974 814 19 Nov 1994 24 Dec 1995 8.9 45-S Honjo et al. [1999]b,c
Arabian Sea MS-4 15.3 61.5 3974 2222 11 Nov 1994 27 Aug 1995 11.1 51-I Honjo et al. [1999]b,c
Arabian Sea MS-4 15.3 61.5 3974 3484 11 Nov 1994 7 Dec 1995 8.9 49-D Honjo et al. [1999]b,c
Arabian Sea CAST 14.9 64.1 3904 2954 10 May 1986 24 Feb 1989 5.21 Haake et al. [1993]
Bay of Bengal Central-s 13.2 84.4 3259 906 Oct 1987 Oct 1988 7.23 Ittekkot et al. [1991]
Bay of Bengal Central-d 13.2 84.4 3259 2282 Oct 1987 Oct 1988 7.15 Ittekkot et al. [1991]
Arabian Sea MS-5 10.0 65.0 4411 2363 28 Nov 1994 24 Dec 1995 3.8 52-I Honjo et al. [1999]b
Arabian Sea MS-5 10.0 65.0 4411 3915 28 Nov 1994 24 Dec 1995 3.3 50-D Honjo et al. [1999]b
Bay of Bengal South-s 4.43 87.3 4017 1040 Oct 1987 Oct 1988 6.49 Ittekkot et al. [1991]
Bay of Bengal South-d 4.43 87.3 4017 3006 Oct 1987 Oct 1988 5.59 Ittekkot et al. [1991]
Southern Ocean ANTARES-M2 52 61.5 4600 1300 13 Feb 1994 31 Jan 1995 2.88 53-I Tréguer, unpub.m
Southern Ocean ANTARES-M2 52 61.5 4600 4025 13 Feb 1994 31 Jan 1995 1.56 51-D Tréguer, unpub.m
C10011
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

deep ocean. On the basis of the geographic and depth


distribution of sediment trap observations, we apply a
first-order approximation of the export zone depth as 100
m at equatorial latitudes (<35°) increasing linearly to 400 m
at high-latitudes (>50°).
[14] (2) Observations are minimally influenced by resus-
pension of benthic surface sediment, typically those from
Reference

depths above the nepheloid layer.


[15] (3) Terrestrial sources of detritus are minimal.
[16] (4) Observations must describe at least one entire
Pilskaln et al. [2004]

year of flux to depth. This includes trap experiments of less


Trull et al. [2001]
Trull et al. [2001]

Tréguer, unpub.m
Tréguer, unpub.m

than a year in duration if the seasonal cycle of flux may


be well approximated. The sediment trap data set assembled
for this study includes 244 annual flux estimates and
153 subannually resolved flux time series used to create
flux climatologies.
[17] Flux climatology construction involves multiple
(mg m2 d1) Labela

52-D
46-S
54-I
55-I

Discrete open-close interval flux to depth data not available in numerical format. Records of flux to depth are digitized from data presented in graphs.

steps depending on the temporal coverage of trap measure-


ments. First, time series are made continuous for the entire
duration of the experiment. Gaps produced during sampling
Excluding KG2 10 Dec. 1985 to 24 Dec 1985 interval of overlap between KG2 and KG3 trap deployments to avoid attenuating seasonality.

bottle failures or between trap deployments are filled with


2.19
1.37
2.47
0.30
0.44
Corg

the linear integral of temporally adjacent observations.


Next, to account for timing inconstancies between trap
experiments, time series are temporally standardized. Tem-
1998
1998
1999
1995
1995

poral standardization involves a method of describing


End

sediment trap measurements that is a combination of the


Collection interval

21 Feb
21 Feb

8 Feb
8 Feb
30 Dec

Data available online (http://www.imars.usf.edu/CAR); F. Müller-Karger, R.Varela, R. Thunell, M. Scranton, and others.

regularly applied techniques.


[18] Authors typically depict sediment trap experiments
Flux corrected for a component potentially associated with pteropod ‘swimmers’ following Collier et al. [2000].

using linear connections between observation midpoints


1997
1997
1998
1994
1994

Climatology open and close dates are estimated as the first days of the month of reported monthly averages.

[Bathmann et al., 1990; Neuer et al., 1997; Honjo et al.,


Data available online (http://usjgofs.whoi.edu/mzweb/data/Honjo/sed_traps.html); S. Honjo and R. Francois.
Start

Data available online (http://www.obs-vlfr.fr/cd_rom_dmtt/ANTARES/A3/a3_parameters.htm); P. Tréguer.


30 Dec
21 Sep
21 Sep

20 Feb
20 Feb

Excluding 10 Feb. 1990 to 8 Feb. 1992 interval of overlap between traps to avoid attenuating seasonality.

1999; Hargrave et al., 2002]. Flux is also frequently


reported in bar-graph format as the average rate during
the open-close intervals of each bottle [Newton et al., 1994;
depth depth

1580
1400
1300
3445
Water Trap

830
(m)

Honjo et al., 2000; Lampitt et al., 2001; Honda et al.,


2002]. Our approach, demonstrated in Figure 3, is a
Due to local cloud cover, remote sensing data used are from 2700 m north of location.
2280
2280
4000
4020
4020

smoothing of the linear style held equal in mass collected


(m)

per bottle interval to the bar-graph style. Dashed lines


Organic carbon estimated as (1.8 * OC = OM) following Kawahata et al. [1998].

connecting bottle midpoints (closed circles) and grey


141.8
141.8
Lon.

73.0
71
71

columns indicate the linear and bar-graph data representa-


tion styles. Open circles designate temporal intersections of
53.8
53.8
62.4

the linear style with sample bottle openings (xo) and clos-
63
63
Lat.

ings (xc). The solid thin line l connects these intersections.


The curve D indicates the ‘‘stretching’’ of line l such that the
integral of D equals the area represented by the bar-graph
Data available online (http://www.whoi.edu/OFP); M. Conte.

style for each bottle:


Multiple trap deployment flux estimates are combined.
Trap ID

Zxc
ANTARES-M3
ANTARES-M3

ð DÞdx ¼ Cflux ; ð3Þ


Labels refer to flux time-series in Figure 8.

xo
PFZ54
PFZ54
PZB-1

where Cflux denotes the average flux rate. The stretch


function D is accomplished by multiplying line l by a daily
cubic spline interpolation of normal distribution constrained
so that the endpoints (open circles) remain constant. The
direction line l stretches to become curve D, up, down, or
Table 1. (continued)

not at all, depends on the flux rate measured in each bottle


Polar Frontal Zone
Polar Frontal Zone
Region

relative to the rates of neighboring bottles. Minimum daily


Southern Ocean
Southern Ocean

flux rates estimated by the stretch function D are


Prydz Bay

constrained not to be negative. Finally, sediment trap


measurements including more than 1 year of observations
are averaged into one climatological year.
m
b

g
h

k
a

e
f

i
j

9 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

[Gust et al., 1994; Buesseler et al., 2000], and influences in


brine and poison addition [Lee and Cronin, 1982; Knauer et
al., 1984]. A standard trap design and methodological
protocol is not adopted by the oceanographic community.
The magnitude of these and other trap biases may be highly
variable and as large as a factor of two or more [Gardner,
2000]. Calibration of sediment trap results have been
attempted [Buesseler et al., 2000; Yu et al., 2001; Scholten
et al., 2001]. In defense of sediment trap technology, we
note that sediment trap results: (1) are interannually consis-
tent with respect to timing, variability, and magnitude of
particle fluxes [Deuser, 1986; Conte et al., 2001]; and
(2) reflect euphotic zone biological variability [Deuser
et al., 1981; Honjo, 1982; Deuser et al., 1990]. We use
sediment trap data sets here because they represent the only
available direct measurements of annual subsurface particle
fluxes through the water column.
[21] To address potential error of sediment trap observa-
tions, estimates global flux to depth include the calibration
Figure 3. Demonstration of the stretch function ((D)dx) rationale suggested by radionuclide studies. The validity of
curve fit technique developed to create sediment trap sediment traps to measure flux to the deep ocean (>1.5 km)
particulate organic carbon flux to depth climatologies. Time is validated by 230Th and 231Pa calibration studies [Scholten
series standardization involves a combination of the et al., 2001; Yu et al., 2001]. Within the upper ocean
regularly applied description techniques: a smoothing of (<1500 km) these radionuclide studies suggest sediment
the linear midpoint connection style (dashed lines and .) traps may often underestimate fluxes. With the exception of
held equal in mass to the bar-graph format (grey columns). the California margin, Yu et al. [2001] found a trapping
Open circles (6) designate temporal intersections of the efficiency of 40% is a typical minimum value for the
linear style with sample bottle openings (xo) and closings pelagic upper ocean. To account for the potential under-
(xc). The solid thin line l shows linear connects these trapping error, we report global estimates of flux to the
intersections. The curve D indicates ‘‘stretching’’ of line l so upper ocean with (observed flux divided by 0.4) and
that the integral of D equals the area represented by the bar- without radiochemical calibration. This trapping efficiency
graph style for each bottle. The direction line l stretches to estimate may be uncertain in part because of the variable
become curve D, up, down, or not at all, depends on the flux incorporation of radionuclides on particles of different sizes
rate measured in each bottle relative to that of its [Buesseler et al., 1995; Yu et al., 2001].
neighboring bottles. [22] Satellite-based estimates of NPP also include sig-
nificant uncertainties [Platt and Sathyendranath, 1993;
4.2. NPP and SST Behrenfeld et al., 2002a]. These uncertainties largely stem
[19] NPP is estimated following Behrenfeld and Falkowski from errors in relating surface chlorophyll biomass to
[1997a] using 7 years of 8-day satellite images available phytoplankton carbon biomass [Behrenfeld et al., 2005].
between 19 August 1997 and 24 June 2004. Data sets applied Furthermore, empirical models used to estimate NPP
include the NOAA/NASA AVHRR Oceans Pathfinder SST from ocean color may not correctly simulate relevant
(http://podaac.jpl.nasa.gov/products/product216.html), phytoplankton physiological variability [Behrenfeld and
NASA SeaWiFs surface chlorophyll concentrations, and Falkowski, 1997b; Behrenfeld et al., 2002b; Banse and
photosynthetically active radiation (http://daac.gsfc.nasa. Postel, 2003]. Ocean color is a property of the uppermost
gov/data/dataset/SEAWIFS/). Images applied display global portion of euphotic zone and additional factors are
coverage with an equal-area grid of 9-km resolution. NPP needed to estimate deeper ocean production. Finally,
climatologies are constructed in manner identical to that of field-based radiocarbon measurements of production used
flux climatologies. During climatology construction, missing to calibrate satellite-derived algorithms include error
production estimates, for example, those obscured by cloud [Sakshaug et al., 1997]. A recent performance analysis
cover, are substituted with the linear interpolation of adjacent indicates various NPP algorithms yield significantly dis-
values. similar results [Campbell et al., 2002]. We acknowledge
algorithms describing production using satellite data may
4.3. Potential Error have significant systematic biases.
[20] Errors associated with methods used to estimate [23] Specific estimates of error regarding satellite-based
marine biogeochemical processes are difficult to assess. NPP and sediment trap flux are not easily quantified and
Field-based oceanography is far from a controlled well- such an exercise is generally avoided in the literature.
calibrated laboratory experiment. The ability of sediment There is no a priori reason to believe errors associated
traps to accurately measure flux has received much attention with sediment trap observations covary with errors asso-
in the scientific literature. Possible trap biases include ciated with satellite-based observations. Thus we infer that
inadvertent ‘‘swimmer’’ capture [Lee et al., 1988], uncertain statistical significance in the results described below is
preservation of trapped particulate matter [Lee et al., 1992], found in spite of, and not because of, error in the data sets
hydrodynamic interactions with traps of various designs used.

10 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

4.4. Seasonal Variation Index seasons of flux to corresponding seasons of production. We


[24] We use the seasonal variation index (SVI) to describe acknowledge this technique allows for a first-order descrip-
temporal variability of the time series. SVI is defined as the tion of production and flux dynamics, as sinking rates may
coefficient of variation, the standard derivation (s(X)) differ between seasons, regions, and depths [Berelson,
normalized to the average (X ), of either the NPP or flux 2002]. Due to the lack of polar wintertime production,
climatologies: winter season p ratios are absent at latitudes >60°.
4.6. Flux Algorithm With Labile and Refractory
sðX Þ
SVI ¼ : ð4Þ Components
X
[28] To quantify relationships between satellite-based
NPP and SST data, and sediment trap flux observations,
[25] SVI is dimensionless. This description of variability annual production ratios are estimated using the following
is statistically similar to the seasonality index, the number of exponential algorithm:
months required to accumulate one half of the total annual
primary production [Berger et al., 1989; Berger and Wefer,  
ze
1990], and the flux stability index, the minimum time taken p ratioðze Þ ¼ prd exp þ prr ð6Þ
rld
for 50% of the annual flux to be collected [Lampitt and
Antia, 1997].
[29] This algorithm was previously derived from an
4.5. Annual and Seasonal Production Ratios empirical fit between regional production and flux measure-
[26] Previous research has compared e ratios, the ratio of ments [Lutz et al., 2002] and is applied here to determine if
flux to export production, to examine the transfer efficiency flux can be estimated using environmental parameters
of POC [Laws et al., 2000; Francois et al., 2002]. A recent derived from satellite observations. This function describes
study shows transfer efficiency described by flux normal- the labile and refractory components of flux to depth below
ized to either production or export yield results of similar the export zone depth ze [depth (z) minus export zone
accuracy [Lutz et al., 2002]. Here following the approach of depth]. The prd and rld coefficients measure flux that is
Suess [1980], we compare the transfer efficiency between available to decay and sinks more slowly. The rld coefficient
regions using annual production (p) ratios, defined as the describes the e-folding remineralization length/depth scale
annual particle flux at depth (z) normalized to annual NPP and measures the rate of sinking detritus degradation. The
in the overlying surface waters: prr coefficient describes the more refractive and more
rapidly sinking portions of flux. Larger prd and prr values
CfluxðzÞ indicate the export of a greater fraction of NPP. Larger rld
p ratioðzÞ ¼ ð5Þ values indicate the labile fraction of export sinks deeper
Cnpp
before regeneration. For simplicity, this equation treats
This study also uses seasonal p ratios to examine subannual sinking and remineralization rates as constants, although
production-to-flux dynamics. Seasonal p ratio creation we acknowledge these factors vary as particles are trans-
involves several steps and is conducted at locations where formed during their descent through the ocean [Berelson,
an entire year of flux time series is available. First, seasons 2002]. For further derivation of equation (6) see Banse
of production are delineated by dividing NPP climatologies [1990] and Lutz et al. [2002].
into autumn, winter, spring, and bloom intervals. The bloom [30] Annual p ratio estimates are used to determine the
production interval describes the average production during coefficients of equation (6) grouped by either SST or SVI.
the 30 continuous days of an annual record with the Coefficient approximation requires the p ratio data be
maximum average production rates. Similarly, the winter divided into groups so that the entire water column is
production interval describes the average production during described for each SVI and SST value evaluated. To best
the 30 continuous days with the minimum average characterize SVI and SST variability, groups used herein are
production rates. Autumn and spring production intervals of equal spacing on a logarithmic-scale for SVI and a linear-
describe average productions during the intervening time scale for SST. The prd, rld, and prr coefficients are deter-
periods between bloom and winter seasons. The durations mined for each group separately. Within each group the prr
of autumn and spring production intervals varies depending coefficient describes the median of the deepest 25% of
on the timing of the bloom and winter seasons, and are p ratios, the depth of which vary slightly between groups.
constrained to be of at least one month in duration. By this Thus for the purpose of this comparison it is assumed that
constraint, 134 time series of production are included in the on an annual basis flux to the deepest traps is refractory
seasonal p ratio analysis. relative to upper ocean traps. The prd coefficient describes
[27] Next, the timings of corresponding sediment trap the maximum shallow p ratio minus the prr coefficient. The
flux seasons are determined by adding a temporal lag to the rld coefficient is determined by fitting equation (6) to the p
timing of the production seasons. The temporal lag is ratio data with the previously determined prd and prr
defined by dividing the trap depth (m) by a sinking rate coefficients. For each coefficient the maximum number of
(m d1). Here we use a sinking rate of 70 m d1 based on groups is used that yields a statistically significant relation-
the 1.5-month lag between surface events observed by ship to either SVI or SST. Hence the number of groups for
satellite and arrival of the consequences of those events in the prd and prr parameters differs depending on the avail-
sediment traps at 3.2 km shown by Deuser et al. [1990]. ability of the shallow and deep trap data within each group.
Finally, seasonal p ratios are determined by normalizing Similarly the grouping for the rld parameter requires a
sufficient number of observations to characterize the entire

11 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 4. The geographic distribution of the seasonal variation index (annual standard deviation
divided by average) of net primary production (NPP). This estimate is derived from NPP climatology
modeled following Behrenfeld and Falkowski [1997a] and constructed from 7 years of satellite-based
estimates of sea surface temperature and chlorophyll. Arrows indicate locations of the meridional
transects of NPP time series shown in Figure 5. Missing data (dark grey) indicates landmasses and
permanent ice cover.

water column. Due to the infrequency of sediment trap [33] The heterogeneous distribution of SVI values at
observations near the local export zone depth, prd values are lower-latitudes indicates provinces where interactions of
deemed minimal approximations. annual climatic phenomenon with the density structure of
upper ocean waters permit seasonal upwelling. In open
5. Results ocean waters, diminished SVI values characterize subtropi-
cal regions of year-round high surface atmospheric pressure
5.1. Patterns of Remotely Sensed Parameters,
and the seasonal limits of the intertropical convergence
NPP, and SST
zone. Reduced SVI values characterize the oligotrophic
[31] The results of our NPP climatology generally agree subtropical ocean gyres and equatorial western Pacific and
with those described by Behrenfeld and Falkowski [1997b]. Atlantic Oceans, where production is relatively constant
Here we briefly describe the first-order characteristics of year-round. Narrow longitudinal bands of moderate SVI
production temporal variability limited to those relating to values resulting from seasonal water mass divergence
patterns shown by our flux to depth climatology. characterize the equatorial upwelling regions of the Pacific
[32] The NPP climatology, as described by the SVI and Atlantic Oceans. Elevated SVI values approximate the
(Figure 4), displays regional- and basin-scale geographic influence of the circumpolar westerlies. Monsoonally influ-
patterns. Temporal trends of the NPP climatology (Figure 5) enced regions of the western Indian Ocean exhibit enhanced
are similar to those found in the literature [Yoder et al., SVI values. Larger SVI values typically characterize the
1993; Longhurst, 1995] and include the well-known open centers of eastern boundary seasonal coastal upwelling and
ocean equatorial near-constant low productivity, the multi- the associated offshore areas where upwelled nutrients are
ple blooms of temperate and monsoonal regions, and the advected. Additionally, SVI values reflect the seasonal input
intense summer bloom of polar regions. A primary compo- of riverine and aeolian nutrients, as well as limitations to
nent of global distribution of NPP SVI is the increase of production, such as mixing of surface waters beneath the
values poleward of roughly 45°S and N, reflecting the photic zone, phytoplankton self-shading, and herbivore
seasonal availability of photosynthetically available radia- activity [Parsons et al., 1977; Broecker and Peng, 1982;
tion and distribution of sea ice. Chester, 1990; Duce and Tindale, 1991].

12 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 5. Climatologies of net primary production (NPP; g Corg m2 yr1) versus latitude. Shown are
time series of NPP located along meridional transects, indicated in Figures 1 and 4, within the (a) central
Pacific (134°W), (b) Atlantic (29°W), and (c) Indian (65°E) Oceans. This estimate uses NPP modeled
following Behrenfeld and Falkowski [1997a] and constructed from 7 years of satellite-based estimates of
sea surface temperature and chlorophyll. Missing data (dark grey) indicates landmasses and permanent
ice cover.

[34] The Southern Hemisphere of the Pacific Ocean [35] For much of the global ocean the SVI of produc-
indicated by transect A of Figures 4 and 5 epitomizes the tion and SST show similar trends (Figures 6 and 7).
transition from low-amplitude seasonality and diminished Poleward of 35° latitude, where waters are cooler than
production of equatorial regions to high-amplitude season- approximately 18°C, SVI, and SST generally covary. At
ality and strong summer blooms of polar regions. Heading lower-latitudes the distribution of SVI and SST varies
south from the equator, between 5° and 15°S, SVI values depending on regional oceanographic circumstances and
decrease as the strength of equatorial upwelling diminishes. differs significantly between ocean basins. Globally,
Continuing south, a latitudinal ribbon of elevated SVI neither variability of production nor SST correlates sig-
values, between 25° and 35°S, indicates where mixing of nificantly with rates of NPP.
subsurface waters increases surface nutrient levels sufficient
to generate a summer bloom. Adjacently poleward, a 5.2. Patterns of Sediment Trap Flux
narrow latitudinal band of minimum SVI values, between [36] The results of the flux to depth climatology are shown
35° and 40°S, indicates where upper water column mixing in Figure 8. Here we focus on regional- and basin-scale
sustains elevated production year-round, without distinct patterns of flux rather than restate findings of original
spring and fall blooms. Farther poleward SVI values in- sources. Global sediment trap sampling is heterogeneous
crease as bloom intensity amplifies. The latitude-dependent in depth and geographic distribution. Often, the water
SVI fluctuations and the northeast trending ‘‘sustained- column is not entirely measured, leaving either shallow,
bloom’’ band of diminished SVI values distinguish the intermediate, or deep waters uncharacterized. Sediment trap
influence of the westerlies, and are less distinctly repeated observations are heavily weighted, by 70%, toward the
at similar latitudes in the western and eastern portions of the Northern Hemisphere. Major portions of the Southern
Northern and Southern Hemispheres, respectively. Hemisphere are uncharacterized, including the deep Antarc-

13 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 6. Latitude versus (a) the Seasonal Variation Index (annual standard deviation divided by
average) of net primary production and (b) sea surface temperature (SST) for the Pacific, Atlantic, and
Indian Oceans. Circles represent data at 1.5° latitude and longitude intervals.

tic circumpolar waters of Atlantic and Indian Oceans. The


Indian Ocean has received particularly little attention, with
less than an eighth of all observations. Large segments of
continental shelf regions are unknown, such as surrounding
South America and much of Antarctica, and bordering the
Indian Ocean. Meagerly characterized open ocean regions
include the equatorial upwelling of the Atlantic and Indian
Oceans, and the moderately productive Southern Hemi-
sphere subtropical convergence. Regions of mesotrophic
productivity have received the majority of attention. The
most productive centers of upwelling regions are not well
characterized. The modestly productive oligotrophic central
ocean gyres are similarly not frequently sampled. Hence
regions characterizing minimum and maximum limits of
global productivity rates largely await description.
[37] Recognition of basin-scale seasonal cycles of flux to
depth is limited by the geographic heterogeneity of sam-
pling and intermittent high-frequency temporal variability
of measurements. Nevertheless, the flux climatology compi-
lation includes five major patterns of seasonality (Figure 8)
similar to those of NPP: polar, equatorial and coastal
upwelling, monsoonal, and subtropical-subpolar. Cycles of
enhanced summertime flux and diminished wintertime flux Figure 7. The seasonal variation index (annual standard
characterize the poles [e.g., Honjo et al., 1987; Bathmann et deviation divided by average) of net primary production
al., 1990; Wefer and Fischer, 1991; Bodungen et al, 1995; versus seas surface temperature (SST) for the Pacific,
Dunbar et al., 1998; Honjo et al., 2000; Schneider et al., Atlantic, and Indian Oceans. Circles represent data at
2000; Takahashi et al., 2000; Wong et al., 1999; Honda 1.5° latitude and longitude intervals.

14 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 8. Climatologies of particulate organic carbon flux to depth (mg Corg m2 d1) derived using
sediment traps versus latitude. Horizontal contours indicate each sediment trap time series of flux to
various depth ranges (shallow 1 km; intermediate >1 and 2.5 km; deep >2.5 km) within the Atlantic,
Pacific, and Indian Oceans. Circles represent the climatological year-day of each sediment trap bottle
observation. Sediment trap experiments are identified in Table 1 by labels located on the right of each
time series.

et al., 2002; Collier et al., 2000; Pilskaln et al., 2004; and production hiatus and summer blooms, respectively. In
additional references in Table 1 and Figure 8]. Polar general, summertime fluxes to shallow, intermediate, and
seasonality is greater than of other regions, with fluxes deep-water depths are 20, 10, and 5 times wintertime fluxes.
ranging from <0.5 to >100 mg Corg m2 d1 during winter Atypical wintertime flux peaks of some polar observations

15 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 9. (a) Seasonal production ratios describing the particulate organic carbon fraction of production
that sinks beneath surface waters associated with each season of production. (b) Seasonal production
ratios normalized to the annual production ratio at each location. Values to the left and right of the dashed
vertical line indicate the factor to which the seasonal p ratios are less or greater than, respectively, the
annual p ratio.

(15 and 19 s) may be due to sediment resuspension [Dunbar et al., 1981; Conte et al., 2001; 42-d, Neuer et al., 1997].
et al., 1998; Collier et al., 2000]. This pattern is also apparent in deep waters of the Pacific
[38] Equatorial seasonality is low and varies little with Ocean [e.g., Figure 8: 9-d and 10-d, Honda et al., 2002; 12-d,
depth [e.g., Honjo et al., 1995; Dymond and Collier, 1988; 15-d, and 16-d, Kawahata et al., 2002; 13-d and 14-d,
Kemp and Knaack, 1996; Kawahata, 2002]. Seasons of Mohiuddin et al., 2004]. Overall, flux peaks are delayed by
maximum flux are generally double the rate of seasons of approximately five days per degree latitude increase. Maxi-
minimum flux. Monsoonal seasonality is moderate and mum flux timings are generally synchronous with the timing
characterized by the multiple flux maxima of the southwest of temperate spring blooms of lower latitudes and summer-
and northeast monsoons [Honjo et al., 1999]. The range of time polar blooms of higher latitudes, with a production-to-
rates between seasons of maximum and minimum flux is flux lag typically between 40 and 80 days. The range of flux
typically a factor of four, with minor attenuation of seaso- rates between season’s maximum and minimum flux is
nality with increasing depth. Regions influenced by coastal roughly a factor of six.
upwelling show relatively enhanced seasonality [e.g., Neuer
et al., 1997; Fischer and Wefer, 1996; Hebbeln et al., 2000; 6. Discussion
Bory et al., 2001]. Maximum fluxes to shallow, interme-
6.1. Limitations Using Seasonal Production Ratios
diate, and deep-water depths are typically 13, 9, and 4 times
minimum fluxes. [40] Sediment trap measurements of flux to depth reflect
[39] Flux seasonality of subtropical-subpolar open ocean variability of surface water production whereby periods of
waters is characterized by a latitude trending pattern, enhanced production generally correspond to periods of
whereby maximum fluxes occur later in the year at enhanced flux to depth [Deuser et al., 1981; Honjo, 1982;
higher-latitudes. This pattern is coherent across multiple Deuser et al., 1990]. We use seasonal production ratios
sediment trap time series, recognizable in the Atlantic (section 4.5) to characterize global-scale subannual variabi-
Ocean in intermediate [e.g., Figure 8: 40-i, 41-i, and 42-i, lity of particle transport efficiency. Inferences using seasonal
Honjo and Manganini, 1993] and deep waters [e.g., Figure 8: production ratios are limited due to methodological and
36-d, Honjo et al., 1987; 38-d and 39-d, Lampitt et al., natural factors. The seasonal interval technique does not
2001; 40-d, Honjo and Manganini, 1993; 41-d, Deuser completely describe subannual variability of production
and flux to depth. For example, dual production blooms

16 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 10. (a) Statistical distribution of seasonal production ratios grouped by depth. Horizontal black
lines designate median values and shaded boxes designate the distribution range 25 and 75% quartiles.
(b) Statistical distribution of seasonal production ratios normalized to the annual production ratio at each
location. Values above and below the dashed horizontal lines indicate the factor to which the seasonal
p ratios are greater or less than, respectively, the annual p ratio. Seasonal p ratio estimates shown are
grouped into the following depth ranges: 1 km wide groups between 0 and 4 km, and >4 km. Quartile
distributions are derived using data shown in Figure 9.

and flux events of monsoonal regions [Honjo et al., 1999] are may further obscure interpretations. Furthermore, the poten-
not completely characterized. Furthermore, the technique is tial inaccuracy of shallow sediment traps may limit the
limited by interannual differences between the timing of interpretation of shallow p ratio estimates. This analysis
production and flux to depth time series used to compile represents an initial step toward characterizing subannual
the climatologies. For example, where production and flux production-to-flux variability.
are influenced by long-term climate phenomenon, such as
ENSO [Kawahata and Gupta, 2004], some bloom produc- 6.2. The Seasonal Sinking Fraction of NPP
tion and flux intervals may be mismatched. Differences in [41] Our results indicate the sinking fraction of produc-
particulate matter sinking rates between seasons and depths tion is not seasonally constant. The global range of seasonal

17 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 11. Latitude versus the seasonal variation index (annual standard deviation divided by average)
of particulate organic carbon flux to various depth ranges [1 km (6), >1 and 2.5 km (+), and >2.5 km
(4)] and the seasonal variation index of net primary production (.) at sediment trap locations.

production ratios (Figures 9a and 10a) spans four orders sinking particulate matter to reach deeper waters. Dimi-
of magnitude, from 0.0003 to 0.93, with much overlap nished p ratios during bloom production relative to other
between seasons. This seasonal range is almost two orders seasons is consistent with a portion of bloom-derived
of magnitude greater than the range of annual p ratios particulate matter being retained to be recycled or fluxed
between different ocean regions (from about 0.001 to 0.1) during latter seasons. Multiseasonal retention and recycling
[Lutz et al., 2002]. A trend toward lower p ratios during the of detritus within the deep ocean is suggested to explain
bloom seasonal interval is indicated; however, an overlap seasonal patterns of suspended particulate matter concen-
between seasonal estimates makes differentiation between trations in the northeast Pacific [Bishop et al., 1999].
seasons difficult. Much of the global variability in seasonal
p ratios reflects differences between regions that exhibit 6.3. Seasonality of NPP and Flux to Depth
different ranges of NPP and different efficiencies of the [44] The balance between seasonality of production and
biological pump [Lutz et al., 2002; Neuer et al., 2002]. In seasonality of flux reverses with latitude (Figure 11). At
order to reduce the influence of variability between loca- higher-latitudes seasonality of production is generally greater
tions and changes between depths, seasonal p ratios are than seasonality of flux. At lower-latitudes seasonality of
normalized to local annual p ratios (Figures 9b and 10b). production is generally less than seasonality of flux. To
Normalized p ratios during bloom intervals are on average account for this latitudinal difference we suggest that
half of those during the rest of the year. The greatest range processes influence flux seasonality discernable at lower-
of normalized p ratios and the bulk of the lowest normalized latitudes, where variability of production is diminished, are
p ratios occur during the bloom season at <1 km. masked by the larger-amplitude production seasonality
[42] The production ratio results indicate a global-scale signal of higher latitudes. Processes that may enhance
synchronicity in the seasonal functioning of the biological low-latitude flux seasonality relative to production include
pump. This pattern may reflect a seasonal change in the phytoplankton mass sedimentation events [Kemp et al.,
biodegradability of sinking particulate matter. Diminished 2000] and transient meteorological forcing of pulsed export
bloom p ratios are consistent with a greater proportion of [Conte et al., 2003]. Processes that may attenuate season-
particulate matter remineralized during enhanced produc- ality between production and flux at high-latitudes include
tion. Increases in flux lability in response to increased the regeneration of detritus retained within surface waters,
production is suggested by previous flux studies involving delayed herbivore growth, and seasonal mixed-layer deep-
sediment traps [Lee and Cronin, 1984; Lohrenz et al., 1992; ening associated with polar winter onset. Overall, the
Haake et al., 1993; Newton et al., 1994; Lampitt and Antia, production-to-flux seasonality signal may be influenced
1997]. Changes of biodegradability may be due to the by horizontal water mass advection [Siegel and Deuser,
seasonal manufacture of skeletal material by phytoplankton. 1997], which may covary with rates of production [Lampitt
For example, growth forms of many Antarctic diatoms and Antia, 1997].
progress seasonally from long, lightly silicified chains 6.4. Parameterization of the Annual Sinking Fraction
(more labile) in the austral spring, to heavily silicified of NPP
valves (more dense and refractory) in the autumn and winter
[El-Sayed and Fryxell, 1993]. [45] Our analysis of SST- and SVI-associated hypotheses
[43] The pattern of seasonal production ratio may addi- includes determining relationships between the remotely
tionally reflect particulate matter retention or the delay of sensed parameters and production ratio data, and comparing

18 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 12. Coefficients used to forecast the annual average particulate organic carbon flux to depth
normalized to overlying production [p ratio(ze)] as a function of satellite-derived parameters, the seasonal
variation index of production (SVI; annual standard deviation divided by average) and sea-surface
temperature (SST; °C). Coefficients indicate the components of flux using equation (6): the labile fraction
of export (prd; a and b), remineralization length scale (rld; c and d), and more refractory and rapidly
sinking fraction of export (prr; e and f). Curve fit coefficient algorithms are reported in Table 2.

the ability of algorithms derived to predict flux. Flux alization in a stepwise manner with little influence until a
predictions are based on algorithms that describe the coef- certain low threshold temperature is reached, where flux is
ficients of equation (6) as a function of the remotely sensed more labile as indicated by the prd and rld coefficients.
parameters, as outlined in section 4.6. Coefficient parame- Alternately, this behavior may reflect SST and SVI covari-
terization involves testing different equation functional ance, characteristic of lower temperatures (Figure 7). SVI
forms (polynomial and exponential), various orders of coefficients and p ratio data groups show somewhat more
equations, and different numbers of data groups to find continuous distributions and include a larger range of rld
the most accurate predictions. Algorithm curve fits were and prr coefficient values. SVI coefficient distributions are
constrained to not allow negative coefficient parameter- consistent with the suggestion that where production is
izations. Figure 12 and Table 2 show the data groupings more variable, sinking PM is more labile and decays more
and coefficient algorithms that produced the most accurate rapidly.
flux predictions. Curve fits shown do not imply complex [47] Flux to depth forecast performance abilities of
relationships, but rather show the simplest and most accu- the SST- and SVI-associated parameterizations and the
rate coefficient algorithms attained. relationship presented by Suess [1980] are compared in
[46] Comparison of SST and p ratio data groups shown in Figure 13. The most accurate forecast performance is
Figure 12 indicates little significant variation in p ratios for found using the SVI-related algorithm (Figure 13). All
much of the range of SST. Dissimilar coefficient behavior is equations overestimate flux to depth and p ratios in the
generally confined to minimal SST values. This discontin- upper ocean (1500 m), although to degrees varying
uous distribution may imply that SST influences reminer- depending on latitude range evaluated. Overestimation of

19 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Table 2. Curve Fit Algorithms Describing Coefficients Used to similar to that of flux within the water column. While
Estimate Annual Particulate Organic Carbon Flux to Depth absolute flux rates differ, the geographic patterns of flux
Normalized to Overlying Production (p ratio(ze)) as a Function and export predicted by the SVI algorithm are generally
of Satellite-derived Parameters, the Seasonal Variation Index of similar to other satellite-derived global estimates [Falkowski
Production (SVI; Annual Standard Deviation Divided by Average) et al., 1998; Laws et al., 2000; Müller-Karger et al., 2005],
and Sea-surface Temperature (SST; °C) with proportionally larger fluxes in regions of enhanced
seasonality (for example, higher latitudes and regions of
Parameter (x) Coefficient algorithmb Nc R2
seasonal upwelling).
SVI prd = (31 x2 + 49 x + 7.8) 103 10 0.92
rld = 1400 exp(0.54 x) 10 0.68 [50] Sediment trap-derived estimates of global export are
prr = (2.6 x2  4.2 x + 4.8) 103 8 0.96 significantly less than those developed using other methods.
SST prd = (0.0060 x4  0.42 x3 + 7 1.0 Multiple causes may influence this discrepancy. Accuracy
10 x2  100 x + 340) 103 of sediment trap-derived estimates of export is limited by
rld = 2.0 x2 + 60 x + 680 8 0.39
prr = (0.010 x2  0.34 x + 6.0) 103 7 0.62 the lack of measurements characterizing the upper mesope-
lagic where greatest rates of subsurface recycling occur. As
Coefficients describe flux using equation (6)a: the labile fraction of export
(prd; Figure 12, A and B), remineralization length scale (rld; Figure 12, C noted in section 4.3, underestimation of upper ocean fluxes
and D), and more refractory and rapidly sinking fraction of export (prr; by sediment traps is suspected for a variety of reasons.
Figure 12, E and F).
a z
While radiogenic calibration narrows the discrepancy be-
Equation (6): p ratio(ze) = prd exp( rl e ) + prr. tween estimates, the difference remains significant. Vari-
b d
The functional forms of coefficient algorithms selected are either
polynomial or exponential fits of lowest order constrained to not yield ability associated with the radiogenic technique may be
negative values and produce the most significant curve fits. responsible [Yu et al., 2001]. Finally, a portion of this
c
The number of data groups used are selected to produce the most discrepancy may be due to export of dissolved organic
significant curve fits. carbon [Carlson et al., 1994] not measurable using sedi-
ment traps.
flux and p ratios is largest using the equation of Suess [51] Sediment trap estimates of flux to the seafloor,
[1980], especially at shallow depths. The SVI- and although not significantly dissimilar from one another,
SST-algorithms perform similarly in the deep ocean are significantly greater than flux estimated using apparent
( 1500 m). A portion of the improved accuracy derived oxygen utilization (AOU) in surface sediments [Jahnke,
using equation (6) may be due to our incorporation of a 1996]. This discrepancy may be due to errors associated
larger range flux and production data available to Suess with sediment trap and AOU-associated flux methods.
[1980]. In the upper ocean the SST algorithm over- The accuracy of traps within the deep ocean may be
estimates high-latitude flux and p ratios at all latitudes influenced by some of the hydrodynamic limitations
more than the SVI algorithm. This depth dependant associated with traps in shallow waters. Resuspension of
discrepancy suggests the exponential form of equation benthic detritus into deep-water traps may be greater than
(6) and the depth scaling of the equation of Suess [1980] anticipated. Multiple factors may limit the accuracy of the
do not always account for the most rapid degradation of AOU-associated flux estimations. The flux and subsequent
detritus in the upper water column. accumulation of settled detritus on the seafloor may be
geographically heterogeneous (for example, favoring to-
6.5. Predictions of Annual Flux to Depth pographic low points) and thus not well characterized by
[48] Coupling between surface and subsurface biogeo- spatial distribution of coring techniques typically used to
chemical process has been proposed by comparing satellite- collect benthic surface sediments. Benthic biological ac-
derived estimates of production and sediment trap-derived tivity and bioturbation may further enhance the geographic
flux to depth [Lampitt and Antia, 1997; Fischer et al., 2000; heterogeneity of benthic remineralization [Tedesco and
Antia et al., 2001; Müller-Karger et al., 2005]. We build on Wanless, 1991; Meadows and Meadows, 1994]. Finally,
this insight by using parameters in addition to the rate of the remineralization of flux arriving on the seafloor may
NPP to further constrain flux to depth. Algorithms deve- occur on timescales not assessed by the methods reported in
loped using the remotely sensed parameters, the SVI of the work of Jahnke [1996].
production and SST, and equation (6) allow for predictions
of flux and p ratios and the assessment of forecast accuracy 7. Conclusions
relative to the equation of Suess [1980].
[49] The accuracy of SVI-associated annual flux and p [52] Theoretical connections have been proposed between
ratio predictions allows for prediction of global particulate variability of upper ocean dynamics and pelagic biogeo-
organic carbon flux using satellite-derived NPP (Figure 14 graphy [Longhurst, 1995], and between biogeography and
and Tables 3 and 4). For much of the global ocean, the biogeochemical cycling within the ocean [Longhurst and
geographic pattern of flux is similar at different depths. Harrison, 1989; Lampitt and Antia, 1997]. In particular,
Enhanced export and flux characterize regions where rates Lampitt and Antia [1997] suggest biogeography, as des-
of production and production seasonality are enhanced. cribed by the plankton climate categories of Longhurst
However, in the central northern Atlantic and Pacific [1995], influences marine biogeochemical cycling. Results
Oceans and in the Southern Ocean where export is presented in this study further establish these findings. Our
enhanced, flux to the seafloor is diminished. Flux to the analysis suggests variability of production, as characterized
seafloor is greatest on continental shelves and within centers by the seasonal variation index, reflects ecosystem-scale
of coastal upwelling. At depths greater than the continental biogeochemical processes. This connection may be because,
shelves the geographic distribution of flux to the seafloor is

20 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Figure 13. Residuals describing the accuracy of algorithms used to forecast annual average particulate
organic carbon flux to depth relative to sediment trap observations grouped by latitude. Residuals indicate
(a) predicted minus observed flux rates (mg Corg m2 d1) and (b) the ratio of predicted to observed p
ratios (flux normalized to production). Flux is estimated as a function of satellite-derived data using the
seasonal variation index (annual standard deviation divided by average) of production- (SVI; 6) and
SST-associated (~) using equation (6), and the equation of Suess [1980] (&).

Table 3. Annual Average Particulate Organic Carbon Export, Flux to Depth, and Flux to the Seafloor in the Global Ocean
Depth Flux (Pg Corg yr1) Methods Reference
Export 1.8; 4.6a SVI-algorithm using sediment trap flux and remotely sensed production this study
1 km 0.88; 2.2a SVI-algorithm using sediment trap flux and remotely sensed production this study
2.5 km 0.31 SVI-algorithm using sediment trap flux and remotely sensed production this study
Seafloor 0.76; 1.6a SVI-algorithm using sediment trap flux and remotely sensed production this study
Export 16 f ratio algorithmb using remotely sensed production Falkowski et al. [1998]
Export 10 Global ocean circulation/biogeochemical inverse model using hydrographic, Schlitzer [2000]
oxygen, nutrient, and carbon data
Export 11 – 13 Pelagic food web models using remotely sensed production Laws et al. [2000]
Seafloor 0.886 ‘Mineral ballast’ sediment trap derived model Klaas and Archer [2002]
Seafloor 0.93 Sediment trap derived flux algorithmc and remotely sensed production Müller-Karger et al. [2005]
Seafloor 0.40 Apparent oxygen utilization of surface sediments Jahnke [1996]
a
Radiogenic calibration.
b
Eppley and Peterson [1979].
c
Pace et al. [1987].

21 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Table 4. Annual Average Particulate Organic Carbon Export, Flux to Depth, and Flux to the Seafloor (Pg Corg yr1) Divided Into Ocean
Basins
Regiona Area (%) Export Flux to 1 km Flux to 2.5 km Flux to seafloor
Pacific Ocean 43 0.60, 1.5b 0.30, 0.76b 0.12 0.27, 0.55b
Atlantic Ocean 21 0.52, 1.3b 0.24, 0.60b 0.078 0.27, 0.60b
Indian Ocean 15 0.25, 0.62b 0.12, 0.29b 0.044 0.085, 0.16b
Arctic waters 1 0.026, 0.065b 0.0082, 0.020b 0.00081 0.045, 0.11b
Southern Ocean 21 0.42, 1.0b 0.20, 0.51b 0.061 0.079, 0.15b
Estimates are derived as a function of a satellite-derived net primary production (NPP) and the Index of Seasonal Variation (annual standard deviation
divided by average) of NPP using equation (6).
a
Ocean basins are divided such that: the Southern Ocean is delineated south of 40°S; Arctic waters are delineated north of 70°N; the Indian Ocean is
delineated east of the Cape of Good Hope (20°E) and west of Sumatra and Australia from Cape Londonderry (127°E) to Melbourne (147°E); the Atlantic
Ocean includes the Mediterranean Sea, which has minor a contribution to global flux.
b
Radiogenic calibration at depths <1.5 km.

by measuring environmental change, SVI of production indicating the relative dominance of solar, climatic, and
reflects pelagic ecosystem structure [Longhurst, 1995]. oceanographic controls on the annual variability of NPP. In
[53] The following conclusions are based on our study of general, seasonal patterns of flux reflect those of production.
the biological pump: The NPP climatology displays sea- Prominent patterns of flux include polar, equatorial upwelling,
sonal patterns coherent over large geographic provinces, coastal upwelling, monsoonal, and within subtropical-

Figure 14. Global ocean forecasts of annual average particulate organic carbon (a) export from surface
waters, (b) flux to the base of the mesopelagic zone (1 km), (c) flux to the center of the bathypelagic zone
(2.5 km), and (d) flux to the seafloor (g Corg m2 yr1). Flux is estimated as a function of a satellite-
derived net primary production (NPP) and the seasonal variation index (annual standard deviation
divided by average) of NPP using equation (6). Flux to the seafloor employs the TerrainBase 5-min
global bathymetry available from the National Geophysical Data Center.

22 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

subpolar regions. Notably, within subtropical-subpolar open [57] Acknowledgments. We thank the following researchers for their
inspiration and advice: James Bishop (Laurence Berkeley National Labo-
ocean regions the timing of maximum flux is delayed by ratory), Kevin Arrigo, Gert van Dijken (for advice on satellite image
approximately 5 days per degree latitude increase. Coher- processing), Pamela Matson, Alexandria Boehm, Alessandro Tagliabue,
ence of flux patterns between multiple widely dispersed Rochelle Labiosa, and Tasha Reddy (Stanford University). This research
was supported by a number of sources, including the NSF ROAVERRS
locations demonstrates the ability of sediment traps to program, Lawrence Livermore National Laboratory DOE Center for
characterize sinking particulate matter. Seasonal produc- Research on Ocean Carbon Sequestration, Stanford University McGee
tion-to-flux analyses indicate POC vertical transfer efficien- and A. W. Mellon Foundations, and International JGOFS Program (Ocean
cy is significantly seasonally variable. In particular, the ratio Biogeochemical Modeling Course, Bangalore, India). Finally, we thank the
J. Geophys. Res. reviewers for their recommendations.
of flux to production during bloom production is typically
half that of the remaining year. Comparison of production
and flux variability shows a latitudinal dependant relation- References
ship. At lower-latitudes seasonality of flux is typically Antia, A. N., B. von Bodungen, and R. Peinert (1999), Particle flux across the
mid-European continental margin, Deep Sea Res. Part 1, 46, 1999 – 2024.
greater than that of production, while at higher latitudes, Antia, A. N., et al. (2001), Basin-wide particulate carbon flux in the Atlan-
seasonality of production is typically greater than that of tic Ocean: Regional export patterns and potential for atmospheric CO2
flux. This reversal of variability may describe a biogeo- sequestration, Glob. Biogeochem. Cycles, 15, 845 – 862.
Armstrong, R. A., C. Lee, J. I. Hedges, S. Honjo, and S. G. Wakeham
graphic distinction in the controls of production-to-flux (2002), A new, mechanistic model for organic carbon fluxes in the ocean,
relationships. based on the quantitative association of POC with ballast minerals, Deep
[54] By analyzing a globally distributed data set of NPP Sea Res. Part II, 49, 219 – 236.
and flux to depth, we show the accuracy of algorithms Arrigo, K. R., D. H. Robinson, D. L. Worthen, R. B. Dunbar, G. R.
DiTullio, M. VanWoert, and M. P. Lizotte (1999), Phytoplankton com-
describing flux is improved by incorporating SVI- and SST- munity structure and the drawdown of nutrients and CO2 in the Southern
related controls and describing labile and refractory sinking Ocean, Science, 283, 365 – 367.
fractions of production. The use of the Behrenfeld and Aydin, M., Z. Top, and D. B. Olson (2004), Exchange processes and water-
mass modifications along the subarctic front in the North Pacific: Oxygen
Falkowski [1997a] global NPP model with equation (6), consumption rates and net carbon flux, J. Mar. Res., 62, 153 – 167.
parameterized as a function of the SVI of production and Banse, K. (1990), New views on the degradation and disposition of organic
SST, yields annual satellite-based estimates of deep-sea particles as collected by sediment traps in the open sea, Deep Sea Res.
Part I, 37, 1177 – 1195.
particulate fluxes with greater skill than the commonly Banse, K., and J. R. Postel (2003), On using pigment-normalized, light-
applied equation of Suess [1980]. Global-scale coherence saturated carbon uptake with satellite-derived pigment for estimating
between seasonal patterns of production and flux to depth column photosynthesis, Global Biogeochem. Cycles, 17(3), 1079,
demonstrate the influence of annual climatic variability on doi:10.1029/2002GB002021.
Bathmann, U. V., R. Peinert, T. T. Noji, and B. von Bodungen (1990),
the marine carbon cycle. Results suggest atmospheric CO2 Pelagic origin and fate of sedimenting particles in the Norwegian Sea,
variability is influenced by changes in ecosystem structure Prog. Oceanogr., 24, 117 – 125.
as well as the rate of production. Behrenfeld, M. J., and P. G. Falkowski (1997a), Photosynthetic rates
derived from satellite-based chlorophyll concentration, Limnol. Oceanogr.,
[55] Differentiation between SVI- and SST-associated 42, 1 – 20.
controls is limited because of their covariance. Correla- Behrenfeld, M. J., and P. G. Falkowski (1997b), A consumer’s guide to
tions found do not prove causality. Furthermore, the phytoplankton primary productivity models, Limnol. Oceanogr., 42,
1479 – 1491.
potential for systematic error in either the NPP or Behrenfeld, M. J., W. E. Esaias, and K. Turpie (2002a), Assessment of
sediment trap flux data may influence the accuracy of primary production at the global scale, in Phytoplankton Productivity:
predictions. Greater confidence in flux to depth forecasts Carbon Assimilation in Marine and Freshwater Ecosystems, edited
may require consideration of additional NPP model esti- by P. J. Williams, D. N. Thomas, and C. S. Reynods, pp. 156 – 186,
Blackwell Malden, Mass.
mates [Campbell et al., 2002], flux to depth techniques Behrenfeld, M. J., E. Marañón, D. A. Siegel, and S. B. Hooker (2002b), A
[Buesseler et al., 2000], or other marine carbon cycle photoacclimation and nutrient based model of light-saturated photosynth-
methodologies [e.g., Aydin et al., 2004; Bishop et al., esis for quantifying oceanic primary production, Mar. Ecol. Prog. Ser.,
228, 103 – 117.
2004]. This study is an initial step toward characterizing Behrenfeld, M. J., E. Boss, D. A. Siegel, and D. M. Shea (2005), Carbon-
the efficiency of the biological pump. based ocean productivity and phytoplankton physiology from space, Glo-
[56] Evaluation of additional flux mechanisms may help bal Biogeochem. Cycles, 19, GB1006, doi:10.1029/2004GB002299.
Berelson, W. M. (2002), Particle settling rates increase with depth in the
refine remotely sensed flux relationships and further improve ocean, Deep Sea Res. Part II, 49, 237 – 251.
simulation of the marine biogenic carbon cycle. For example, Berger, W. H., and G. Wefer (1990), Export production: Seasonality and
microbial heterotrophic activity responds to the rate of NPP intermittency, and paleoceanographic implications, Palaeogeogr. Palaeo-
[Lucas et al., 1986] and thus may influence the SVI-derived climatol. Palaeoecol., 89, 245 – 254.
Berger, W. H., K. Fischer, C. Lai, and G. Wu (1989), Ocean productivity
relationship. Differences in the mineral ballast of phytoplank- and paleoproductivity an overview, in Productivity of the Oceans: Pre-
ton may contribute to seasonal variability of particulate sent and Past, edited by W. H. Berger, V. S. Smetacek, and G. Wefer,
organic carbon flux and remineralizaiton [Armstrong et al., pp. 1 – 34, Dahlem Workshop Reports, John Wiley, Hoboken, N. J.
Bishop, J. K. B. (1989), Regional extremes in particulate matter composi-
2002; Klaas and Archer, 2002]. Additional factors that may tion and flux: Effects on the chemistry of the ocean interior, in Produc-
influence the efficiency of the biological pump include tivity of the Ocean: Present and Past, edited by W. H. Berger,
plankton community structure variability [Boyd and Newton, V. Smetacek, and G. Wefer, pp. 117 – 137, John Wiley, Hoboken, N. J.
Bishop, J. K. B., J. C. Stepian, and P. H. Wiebe (1986), Particle matter
1995; Arrigo et al., 1999], phytoplankton taxa-specific distributions, chemistry and flux in the Panama Basin: Response to
particulate matter biodegradability [Dunbar et al., 2003], environmental forcing, Prog. Oceanogr., 17, 1 – 59.
the vertical migration of zooplankton [Bishop et al., 1986; Bishop, J. K. B., S. E. M. Calvert, and Y.-S. Soon (1999), Spatial and
Bishop, 1989], physical forcings of surface waters [Conte et temporal variability of POC in the northeast subarctic Pacific, Deep
Sea Res. Part II, 46, 2699 – 2733.
al., 2001; Fischer et al., 1996b], multiseasonal recycling and Bishop, J. K. B., T. J. Wood, R. E. Davis, and J. T. Sherman (2004),
retention of subsurface particulate matter [Bishop et al., Robotic observations of enhanced carbon biomass and export at 55°S
1999], and mass sedimentation events [Kemp et al., 2000]. during SOFeX, Science, 304, 417 – 420.

23 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Bodungen, B. V., et al. (1995), Pelagic processes and vertical flux of Eppley, R., and B. Peterson (1979), Particulate organic matter flux and
particles: An overview of a long-term comparative study in the Norwe- planktonic new production in the deep ocean, Nature, 282, 677 – 680.
gian Sea and Greenland Sea, Geol. Rundsch., 84, 11 – 27. Fabres, J., A. Calafat, A. Sanchez-Vidal, M. Canals, and S. Heussner
Bory, A., C. Jeandel, N. Leblond, A. Vangriesheim, L. Beaufort, (2002), Composition and spatio-temporal variability of particle fluxes
C. Rabouille, E. Nicolas, K. Tachikawa, and P. Buat-Ménard (2001), in the Western Alboran Gyre, Mediterranean Sea, J. Mar. Syst., 33-34,
Downward particle fluxes within different productivity regimes off the 431 – 456.
Mauritanian upwelling zone (EUMELI program), Deep Sea Res. Part I, Falkowski, P. G., R. T. Barber, and V. Smetacek (1998), Biogeochemical
48, 2251 – 2282. controls and feedbacks on ocean primary production, Science, 281, 200 –
Boyd, P. W., and P. P. Newton (1995), Evidence of the potential influence of 206.
planktonic community structure on the interannual variability of particu- Field, C. B., M. J. Behrenfeld, J. T. Randerson, and P. Falkowski (1998),
late organic carbon flux, Deep Sea Res. Part I, 42, 619 – 639. Primary production of the biosphere: Integrating terrestrial and oceanic
Broecker, W. S., and T. Peng (1982), Tracers in The Sea, 690 pp. components, Science, 281, 237 – 240.
Lamont-Doherty Geological Observatory, Columbia Univ., N.Y. Palisades. Fischer, G., and G. Wefer (1996), Seasonal and interannual particle fluxes
Buesseler, K. O., J. A. Andrews, M. C. Hartman, R. Belastock, and F. Chai in the eastern equatorial Atlantic from 1989 to 1991: ITCZ migrations
(1995), Regional estimated of the export flux of particulate organic car- and upwelling, in Particle Flux in the Ocean, SCOPE, vol. 57, edited by
bon derived from thorium-234 during the JGOFS EqPac program, Deep V. Ittekkot, P. Schafer, S. Honjo, and P. J. Depetris, chap. 10, pp. 199 –
Sea Res. Part II, 42, 777 – 804. 214, John Wiley Hoboken, N. J.
Buesseler, K. O., D. K. Steinberg, A. F. Michaels, R. J. Johnson, J. E. Fischer, G., S. Neuer, G. Wefer, and G. Krause (1996a), Short-term sedi-
Andrews, J. R. Valdes, and J. F. Price (2000), A comparison of the mentation pulses recorded with a fluorescence sensor and sediment traps
quantity and composition of material caught in a neutrally buoyant versus at 900-m depth in the Canary Basin, Limnol. Oceanogr., 41, 1354 – 1359.
surface-tethered sediment trap, Deep Sea Res. Part I, 47, 169 – 187. Fischer, G., B. Donner, B. Davenport, V. Ratmeyer, and G. Wefer (1996b),
Campbell, J., et al. (2002), Comparison of algorithms for estimating Distinct year-to-year particle variations off Cape Blanc during 1988 –
ocean primary production from surface chlorophyll, temperature, and 1991: Relation to d 180-deduced sea-surface temperature and trade winds,
irradiance, Global Biogeochem. Cycles, 16(3), 1035, doi:10.1029/ J. Mar. Res., 54, 73 – 98.
2001GB001444. Fischer, G., V. Ratmeyer, and G. Wefer (2000), Organic carbon fluxes in the
Carlson, C. A., H. W. Ducklow, and A. F. Michaels (1994), Annual flux of Atlantic and the Southern Ocean: Relationship to primary production
dissolved organic carbon from the euphotic zone in the northwestern compiled from satellite radiometer data, Deep Sea Res. Part II, 47,
Sargasso Sea, Nature, 371, 405 – 408. 1961 – 1997.
Chester, R. (1990), Marine Geochemistry, 698 pp., Unwin Hyman, London. Fischer, G., R. Gersonde, and G. Wefer (2002), Organic carbon, biogenic
Collier, R., J. Dymond, S. Honjo, S. Manganini, R. Francois, and R. Dunbar silica and diatom fluxes in the marginal winter sea-ice zone and the Polar
(2000), The vertical flux of biogenic and lithogenic material in the Ross Front Region: Interannual variations and differences in composition,
Sea: Moored sediment trap observations 1996 – 1998, Deep Sea Res. Part Deep Sea Res. Part II, 49, 1721 – 1745.
II, 47, 3491 – 3520. Francois, R., S. Honjo, R. Krishfield, and S. Manganini (2002), Factors
Conte, M. H., N. Ralph, and E. H. Ross (2001), Seasonal and interannual controlling the flux of organic carbon to the bathypelagic zone of the ocean,
variability in deep ocean particle fluxes at the Oceanic Flux Program Glob. Biogeochem. Cycles, 16(4), 1087, doi:10.1029/2001GB001722.
(OFP)/Bermuda Atlantic Times Series (BATS) site in the western Sar- Gardner, W. D. (2000), Sediment trap sampling in surface waters, in The
gasso Sea near Bermuda, Deep Sea Res. Part II, 48, 1471 – 1505. Changing Ocean Carbon Cycle—A Midterm Synthesis of the Joint
Conte, M. H., T. D. Dickey, J. C. Weber, R. J. Johnson, and A. H. Knap Global Ocean Flux Study, edited by R. B. Hanson, H. W. Ducklow,
(2003), Transient physical forcing of pulsed export of bioreactive materi- and J. C. Field, pp. 240 – 281, Cambridge Univ. Press. New York.
al to the deep Sargasso Sea, Deep Sea Res. Part I, 50, 1157 – 1187. Guieu, C., M. Roy-Barman, N. Leblond, C. Jeandel, M. Souhaut, B. Le
Deuser, W. G. (1986), Seasonal and interannual variations in deep-water Cann, A. Dufour, and C. Bournot (2005), Vertical particle flux in the
particle fluxes in the Sargasso Sea and their relation to surface hydro- northeast Atlantic Ocean (POMME experiment), J. Geophys. Res., 110,
graphy, Deep Sea Res., 33, 225 – 246. C07S18, doi:10.1029/2004JC002672.
Deuser, W. G., E. H. Ross, and R. F. Anderson (1981), Seasonality in the Gupta, L. P., and H. Kawahata (2000), Amino acid and hexosamine com-
supply of sediment to the deep Sargasso Sea and implications for the position and flux of sinking particulate matter in the equatorial Pacific at
rapid transfer of matter to the deep ocean, Deep Sea Res. Part A, 28, 175°E longitude, Deep Sea Res. Part I, 47, 1937 – 1960.
495 – 505. Gust, G., A. F. Michaels, R. Johnson, and W. G. Deuser (1994), Mooring
Deuser, W. G., F. E. Müller-Karger, R. H. Evans, O. B. Brown, W. E. line motions and sediment trap hydromechanics: In situ intercomparison
Esaias, and G. C. Feldman (1990), Surface-ocean color and deep-ocean of three common deployment designs, Deep Sea Res. Part I, 4l, 831 –
carbon flux: How close a connection?, Deep Sea Res. Part A, 37, 1331 – 857.
1343. Haake, B., V. Ittekkot, T. Rixen, V. Ramaswamy, R. R. Nair, and
Duce, R. A., and N. W. Tindale (1991), Atmospheric transport of iron and W. B. Curry (1993), Seasonality and interannual variability of particle
its deposition in the ocean, Limnol. Oceanogr., 36, 1715 – 1726. fluxes to the deep Arabian Sea, Deep Sea Res. Part I, 40, 1323 – 1344.
Dunbar, R. B., A. R. Leventer, and D. A. Mucciarone (1998), Biogenic Hargrave, B. T., I. D. Walsh, and D. W. Murray (2002), Seasonal and spatial
sediment fluxes in the Ross Sea, Antarctica: Atmospheric and sea ice patterns in mass and organic matter sedimentation in the North Water,
forcing, J. Geophys. Res., 103, 30,741 – 30,759. Deep Sea Res. Part II, 49, 5227 – 5244.
Dunbar, R. B., K. R. Arrigo, M. J. Lutz, G. R. DiTullio, A. R. Leventer, Hebbeln, D. (2000), Flux of ice-rafted detritus from sea ice in the Fram
M. P. Lizotte, M. L. Van Woert, and D. H. Robinson (2003), Non-Redfield Strait, Deep Sea Res. Part II, 47, 1773 – 1790.
production and export of marine organic matter: A recurrent part of the Hebbeln, D., M. Marchant, and G. Wefer (2000), Seasonal variations of the
annual cycle in the Ross Sea, Antarctica, in Biogeochemistry of the Ross particle flux in the Peru-Chile current at 30°S under ‘‘normal’’ and El
Sea, edited by G. DiTullio and R. Dunbar, Antarctic Research Series, Nino conditions, Deep Sea Res. Part II, 47, 2101 – 2128.
Monograph 78, pp. 179 – 196, AGU, Washington, D. C. Hedges, J. I., F. S. Hu, A. H. Devol, H. E. Hartnett, E. Tsamakis, and
Dymond, J., and R. Collier (1988), Biogenic particle fluxes in the equatorial R. G. Keil (1999), Sedimentary organic matter preservation: A test for
Pacific: Evidence for both high and low productivity during the 1982 – selective degradation under oxic conditions, Am. J. Sci., 299, 529 – 555.
1983 El Niño, Glob. Biogeochem. Cycles, 2, 129 – 137. Honda, M. C. (2001), Studies of carbon cycles in the NW Pacific Ocean
Dymond, J., and M. Lyle (1985), Flux comparisons between sediment and by sediment trap and C14 data, Ph.D. thesis, D. Hokkaido University,
sediment traps in the eastern Pacific: Implications for atmospheric CO2 Sapporo, Japan.
variations during the Pleistocene, Limnol. Oceanogr., 30, 699 – 712. Honda, M. C., K. Imai, Y. Nojiri, F. Hoshi, T. Sugawara, and M. Kusakabe
Dymond, J., and M. Lyle (1994), Particle fluxes in the ocean and implica- (2002), The biological pump in the northwestern North Pacific based on
tions for sources and preservation of ocean sediments, in Studies in fluxes and major component of particulate matter obtained by sediment
Geophysics: Material Fluxes on the Surface of the Earth, edited by trap experiments (1997 – 2000), Deep Sea Res. Part II, 49, 5595 – 5626.
NRC, pp. 125 – 142, National Academy Press, Washington, D. C. Hong, G. H., S.-M. Choe, M.-S. Suk, J.-Y. Na, I. C. Sin, C. S. Chung, and
Dymond, J., R. Collier, J. McManus, S. Honjo, and S. Manganini (1997), S. H. Kim (1996), Annual biogenic particle fluxes to the interior of the
Can the aluminum and titanium contents of ocean sediments be used to East Japan Sea, a large marginal sea of the Northwest Pacific, in Pro-
determine the paleoproductivity of the oceans?, Paleoceanography, 12, ceedings of the International Marine Science Symposium, edited by
586 – 593. S. Tsunogai, pp. 300 – 321, Japan Marine Science Foundation, Mutsu,
El-Sayed, S. Z., and G. A. Fryxell (1993), Phytoplakton, in Antarctic Aomori, Japan.
Microbiology, edited by E. I. Friedman, pp. 65 – 122, Wiley-Leiss, New Honjo, S. (1978), Sedimentation of materials in the Sargasso Sea at a 5,367 m
York. deep station, J. Mar. Res., 36, 609 – 625.

24 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Honjo, S. (1982), Seasonality and interaction of biogenic and lithogenic Laws, E. A., P. G. Falkowski, W. O. Smith, H. Ducklow, and J. J. McCarthy
particulate flux at the Panama Basin, Science, 218, 883 – 884. (2000), Temperature effects on export production in the open ocean,
Honjo, S. (1996), Fluxes of particles to the interior of the open oceans, Glob. Biogeochem. Cycles, 14, 1231 – 1246.
in Particle Flux in the Ocean, SCOPE, vol. 57, edited by V. Ittekkot, Le Quéré, C., et al. (2005), Ecosystem dynamics based on plankton func-
P. Schafer, S. Honjo, and P. J. Depetris, chap. 7, pp. 91 – 154, John tional types for global ocean biogeochemistry models, Glob. Change
Wiley Hoboken, N. J. Biol., 11, 2016 – 2040.
Honjo, S., and S. J. Manganini (1993), Annual biogenic particle fluxes to Lee, C., and C. Cronin (1982), The vertical flux of particulate organic
the interior of the North Atlantic Ocean: Studied at 34°N 21°W and 4S°N nitrogen in the sea: Decomposition of amino acids in the Peru upwelling
21°W, Deep Sea Res. Part II, 40, 587 – 607. area and the equatorial Atlantic, J. Mar. Res., 40, 227 – 251.
Honjo, S., S. J. Manganini, A. Karowe, and B. L. Woodward (1987), Lee, C., and C. Cronin (1984), Particulate amino acids in the sea: Effects of
Particle fluxes, North-Eastern Nordic Seas: 1983 – 1986, in Woods primary productivity and biological decomposition, J. Mar. Res., 42,
Hole Oceanogr. Inst. Tech. Rept. WHOI-87-17, 84 pp., Woods Hole, 1075 – 1097.
Massachusetts. Lee, C., S. G. Wakeham, and J. I. Hedges (1988), The measurement of
Honjo, S., J. Dymond, R. Collier, and S. J. Manganini (1995), Export oceanic particle fluxes—Are ‘swimmers’ a problem? (review and com-
production of particles to the interior of the equatorial Pacific Ocean ment), Oceanography, 1, 34 – 36.
during the 1992 EqPac experiment, Deep Sea Res. Part II, 42, 831 – 870. Lee, C., S. G. Wakeham, and N. Zhu (1992), Effectiveness of various
Honjo, S., J. Dymond, W. Prell, and V. Ittekkot (1999), Monsoon-controlled treatments in retarding microbial activity in sediment trap material and
export? fluxes to the interior of the Arabian Sea, Deep Sea Res. Part II, their effects on the collection of swimmers, Limnol. Oceanogr., 37, 117 –
46, 1859 – 1902. 130.
Honjo, S., R. Francois, S. Manganini, J. Dymond, and R. Collier (2000), Lohrenz, S. E., G. A. Knauer, V. L. Asper, M. Tuel, A. F. Michaels, and
Particle fluxes to the interior of the Southern Ocean in the Western Pacific A. H. Knap (1992), Seasonal and interannual variations in photosynthetic
sector along 170°W, Deep Sea Res. Part II, 47, 3521 – 3548. variability in primary production and particle flux in the northwestern
Ittekkot, V., R. Nair, S. Honjo, V. Ramaswamy, M. Bartsch, S. J. Manganini, Sargasso Sea: US JGOFS Bermuda Atlantic Time-Series, Deep Sea Res.
and B. N. Desai (1991), Enhanced particle fluxes in Bay of Bengal induced Part A, 39, 1373 – 1391.
by injection of water, Nature, 351, 385 – 387. Longhurst, A. (1995), Seasonal cycles of pelagic production and consump-
Jahnke, R. A. (1996), The global ocean flux of particulate organic carbon: tion, Prog. Oceanogr., 36, 77 – 167.
Areal distribution and magnitude, Glob. Biogeochem. Cycles, 10, 71 – 88. Longhurst, A., and W. G. Harrison (1989), The biological pump: Profiles of
Jianfang, C., Z. Lianfu, M. G. Wiesner, C. Ronghua, Z. Yulong, and K. H. plankton production and consumption in the upper ocean, Prog. Ocea-
Wong (1998), Estimations of primary productivity and export production nogr., 22, 47 – 123.
in the South China Sea based on sediment trap experiments, Chin. Sci. Lucas, M. J., S. J. Painting, and D. G. Muir (1986), Estimates of carbon
Bull., 43, 585 – 586. flow through bacterioplankton in the S. Benguela upwelling region based
Jickells, T. D., P. P. Newton, P. King, R. S. Lampitt, and C. Boutle (1996), on 3H-thymidine incorporation and predator-free incubations, IFREMER,
Comparison of sediment trap records of particle fluxes from 19° to 48°N Actes Colloq., 3, 375 – 383.
in the northeast Atlantic and their relation to surface water productivity, Lutz, M. J., R. B. Dunbar, and K. Caldeira (2002), Regional variability in
Deep Sea Res. Part I, 43, 971 – 986. the vertical flux of particulate organic carbon in the ocean interior, Glob.
Kawahata, H. (2002), Suspended and settling particles in the Pacific, Deep Biogeochem. Cycles, 16, 1037 – 1055.
Sea Res. Part II, 49, 5647 – 5664. Martin, J. H., G. A. Knauer, D. M. Karl, and W. W. Broenkow (1987),
Kawahata, H., and L. P. Gupta (2004), Settling particles flux in response to VERTEX: Carbon cycling in the northeast Pacific, Deep Sea Res. Part A,
El Niño/Southern Oscillation (ENSO) in the equatorial Pacific, in Global 34, 267 – 285.
Environmental Change in the Ocean and on Land, edited by M. Shiyomi Meadows, A., and P. S. Meadows (1994), Bioturbation in deep sea Pacific
et al., pp. 95 – 108, Terra Sci., Tokyo. sediments, J. Geol. Soc., 150, 361 – 375.
Kawahata, H., and H. Ohta (2000), Sinking and suspended particles in the Mohiuddin, M. M., A. Nishimurab, Y. Tanakab, and A. Shimamotoc
Southwest Pacific, Mar. Freshwater Res., 51, 113 – 126. (2004), Seasonality of biogenic particle and planktonic foraminifera
Kawahata, H., M. Yamamuro, and H. Ohta (1998), Seasonal and vertical fluxes: Response to hydrographic variability in the Kuroshio Extension,
variations of sinking particle fluxes in the west Caroline Basin, Oceanol. northwestern Pacific Ocean, Deep Sea Res. Part I, 51, 1659 – 1683.
Acta, 21, 521 – 532. Müller-Karger, F., R. Varela, R. Thunell, Y. Astor, H. Zhang, R. Luerssen,
Kawahata, H., A. Suzuki, and H. Ohta (2000), Export fluxes in the Western and C. Hua (2004), Processes of coastal upwelling and carbon flux in the
Pacific Warm Pool, Deep Sea Res. Part I, 47, 2061 – 2091. Cariaco Basin, Deep Sea Res. Part II, 51, 927 – 943.
Kawahata, H., A. Nishimura, and M. K. Gagan (2002), Seasonal change in Müller-Karger, F. E., R. Varela, R. Thunell, R. Luerssen, C. Hu, and J. J.
foraminiferal production in the western equatorial Pacific warm pool: Walsh (2005), The importance of continental margins in the global car-
Evidence from sediment trap experiments, Deep Sea Res. Part II, 49, bon cycle, Geophys. Res. Lett., 32, L01602, doi:10.1029/2004GL021346.
2783 – 2800. Neuer, S., V. Ratmeyer, R. Davenport, G. Fischer, and G. Wefer (1997),
Kemp, S., and H. Knaack (1996), Vertical particle flux in the western Deep water particle flux in the Canary Island region: Seasonal trends in
Pacific below the north equatorial current and the equatorial counter relation to long-term satellite derived pigment data and lateral sources,
current, in Particle Flux in the Ocean, SCOPE, vol. 57, edited by Deep Sea Res. Part I, 44, 1451 – 1466.
V. Ittekkot P. Schafer S. Honjo, and P. J. Depetris, chap. 18, pp. 313 – Neuer, S., R. Davenport, T. Freudenthal, G. Wefer, O. Llinás, M. J. Rueda,
324, John Wiley, Hoboken, N. J. D. K. Steinberg, and D. M. Karl (2002), Differences in the biological
Kemp, A. E. S., J. Pike, R. B. Pearce, and C. B. Lange (2000), The ’’Fall carbon pump at three subtropical ocean sites, Geophys. Res. Lett., 29(18),
dump’’: A new perspective on the role of a ‘‘shade flora’’ in the annual 1885, doi:10.1029/2002GL015393.
cycle of diatom production and export flux, Deep Sea Res. Part II, 47, Newton, P. P., R. S. Lampitt, T. D. Jickells, P. King, and C. Boutle (2004),
2129 – 2154. Temporal and spatial variability of biogenic particle fluxes during the
Klaas, C., and D. E. Archer (2002), Association of sinking organic matter JGOFS northeast Atlantic process studies at 47°N, 20°W, Deep Sea
with various types of mineral ballast in the deep sea: Implications for the Res. Part I, 41, 1617 – 1642.
rain ratio, Glob. Biogeochem. Cycles, 16, 63 – 77. Nodder, S. D., and L. C. Northcote (2001), Episodic particulate fluxes at
Knauer, G. A., D. M. Karl, J. H. Martin, and C. N. Hunter (1984), In situ southern temperate mid latitudes (42 – 45°S) in the Subtropical Front
effects of selected preservatives on total carbon, nitrogen and metals region, east of New Zealand, Deep Sea Res. Part I, 48, 833 – 864.
collected in sediment traps, J. Mar. Res., 42, 445 – 462. Pace, M., G. Knauer, D. Karl, and J. Martin (1987), Primary production,
Kuss, J., and K. Kremling (1999), Particle trace element fluxes in the deep new production and vertical flux in the eastern Pacific Ocean, Nature,
northeast Atlantic Ocean, Deep Sea Res. Part I, 46, 149 – 169. 325, 803 – 804.
Lampitt, R. S. (1992), The contribution of deep-sea macroplankton to Parsons, T. R., M. Takahashi, and B. Hargrave (1977), Biological oceano-
organic remineralization: Results from sediment trap and zooplankton graphic processes, 300 pp., Elsevier, New York.
studies over the Madeira Abyssal Plain, Deep Sea Res. Part A, 39, 221 – Pilskaln, C. H., J. B. Paduan, F. P. Chavez, R. Y. Anderson, and W. M.
233. Berelson (1996), Carbon export and regeneration in the coastal upwelling
Lampitt, R. S., and A. N. Antia (1997), Particle flux in deep seas: Regional system of Monterey Bay, central California, J. Mar. Res., 54, 1149 –
characteristics and temporal variability, Deep Sea Res. Part I, 44, 1377 – 1178.
1403. Pilskaln, C. H., S. J. Manganini, T. W. Trull, L. Armand, W. Howard, V. L.
Lampitt, R. S., B. J. Bett, K. Kiriakoulakis, E. E. Popova, O. Ragueneau, Asper, and R. Massom (2004), Geochemical particle fluxes in the South-
A. Vangriesheim, and G. A. Wolff (2001), Material supply to the abyssal ern Indian Ocean seasonal ice zone: Prydz Bay region, East Antarctica,
seafloor in the Northeast Atlantic, Prog. Oceanogr., 50, 27 – 63. Deep Sea Res. Part I, 51, 307 – 332.

25 of 26
C10011 LUTZ ET AL.: SEASONALITY OF THE BIOLOGICAL PUMP C10011

Platt, T., and S. Sathyendranath (1993), Fundamental issues in measurement Sediment trap results from the Cariaco Basin, Limnol. Oceanogr., 45,
of primary production, ICES Mar. Sci. Symp., 197, 3 – 8. 300 – 308.
Pudsey, C. J., and P. King (1997), Particle fluxes, benthic processes and the Trull, T. W., S. G. Bray, S. J. Manganini, S. Honjo, and R. Francois (2001),
palaeoenviron-mental record in the northern Weddell Sea, Deep Sea Res. Moored sediment trap measurements of carbon export in the Subantarctic
Part I, 44, 1841 – 1876. and Polar Frontal Zones of the Southern Ocean, south of Australia,
Ramseier, R. O., E. Bauerfeind, and R. Peinert (1999), Sea-ice impact on J. Geophys. Res., 106, 31,489 – 31,509.
long term particle flux in the Greenland Sea’s Is Odden-Nordbukta Tsunogai, S., and S. Noriki (1991), Particulate fluxes of carbonate and
region, 1985 – 1996, J. Geophys. Res., 104, 5329 – 5343. organic carbon in the ocean. Is the marine biological activity working
Rivkin, R. B., and L. Legendre (2001), Biogenic carbon cycling in the as a sink of the atmospheric carbon?, Tellus, 43B, 256 – 266.
upper ocean: Effects of microbial respiration, Science, 291, 2398 – 2400. Tyrell, T. (1999), The relative influences of nitrogen and phosphorus on
Rühlemann, C., P. J. Müller, and R. R. Schneider (1999), Organic carbon oceanic primary production, Nature, 400, 525 – 531.
and carbonate as paleoproductivity proxies: Examples from high and low Volk, T., and M. I. Hoffert (1985), Ocean carbon pumps: Analysis of relative
productivity areas of the tropical Atlantic, in Use of Proxies in Pale- strengths and efficiencies in ocean-driven atmospheric CO2 changes, in
Oceanogrraphy: Examples from South Atlantic, edited by G. Fisher The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to
and G. Wefer pp. 315 – 344, Springer, New York. Present, edited by E. T. Sundquist and W. S. Broecker, Geophysical
Sakshaug, E., A. Bricaud, Y. Dandonneau, P. G. Falkowski, D. A. Kiefer, Monograph Series, vol. 32, pp. 99 – 110, AGU, Washington, D. C.
L. Legendre, A. Morel, J. Parslow, and M. Takahashi (1997), Parameters Wefer, G., and G. Fischer (1991), Annual primary production and export
of photosynthesis: Definitions, theory and interpretation of results, flux in the Southern Ocean from sediment trap data, Mar. Chem., 35,
J. Plankton Res., 19, 1637 – 1670. 597 – 613.
Sanchez-Vidal, A., A. Calafat, M. Canals, and J. Fabres (2004), Particle Wefer, G., and G. Fischer (1993), Seasonal patterns of vertical particle flux
fluxes in the Almeria-Oran Front: Control by coastal upwelling and sea in the equatorial and coastal upwelling areas of the eastern Atlantic, Deep
surface circulation, J. Mar. Syst., 52, 89 – 106. Sea Res. Part I, 40, 1613 – 1645.
Sarmiento, J. L., and C. Le Quéré (1996), Oceanic carbon dioxide uptake in Wefer, G., G. Fischer, D. Füetterer, and R. Gersonde (1988), Seasonal
a model of century-scale global warming, Science, 274, 1346 – 1350. particle flux in the Bransfield Strait, Antarctica, Deep Sea Res. Part A,
Sarmiento, J. L., T. M. C. Hughes, R. J. Stouffer, and S. Manabe (1998), 35, 891 – 898.
Simulated response of the ocean carbon cycle to anthropogenic climate Wefer, G. G. Fischer, D. K. Füetterer, R. Gersonde, S. Honjo, and
warming, Nature, 393, 245 – 249. D. Ostermann (1990), Particle sedimentation and productivity in Antarctic
Schlitzer, R. (2000), Applying the adjoint method for biogeochemical mod- waters of the Atlantic sector, in Geological History of the Polar Oceans:
eling: Export of particulate organic matter in the world ocean, in Inverse Arctic Versus Antarctic, edited by U. Bleil and J. Thiede, pp. 363 – 379,
Methods in Global Biogeochemical Cycles, edited by P. Kasibhatla, Springer, New York.
M. Heimann, P. Rayner, N. Mahowald, R. G. Prinn, and D. E. Hartley, Wiesner, M. G., L. Zheng, H. K. Wong, Y. Wang, and W. Chen (1996),
pp. 107 – 124, Geophysical Monograph Series, vol. 114, AGU, Fluxes of particulate matter in the South China Sea, in Particle Flux
Washington, D. C. in the Ocean, SCOPE, vol. 57, edited by V. Ittekkot, P. Schafer,
Schmittner, A. (2005), Decline of the marine ecosystem caused by a reduc- S. Honjo, and P. J. Depetris, chap. 16, pp. 293 – 312, John Wiley
tion in the Atlantic overturning circulation, Nature, 434, 628 – 633. Hoboken, N. J.
Schneider, B., K. Nagel, and U. Struck (2000), Carbon fluxes across the Wong, C. S., F. A. Whitney, I. Tsoy, and A. Bychkov (1994), Opal pump
halocline in the eastern Gotland Sea, J. Mar. Syst., 25, 261 – 268. and subarctic carbon removal, in 1994 Sapporo IGBP Symposium, Sap-
Scholten, J. C., et al. (2001), Trapping efficiencies of sediment traps from poro, Japan, 339-347.
the deep eastern North Atlantic: The 230th calibration, Deep Sea Res. Wong, C. S., F. A. Whitney, D. W. Crawford, K. Iseki, R. J. Matear, W. K.
Part II, 48, 2383 – 2408. Johnson, J. S. Page, and D. Timothy (1999), Seasonal and interannual
Siegel, D. A., and W. G. Deuser (1997), Trajectories of sinking particles in variability in particle fluxes of carbon, nitrogen and silicon from time
the Sargasso Sea: Modeling of statistical funnels above deep-ocean sedi- series of sediment traps at Ocean Station P, 1982 – 1993: Relationship to
ment traps, Deep Sea Res. Part I, 44, 1519 – 1541. changes in subarctic primary productivity, Deep Sea Res. Part II, 46,
Sigman, D. M., and E. A. Boyle (2000), Glacial/interglacial variations in 2735 – 2760.
atmospheric carbon dioxide, Nature, 407, 859 – 869. Yoder, J. A., C. R. McClain, G. C. Feldman, and W. E. Esaias (1993),
Six, K. D., and E. Maier-Reimer (1996), Effects of plankton dynamics on Annual cycles of phytoplankton chlorophyll concentrations in the global
seasonal carbon fluxes in an ocean general circulation model, Glob. ocean—A satellite view, Glob. Biogeochem. Cycles, 7, 181 – 193.
Biogeochem. Cycles, 10, 559 – 583. Yu, E. F., R. Francois, M. P. Bacon, S. Honjo, A. P. Fleer, S. J. Manganini,
Suess, E. (1980), Particulate organic carbon flux in the oceans—Surface M. M. R. VanderLoeff, and V. Ittekot (2001), Trapping efficiency of
productivity and oxygen utilization, Nature, 288, 260 – 263. bottom-tethered sediment traps estimated from intercepted fluxes of
230
Takahashi, K., N. Fujitani, M. Yanada, and Y. Maita (2000), Long-term Th and 231Pa, Deep Sea Res. Part I, 48, 865 – 889.
biogenic particle fluxes in the Bering Sea and the central subarctic Pacific
Ocean, 1990 – 1995, Deep Sea Res. Part I, 47, 1723 – 1759.
Tedesco, L. P., and H. R. Wanless (1991), Generation of sedimentary
fabrics and facies by repetitive excavation and storm infilling of burrow 
networks, Holocene of South Florida and Caicos Platform, B.W.I., M. J. Behrenfeld, Department of Botany and Plant Pathology, Cordley
Palaios, 6, 326 – 343. Hall 2082, Oregon State University, Corvallis, OR 97331-2902, USA.
Thunell, R. C. (1998a), Particle fluxes in a coastal upwelling zone: Sedi- K. Caldeira, Department of Global Ecology, Carnegie Institution of
ment trap results from Santa Barbara Basin, California, Deep Sea Res. Washington, 260 Panama St., Stanford, CA 94305, USA.
Part II, 45, 1863 – 1884. R. B. Dunbar and M. J. Lutz, Department of Geological and
Thunell, R. C. (1998b), Seasonal and annual variability in particle fluxes in Environmental Sciences, Stanford University, Stanford, CA 94305-2115,
the Gulf of California: A response to climate forcing, Deep Sea Res. Part USA. (michael.jlutz@gmail.com)
I, 45, 2059 – 2083.
Thunell, R., R. Varela, M. Llano, J. Collister, and R. Bohrer (2000),
Organic carbon fluxes and regeneration rates in an anoxic water column:

26 of 26

You might also like