You are on page 1of 33

Chapter 1

Why Dissolved
Organics Matter
John I. Hedges
School of Oceanography, University of Washington, Seattle, Washington

I.
II.
III.
IV.
V.

Introduction
DOM Research Pre-1970
DOM Research in the 1970s
DOM Research in the 1980s
"New" DON and DOC

VI. Why Dissolved Organics


Matter
VII. What did we Learn?
References

I. INTRODUCTION
As this book attests, research on dissolved organic matter (DOM) in seawater
has burgeoned in the past decade. This increase in activity is evident not only from
the growing number of articles published each year in the scientific literature,
but also from the topical breadth and broad integration of present research. The
oceanographic community's perception of DOM has evolved from an emphasis
on a dilute and largely separate pool of remarkably old and static substances to
the current view of a dynamic assemblage of organic molecules that interact with
each other, trace metals, and living organisms over a broad continuum of space and
time scales. The sparingly reactive components of this molecular continuum that
persist and change on time scales sampled by conventional oceanographic surveys
represent a small molecular outcrop of a churning mass of molecules through
which much of the total primary production of the ocean cycles.
To better understand what is to come in this chapter and book, it is useful to keep
in mind that investigations of DOM in seawater have followed two fundamentally
Biogeochemistry of Marine Dissolved Organic Matter
Copyright 2002, Elsevier Science (USA). All rights reserved.

John I. Hedges

different strategies. The first is a holistic approach focusing primarily on the total
concentration, bulk properties, and collective behavior of the entire mixture of
molecules that make up the operationally defined DOM pool. Examples would be
measurements of the total dissolved organic carbon (DOC) or dissolved organic
nitrogen (DON) concentrations, determinations of bulk spectral or isotopic compositions, and estimates of cumulative oxygen and nutrient changes attending microbial attack of the entire organic mixture. This strategy has the major advantage
of yielding characteristics that are representative of the entire DOM pool, but the
information obtained is typically limited and highly biased toward the less reactive
components of the mixture that accumulate over time. In contrast, the reductionist
approach has been to target selected fractions of the total mixture for detailed analyses of specific features that might then be meaningfully extrapolated back to the
bulk pool. The most common form of reductionism is the chromatographic analysis
of specific biochemical components of seawater DOM. This particular strategy can
yield a wealth of information on structural features, stereochemistries, and reaction
pathways and dynamics. However, molecular-level analyses are highly selective
for individual biochemical classes (or subclasses), which in turn often comprise
a tiny, and not necessarily representative, fraction of bulk DOM. Thus, major uncertainties arise in extrapolating from detailed molecular-level information to the
whole DOM pool, and especially to its emergent properties.
This introductory chapter emphasizes the oceanographic community's perceptions of the entire DOM pool from a bulk chemical perspective, bringing in biochemical and microbiological information primarily as it pertains to the larger view.
While this focus on collective properties necessitates that substantial advances at
the biochemical level will not be highlighted, it does allow better historic continuity and further development of broad issues pertaining to oceanography in
general. This chapter recaps selected experimental and conceptual developments
extending from the last century up through the Seattle DOC/DON Workshop Report (Hedges and Lee, 1993) that have led to the modem dynamic view of oceanic
DOM presented in the following chapters.

11. DOM RESEARCH PRE-1970


By 1970, study of seawater DOM had already been under way for almost a
century (see review by Kalle, 1966). Glass fiber or silver filters available in the
mid-20th century had minimal pore sizes of ^0.45-1.0 /xm and became the basis
of the operational definition that "dissolved" materials pass such filters whereas
"particulate" matter does not (Fig. 1). This definition persists to today, although we
now know that seawater contains a continuum of discrete units stretching from the
size of whales to that of a water molecule, with no discemable break in abundance
in the micrometer range (Sharp, 1973a). The traditional definition can be useful.

W/zy Dissolved Organics Matter


mm

urn

I
Meters

nm

10-^I III I 10-^


mill III

10-5
10-^
IN III nil, lllllilJI

10-^

Partlcuiate

10-8
iiiiiii III

lo-^
mill I

III

10-^

Dissolved
Colloids

^fel

I
[

Viruses

Macyomolecutes "[

Sand
'^rCliyJ
Screen
Sieves

^^
-^iijfi'lWr,.,L^
Papers f
Ultrafilters

Figure 1

Molecular

-^

^-

Sieves

The continuum of sizes and separation methods for organic matter in seawater.

however, because particles smaller than 1/xm are not prone to sink (Duursma,
1961) and all living organisms other than viruses and small bacteria fall into the
particulate fraction. Colloidal particles, constituting the upper size range (0.0011.0 jjim) of the DOM continuum, correspond in minimal size to approximately a
six-sugar oligosaccharide (Fig. 1).
Following several largely unsuccessful early attempts (e.g., Piitter, 1909; Raben,
1910) to quantify the dissolved organic contents of seawater, Krogh and Keys
(1934) published comparatively reproducible methods for the determination of
both DON and DOC in seawater. The DON method was based on a micro-Kjeldahl
(sulfuric acid hydrolysis) procedure, whereas DOC was quantified (after chloride
removal) by wet oxidation in aqueous chromic acid. Using these methods, Krogh
(1934a) measured the first full water column profiles of DOC and DON in the open
ocean off Bermuda. He found uniform concentrations of organic material from the
surface down and concluded that seawater DOM is chronologically old, chemically
and biochemically inert, and insignificant as a food source for organisms in the
deep sea (Krogh, 1934b). The following year, however, Waksman and Carey (1935)
demonstrated in a series of culturing experiments that bacteria decompose DOM
from surface seawater in a matter of days, with attending increases in inorganic
nitrogen and decreases in dissolved oxygen.
Kalle (1937) used UV absorption to detect yellow organic substances in the
waters of the North Sea and open North Atlantic. Although spectroscopically similar to DOM in rivers, seawater "gelbstoff" was recognized to have a predominant

John I. Hedges

marine origin (Kalle, 1949). Kalle (1949) also reported an organic component of
seawater DOM that gives a bluefluorescencewhen irradiated with long-wavelength
ultraviolet light and appears to have a predominantly terrestrial source. Early attempts to isolate seawater DOM by sorption onto charcoal (Wilson and Armstrong, 1952; Johnston, 1955) or extraction with nonpolar solvents (Slowley et al,
1959; Chanu, 1959) were successful, although subsequent chemical characterizations were primarily limited to demonstrating UV absorbance and the presence of
trace amounts of fatty acids (Jeffrey and Hood, 1958). Various laboratory experiments (e.g., Fogg and Boalch, 1958) demonstrated that marine algae (especially
phaeophyta) are potential direct sources of seawater DOM. At this time, amino
acids and carbohydrates were known to spontaneously condense (although at elevated temperatures) to produce melanoidin polymers (Maillard, 1913) that exhibit
many of the spectral qualities of marine DOM (Kalle, 1966). By the early 1960s,
DOC was measured at concentrations on the order of 1 mg/L (83.3 /xM) and found
to be more concentrated in surface ocean water than at depth (Kay, 1954; Plunkett
and Rakestraw, 1955; Duursma, 1961). In addition, a variety of component biochemicals, including simple sugars, low-molecular-weight acids, and vitamin B12,
had been detected in seawater (Vallentyne, 1957; Hood, 1970; Duursma, 1965).
A wave of pioneering field studies during the 1960s served mainly to strengthen
the perception of deep-ocean DOM as a largely static pool. Improved wet chemical
oxidation methods for seawater DOM (e.g., the persulfate adaptation of Menzel
and Vacarro, 1964) became the basis for extensive surveys of DOC concentrations
in various oceans (e.g., Menzel, 1964; Menzel and Ryther, 1968). Menzel's 1964
study of DOC distributions in the western Indian Ocean was by far the most
extensive to that time with respect to the number of stations (39) and depths
(1-2000 m) sampled. In addition to synoptic temperature and salinity data for
each sample, this study included ^"^C-based measurements of primary production
under simulated euphotic zone conditions. No apparent correlation between DOC
concentration and primary production rates was observed in surface ocean waters.
Although DOC concentrations below 200 m ranged geographically between 0.2
and 2 mg/L ('^ 15-170 /xM), these gradients covaried linearly with salinity and
thus appeared to result primarily from mixing of different water masses with
characteristically different DOC signatures. Menzel (1964) concluded, "carbon in
solution and in particulate form in the ocean is extremely stable and subject to
limited change by biological activity."
Menzel and Ryther (1968) soon published a more detailed study of dissolved
and particulate organic carbon (POC) distributions in discrete water samples collected over the entire water column at 14 stations in the southern Atlantic Ocean.
In contrast to the Indian Ocean survey, dissolved oxygen was directly measured
for each sample, along with temperature and salinity. Relatively constant DOC
concentrations (35 5 /xM) were observed at depths greater than 500 m throughout the South Atlantic (Fig. 2). Suspended POC accounted for roughly 1% of

Why Dissolved Organics Matter

DOC, |LiM
0

20

40
I

60
^1

80

100

Figure 2 Vertical profile of DOC in the southwest Atlantic Ocean (after Menzel and Ryther, 1968).
Arrow lengths indicate the range of measured values, with the profile line passing through the mean
value for that depth. Values in parentheses represent the total number of multiple analyses at one depth.

DOC below 500 m depth and also was essentially invariant. Dissolved O2 varied
linearly versus salinity (Fig. 3) at the core of Antarctic intermediate water in all profiles. This observation of minimal DOC variation (Fig. 2) over a substantial oxygen
gradient of > 100 /xM (Fig. 3) supported previous evidence for Httle or no DOC
respiration below 500 m (Menzel and Ryther, 1970). By comparison, the theoretical OC/O2 ratio for respiration of "average marine plankton" is 106/138 = 0.77
(Redfield et al, 1963), whereas the best current estimate is near 0.70 (Anderson,
1995). Russian researchers at this time were measuring DOC concentrations by
high-temperature combustion of freeze-dried samples. Although this method
indicated concentrations that were approximately three times higher than those
obtained with persulfate (see review by Starikova, 1970), minimal changes in
deep-ocean DOC profiles were nonetheless noted (Skopintsev, 1966).
Independent evidence that deep-sea DOM is refractory came from a variety
of other sources. Barber (1968) demonstrated that DOM concentrated fivefold
from deep-ocean water was not measurably utilized by marine bacteria and argued

John I. Hedges
F"^^T^T

JUU

-^
H

-1

250 m''-,.

"X,

200

' 1

150
0.200

^ ^ ^ ^ ^ 1

^^^^^^^^^

0.300
0.400
Salinity - 34.000

0.500

Figure 3 Measured dissolved oxygen versus the salinity in excess of 34.000 at the core of the
Antarctic intermediate water in the southwest Atlantic Ocean (based on data from Table 1 of Menzel and
Ryther, 1968). The equation of the best-fit line (r^ = 0.992) to the 13 data points is O2 = (-439 12) x
(S-34) + (374 4).

against previous speculation (Jaanasch, 1967) that seawater DOM might simply
be too dilute to serve as a suitable substrate. P. M. Williams (1968b) showed
that the stable carbon isotopic composition of DOC is consistent with a predominantly marine origin and essentially constant throughout the water column of
the San Diego trough. The definitive experiment of the decade, however, was the
demonstration by Williams et al (1969) that the radiocarbon content of dissolved
organic matter from the deep Pacific Ocean corresponds to a radiocarbon "age" of
roughly 3400 years BR If this radiocarbon "age" is assumed to represent a mean
residence time (Williams et al, 1969), it corresponds to a steady-state flux of
roughly 0.2 x 10^^ g C/year through the ocean DOC pool (650-700 x 10^^ g C).
Critically, this small flux would necessitate that only 0.4% of global primary production enters the marine DOC pool per year. Although a flux of this order could
be supported by riverine DOC discharge alone (Williams, 1971; Mantoura and
Woodward, 1983), the stable carbon isotopic composition of seawater DOC points
toward a marine origin (WilHams, 1968a). WiUiams (1971) concluded that the
predominant uncharacterized fraction of seawater DOM is humic-like and thus
intrinsically unreactive.
At the same time, parallel evidence was accumulating that an appreciable fraction of DOM in surface ocean waters can be physically and biologically reactive
under at least some conditions. Natural slicks were observed to form and disperse

Why Dissolved Organics Matter


rapidly at the ocean surface (Ewing, 1950; Jarvis, 1967) and to contain a variety of
surface-active organic materials (Garrett, 1967, 1970) that could be concentrated
by a dipped screen (Garrett, 1965), rotating drum (Harvey, 1966) or, "bubble microtome" (Maclntyre, 1966). In a series of experiments, Sieburth and Jensen (1968,
1969) demonstrated exudation of DOM by phaeophyta (kelp) and associated formation of sea surface slicks (Sieburth and Conover, 1965). Duursma (1961,1963,
1965) observed greater than twofold seasonal variation of DOC in surface waters
of the North Sea. This indication of cycling on a monthly time scale suggested the
possible use of DOC as an indicator of primary production. In contrast to results
for deep water, Barber (1968) found that DOM concentrated from surface seawater
exhibited a relatively short half-life (1-2 months) with respect to bacterial remineralization. The list of chromatographically measured biochemicals also increased
substantially and the more abundant fatty acids, amino acids, and sugars had been
quantified in surface waters and over a few deep-sea profiles (Holm-Hansen et al,
1966; Duursma, 1965; WiUiams, 1971). However, only about 10% of the DOC in
surface and subsurface waters could be accounted for as individually measurable
biochemical types, even when results from separate studies were added together
(Williams, 1971). Although potentially labile biochemicals were evident, their low
concentration was taken as additional evidence for a largely refractory pool of bulk
DOM. The decade closed with a short conmiunication by Riley and Taylor (1969)
describing how fatty acids and humic substances can be recovered from acidified
seawater (pH 2) by sorption onto a cross-linked polystyrene resin called Amberlite
XAD-1.

III. DOM RESEARCH IN THE 1970s


The perception of a labile DOM component in the surface ocean accompanied
by largely inert DOM (marine humus) that predominates below ~500 m continued to develop in the 1970s. The decade opened with the report by Williams and
Gordon (1970) that the stable isotope composition of DOC at multiple stations
in the northeast Pacific Ocean is remarkably uniform (5^^C = -22.6 db 0.6%^ )
and independent of depth and time, as well as dissolved O2 and DOC concentrations. The observation that these values were similar to those of local POC and
marine plankton pointed toward a predominant marine origin of oceanic DOM.
This inference was supported by a very different 8^^C value of 28.5%o measured
for DOM from the Amazon River (Williams, 1968a). Although rivers discharge
DOC at a rate sufficient to support the entire marine pool (WiUiams, 1971), the
much more ^^C-enriched composition of marine DOC indicates that land-derived
DOC must be rapidly removed or profoundly changed in its stable carbon isotopic composition. Minimal changes in the S^^C of marine DOC in depth profiles

John I. Hedges
lUUI

80

A 24.0-24.5
I
1
25.0 - 25.5 1
D 25.5-26.0
1
26.0 - 26.5 1
A 26.5 - 27.0 . , ^ ^^ ^ 1
X >27.5
-^^^O-^^-^l

3<
2

d
o
Q

WSjlii

60

=1

Ox
*
0

<24.0
24.5-25.0

^ v ^ A^ A-"AAA

A ^ . ^ 1 jAiWP 1

40
XX *

JT^X

20

A* -^

AD0C/A02 = -1 /3 X (106/138)

LJ
100

1
200

300

AOU, ^iM
Figure 4 DOC concentration versus AOU in waters of the western North Pacific Ocean (after Ogura,
1970). Different symbols correspond to various isopycnal intervals within the seawaters that were
sampled. The line of best fit was by the original author.

characterized by correspondingly large variations in dissolved O2 were interpreted


by Williams and Gordon (1970) to suggest that, "dissolved organic matter is stable
to biochemical fractionation and oxidation."
In a milestone paper that in many ways foreshadowed controversies to come,
Ogura (1970) reported a weak overall relationship (Fig. 4) between DOC and
apparent oxygen utilization (AOU) in shallow (at < 26.5) waters of the western
North Pacific (east of Japan). His DOC measurements, by the standard persulfate
method of Menzel and Vaccaro (1964), gave overall concentrations and depth
trends similar to those reported by Menzel and Ryther (1968) in the South Atlantic.
No relationship between DOC and AOU was apparent in deeper waters (at > 26.5),
where AOU increased substantially against a relatively constant background DOC
near 40 /xM (Fig. 4). By directly combining DOC and AOU measurements for the
first time, Ogura demonstrated that approximately half of the total DOC in shallow
waters of the western North Pacific is labile, whereas the other half resembles deepocean DOC in both its concentration and minimal reactivity. An inverse DOC/AOU
trend did occur in shallower waters, but at only one-third of the 106/138 molar
ratio expected from Redfield stoichiometry (Redfield et al, 1963) for complete
remineralization of average marine plankton. Ogura inferred that the organic matter
being respired in the upper water column was either compositionally very different
from Redfield plankton or that another O2 sink, such as the "particulate organic

Why Dissolved Organics Matter


substances" existed throughout the water column. Soon thereafter Craig (1971)
reported field evidence for measurable O2 consumption on the time scale of water
mass mixing in the deep Pacific Ocean, a result that was supported by later culture
experiments (Williams and Carlucci, 1976). Within half a decade, McCave (1975)
published a theoretical paper indicating that rare, fast-sinking large particles might
compose an as-yet-unsampled source of labile organic matter to the deep ocean
which could explain Ogura's and Craig's observations of in situ O2 consumption.
Improved methods for DOC analysis were also a main research thrust in the
1970s. New methods for high-temperature combustion of dry samples generated
by freeze-drying (Gordon and SutcHffe, 1973) and aspirating (Mackinnon, 1978)
seawater were described. Sharp (1973b) reported the first practical adaptation of
an earlier Hquid injection method (van Hall et al, 1963) for rapid high-temperature
combustion of seawater DOC. These high-temperature combustion methods typically measured greater DOC concentrations than were obtained by prevalent
wet chemical methods involving persulfate (Menzel and Vaccaro, 1964) or UV
(Williams, 1969a,b) oxidation. This observation supported previous Russian findings based on dry combustion methods that the reservoir of DOC in the ocean
might be up to twice as large as previously accepted (Sharp, 1973b).
A second major analytical thrust of the 1970s was development of ultrafiltration to separate and characterize colloidal fractions of seawater DOM. Using
Millipore and Diaflo membranes with nominal pore sizes of 0.025 and 0.003 /xm.
Sharp (1973a) separated roughly 10 and 20% of total DOC, respectively. He noted
that colloidal DOC is approximately an order of magnitude more abundant than
suspended POC, leading to a monotonic increase in the cumulative fraction of
seawater organic matter found in progressively smaller size fractions. Similar
molecular size distributions were reported for profiles from the Gulf of Mexico by
Maurer (1976), who inferred from colorimetric analyses that proteins and carbohydrates accounted for approximately 15 and 20%, respectively, of colloidal DOM
throughout the water column (see also Ogura, 1977).
Another advance was the development of improved hydrophobic sorption methods for large-scale isolation and subsequent characterization of milligram amounts
of salt-free DOM. Kerr and Quinn (1975), building on the pioneering research of
Jeffrey and Hood (1958), isolated seawater DOM by adsorption onto charcoal and
demonstrated its lower UV absorbance versus sedimentary and estuarine DOM.
The tour de force of this era, however, was the work of Stuermer and Harvey
(1974,1977), who carried out the first detailed analysis of seawater DOM isolated
from Sargasso Sea and coastal seawater onto XAD-2, a microporous polystyrene
resin. This work built on the isolation methodology worked out by Riley and
Taylor (1969) and Mantoura and Riley (1975), which on average recovered less
than 10% of the total organic carbon in ocean water (Stuermer and Harvey, 1977).
Because the tradition at that time was to view structurally uncharacterized materials in soils, sediments, and natural waters as "humic substances," the first paper

10

John I. Hedges

in this series (Stuermer and Harvey, 1974) was entitled "Humic Substances from
Seawater." The isolated seawater humic substances (SHS) where found to be
low in molecular weight (predominantly <700 amu), highly aliphatic (atomic
H/C = 1.6), oxygen rich (0/C = 0.55), and nitrogen poor (C/N = 15). The latter
value, however, must be considered with caution because all isolates were eluted
from XAD-2 with anmionium hydroxide (pH 11.6) solution from which nitrogen
may have been added. SHS had 8^^C values near 23%o and, like bulk seawater
DOC, appeared to be marine derived. Later ^^C NMR (Stuermer and Payne, 1976)
and chemical degradation (Stuermer and Harvey, 1978) studies confirmed the
highly aliphatic character of these SHS isolates, although the NMR spectra were
noisy. Overall the DOM fractions isolated from seawater were markedly different
in their compositional signature from riverine, soil, and sedimentary counterparts
(e.g.. Fox, 1983). Nevertheless, the "humic" label stuck and conferred an image
of extreme structural alteration and low substrate quality.
Following the lead of Garrett (1965,1967) and others, research on the concentration and reactions of organic matter at the air-sea interface blossomed in the 1970s
(see reviews by Liss, 1975; Hunter and Liss, 1981). Commonly employed isolation methods involved dipping glass (Harvey and Burzell, 1972) or Teflon (Garrett
and Barger, 1974) plates, or plastic funnels (Morris, 1974) through the sea surface
microlayer. A particularly creative approach for both collecting and characterizing
surface ocean films was sorption onto a germanium prism within which infrared
light could be attenuated via multiple internal reflections (Baier, 1972). The resulting IR spectra of most natural films were characterized by the presence of bands at
3400,1650 and 1100 cm~^ that are typical of carbohydrate-like material, as well as
bands at 3300 plus 1530 and 1440 cm~^ indicating protein and carboxyl groups, respectively (Baier et al, 1974). Little if any absorbance at 2950 and 2850 cm"^ was
observed that would demonstrate the presence of C-H stretching in lipids. Baier
et al (1974) concluded that glucoprotein-like substances concentrate at the seaatmosphere interface, although lipids could be observed when slicks are apparent.
This conclusion was largely supported by molecular-level analyses that showed
major carbohydrate and protein components, and minor lipid concentrations, in
sea surface films (Hunter and Liss, 1981).
The 1970s also saw increased attention to the reactions of bulk seawater DOM
and its biochemical content (see reviews by Williams, 1975; Ogura, 1977). Trace
metals, and copper in particular, were observed in complexes with DOM (e.g.,
Siegel, 1971). Sensitive methods for analysis of amino acids were developed, including means for separating D versus L enantiomers of selected compounds (Lee
and Bada, 1975,1977). Observation of elevated D/L aspartic acid ratios in Sargasso
Sea waters was one of thefirstdirect indications of a possible bacterial cell wall origin (Bada and Lee, 1977). Decomposition of organic matter dissolved in seawater
was observed to advance with selective loss of measurable biochemicals and selective preservation of the high-molecular-weight fraction (Ogura, 1975). By analogy

Why Dissolved Organics Matter

11

with both terrestrial systems (Flaig, 1964) and laboratory simulations (Maillard,
1913; Hodge, 1953), such trends were widely viewed as being indicative of "heteropolycondensation" reactions yielding structurally complex and enzymatically
intractable dissolved humic substances (Degens, 1970).

IV. DOM RESEARCH IN THE 1980s


Efforts to improve methods for DOM isolation and characterization continued
as a major research direction into the 1980s, with a special emphasis on NMR methods. Fu and Pocklington (1983) reported that the sequential passage of acidified
seawater onto XAD-2 resin and activated charcoal columns, followed by sequential elutions with methanol and NH4OH, recovered the equivalent of 90-100% of
the initial DOC. Little and Jacobs (1985) demonstrated that DOM can be adsorbed
from seawater directly onto titanium foil, which could then be pyrolyzed to products that are quite different from those generated by DOM isolated with XAD-2
resin. Wilson et al (1983) pubHshed some of the first high-quality ^H (solution)
and ^^C NMR (solution and solids) spectra of coastal seawater DOM isolated
by sorption onto XAD-2 resin. These spectra showed evidence for carbohydrate,
highly branched alkyl chains, and (to a lesser extent) aromatic carbon as major
structural units of seawater DOM and algal exudates. Complementary analysis of
the same samples by pyrolysis gas chromatography-mass spectrometry (Wilson
et al, 1983; Gillam and Wilson, 1985) indicated phenol- and carbohydrate-based
structural units. Harvey et al (1983) published proton NMR spectra of seawater
humic and fulvic acids, which were both characterized by resonances of aliphatic
structures. They proposed that such macromolecules might be formed by lightcatalyzed, oxidative cross-linking of polyunsaturated fatty acids or glycerides via
a free-radical reaction. This pathway, however, did not alone explain the presence
nitrogen or the fluorescence of seawater DOM (Laane, 1984).
Another major research theme of the 1980s was the study of fluorescent substances dissolved in seawater. Following the lead of Kalle (1949), one line of research involved tracing highlyfluorescentriverineDOM into the coastal ocean. For
example, Willey and Atkinson (1982) were able to sensitively trace (~ 1 part in 200
parts seawater) discharges of the Savannah and Ogeechee Rivers (Georgia, U.S.A.)
through their respective estuaries and into adjacent coastal waters with salinities
greater than 30 ppt. The coastal plain Ogeechee River exhibited approximately
twice the fluorescence of the piedmont Savannah River, allowing contributions of
these two sources to be distinguished off the Georgia coast. Similar tracer applications were described by Hayase et al (1987) and by Laane (1982) for riverine organic matter discharged into Tokyo Bay and the Ems-DoUart estuary (Netherlands/
Germany), respectively. Donard et al (1989) described a high-sensitivity method
based on laser excitation sources for fluorescent DOM in the Mediterranean Sea

12

John I. Hedges

Dark Control
\.9^.^..^./....

.-

Dark Control

Sunlight (unsterilized)

Sunlight (sterilized)
20 h

10

15

20

25

Days of Treatment
Figure 5 Decrease of the natural fluorescence of unsterilized and sterilized coastal seawater during
exposure to sunlight. A dark control showed no change in fluorescence over the 20-day experiment
(after Kramer, 1979).

and detected different spectral signatures between coastal and offshore waters. The
second line of research built on Kramer's (1979) observation that the natural fluorescence of seawater DOM is lowest in the upper water column and is decreased
substantially (>50%) during exposure to sunUght over a period of days (Fig. 5).
Hayase et al. (1988) later demonstrated that vertical profiles of seawater fluorescence in the North Pacific closely parallel those of dissolved nutrients, suggesting
that fluorescent DOM is released as degrading organic substances streaming from
sinking particles into ambient seawater. They showed that the fluorescence of
deep-sea DOM is also readily bleached by UV light, thereby helping to explain
the minimal fluorescence of surface seawater throughout the open ocean.
The previous recognition of photobleaching and the photosensitizing properties
of seawater (Momzikoff et al, 1983; Zepp et al, 1985) led to a number of studies
in the later 1980s on the photochemical reactivity of seawater DOM. For example,
Mopper and Stahovec (1986) demonstrated in laboratory andfieldexperiments that
low-molecular-weight carbonyl compounds such as formaldehyde, acetone, and
glyoxal can be formed by the photolysis of seawater DOM. This process was most
pronounced in the surface ocean and in coastal regions where terrigenous DOM
prevails (Kieber and Mopper, 1987). The photodegradation products of seawater
DOM were rapidly taken up by marine microbes, thereby providing a pathway
for the cycling of relatively refractory seawater DOM into planktonic food webs
(Kieber ^? a/., 1989).

Why Dissolved Organics Matter

13

Molecular highlights of the 1980s include the application of highly sensitive


analytical methods for individual amino acids (Garrasi et al, 1979) and sugars (Mopper et al, 1980) to determine the composition and dynamics of these
DOM components in their free and combined (polymeric) forms. For example,
Liebezeit et al (1980) found that individual free carbohydrates (<5-650 nmol)
and amino acids (<l-30 nmol) are enriched in the upper boundaries of density
gradients in the Sargasso Sea, apparently due to high particle concentrations and
heterotrophic activities of these layers. Dissolved free amino acids (including aspartic and glutamic acid, threonine, serine, glycine, and alanine) turned over more
rapidly than corresponding sugars. Ittekkot et al (1981) demonstrated the release
of large amounts of dissolved combined sugars toward the end of an algal bloom
in the North Sea. As observed elsewhere (e.g., Liebezeit et al, 1980), glucose
and fructose composed more than 90% of the free sugars, whereas glucose, mannose, galactose, and xylose predominated in the combined fraction. Sakugawa and
Handa (1985) and Sakugawa et al (1985) isolated dissolved polysaccharides from
coastal seawater by dialysis and sorption onto charcoal. A variety of dissolved
oligosaccharides and polysaccharides were observed, including a glucan and a
heteropolysaccharide containing glucose, galactose, mannose, xylose, fucose, and
rhamnose (Sakugawa and Handa, 1985). Later, Meyers-Schulte and Hedges (1986)
detected small amounts of chemically recognizable lignin, and hence terrigenous
DOM, in waters of the East Equatorial Pacific.
In 1987, WilUams and Druffel measured the first detailed profiles of the radiocarbon content of DOC and DIC in water samples collected throughout a 5700-m
water column of the central North Pacific (Fig. 6). They confirmed the great radiocarbon age (equivalent to >6000 years) of DOC in the deep ocean that was
first reported almost two decades before by Williams et al (1969). Williams and
Druffel (1987) also demonstrated that the A^^C profiles of DOC and dissolved
inorganic carbon (DIC) are remarkably parallel in the Pacific Ocean (Fig. 6) and
concluded that the same transport and recycling processes control the radiocarbon distributions of both dissolved carbon pools. However, the approximately
300%o lower A^'^C values of DOC versus coexisting DIC suggested that a fraction of DOC is recycled within the ocean on much longer time scales than for
DIC, which largely equilibrates with the atmosphere on a 500- to 1000-year time
scale.

V. "NEW DON AND DOC


A series of events that was to galvanize the oceanographic community began
in the mid-1980s with publication of a paper by Suzuki et al (1985) on the determination of total dissolved nitrogen (TDN) in seawater. The method was based
on high-temperature (680 C) catalytic oxidation (HTCO) of a small (200 /xL)

14

John L Hedges

A^^C, o/oo
-600

200

lOOOt

6000
Figure 6 Profiles of the relative radiocarbon contents (A^^C) of dissolved inorganic carbon (DIC)
and dissolved organic carbon (DOC) in the water column of the central North Pacific Ocean (data from
Table 1 of Druffel et al, 1989).

sample of filtered seawater directly injected into a quartz column filled with
platinum-impregnated (3 wt%) alumina (AI2O3). High purity O2 acted both as
a carrier gas and an oxidizing agent. Injected organic and inorganic nitrogen compounds were all converted to nitrogen dioxide that then combined with a chromogenic reagent, A^-(l-naphthyl)-ethylenediamine dihydrochloric acid, to absorb
light at 545 nm. Total organic nitrogen was determined as the difference between
total dissolved nitrogen (measured as above) and total inorganic nitrogen (nitrate +
nitrite + ammonium). The method description included extensive tests of the effects of combustion column temperature, matrix ions (Cl~, F~, and 804"), and
injection sample volume (5 to >200 /xL) on linearity (0-13 mM), precision (<2%
for TDN), and analytical error (<2% for TDN). Conversion efficiencies of the analytical system for a variety of inorganic (nitrate) and organic (e.g., caffeine, oxine,
and albumin) substances, as well as molecular size fractions of seawater DOM,
were found to be >95%. The only exception was sulfathiazole (<65%). TDN
measurements of reference materials and natural seawater samples by the HTCO
method compared well with those obtained by pyrolytic oxidation of freeze-dried
sea salt, whereas nitrate analyses following aqueous persulfate oxidation gave
much lower (20-90% recovery) results. This pattern of higher TDN values measured by HTCO versus persulfate was also observed when these two methods

Why Dissolved Organics Matter

15

were compared for water column profiles at two different sites in the western
North Pacific (Suzuki et aU 1985).
In retrospect, some aspects of the Suzuki et al paper were problematic. For
example, all seawater samples were processed through Millipore HA filters that
were made of mixed acetate and nitrate esters of cellulose and hence were potential
sources of both DOC and TDN. Blank runs for all analyses were made without
sample injection and thus were not procedurally representative and would not have
picked up contamination derived from filtration or released by injected water from
alumina. The Sephadex gel-filtration method (Sugimura and Suzuki, 1983) used
to size fractionate DOM in seawater samples indicated that less than 90% of the
component nitrogen resides in molecules < 1500 amu. This was a surprising result
because previous (Sharp, 1973a; Maurer, 1976) and following studies (Carlson
et al, 1985; Benner et al, 1992) indicated that 75% or more of total seawater
DOC has a molecular weight less than ^1000 amu. Organic nitrogen would have
to be very strongly concentrated in higher-molecular-weight molecules to account
for this contrast in DON versus DOC distribution (see later discussion). Most
troubling, however, TDN concentrations (30^0 /xM) measured in the surface
100 m at two different sites in the western North Pacific (e.g.. Fig. 7) were much
higher than previously reported (e.g., Duursma, 1961; Williams, 1975; Jackson
and Williams, 1985). The directly measured large increases in DIN with depth
(^40 /zM) produced corresponding decreases in DON concentration (measured as
TDN-DIN) on the order of 35 /xM at both sites. The negative correlations between
DON and AOU at these two study sites were strong (r^ > 0.95, Fig. 7) and averaged
0.13, compared to a theoretical molar AN/AAOU of 0.12 for the respiration
of average marine plankton (Redfield et al, 1963). This result suggests that most
of the cumulative respiration evident in the upper 1000 m of each water column
was supported by oxidation of DOM, as opposed to sinking organic particles.
The second paper in this series focused on a derivative high-temperature catalytic oxidation method for DOC analysis (Sugimura and Suzuki, 1988) and began
a cascade of events that carried over into the 1990s and much of the research reported in the following chapters. The DOC paper was similar in many ways to the
DON pubHcation of 3 years before. The major differences were that water samples
were first passed through a Nuclepore (polycarbonate) membrane filter, acidified
with H3PO4, and purged to remove DIC. The same Pt-catalyzed combustion column and procedure were used as for TDN. The generated CO2 was analyzed with
a nondispersive infrared detector. The procedural blank was estimated by injecting
different volumes of seawater and extrapolating the measured responses back to
zero volume. The value of this blank, however, was not reported. Oxidation efficiencies to CO2 were determined for many of the same organic substances (e.g.,
caffeine, oxine, and albumin) tested in the 1985 paper and again found to be >95%,
with the exception of sulfathiazol (<65%). The trend in measured DOC concentration versus combustion temperature was very different for surface versus deep

John I. Hedges

16

Dissolved N, |iM
0

10

20

30

40

50

Figure 7 Representative plots (for the profile taken at 39 59' N, 144 02' E) of total dissolved nitrogen
(TDN), dissolved inorganic nitrogen (DIN), and dissolved organic nitrogen (DON) concentrations
versus AOU (data from Table VII of Suzuki et al, 1985).

waters, with approximately half of the DOM in surface water requiring temperatures in excess of 600C (versus 500C for deep-ocean DOM) to obtain complete
oxidation. At an optimal temperature of 680C, the HTCO analyzer reportedly
yielded linear responses over the range of 0-2000 //M of DOC (5- to 400 /nL
injection volumes) and a precision of 2% for glucose and natural seawater.
Sugimura and Suzuki (1988) also reported extensive comparisons of their
HTCO method to the conventional persulfate procedure (Menzel and Vaccaro,
1964). For surface seawater from the North Pacific (30 N, 147 W) they measured
DOC concentrations that were approximately 2.5 times higher than the concentration of 100 fxM obtained by persulfate oxidation and values for deeper waters that
averaged almost twice the persulfate-derived concentrations of ~70 fiM. Overall
differences between DOC concentrations measured by HTCO and persulfate were
even greater (~3x overall) for larger size fractions of a surface seawater, with
most of the discrepancies occurring in the higher-molecular-weight (>4000 amu)
fractions. These results were supported by the finding that over 60% of the initial
DOC (232 IJM by HTCO) in an unspecified seawater sample subjected for 2 h

Why Dissolved Organics Matter

17

DU-

"""^'"""^'^^

45^
Z
O

A--'''

G
CI

fz

40-

.."'

1
1

* < #-"

.'

35-

.,-''

3025 H

100

200

300

400

500

DOC, |LiM
Figure 8 DON versus DOC in surface waters from the North Pacific and East China Sea (data from
Table VII of Sugimura and Suzuki, 1988). The five open squares indicate high-DOC waters with a
salinity less than 33 that were collected off the mouth of the Changjiang (Yangtze) River. The equation
of the best-fit line (r^ = 0.625) to the other 32 data points (sohd squares) is DON = (0.554 0.078) x
D O C + (25.9 1.5).

to persulfate oxidation still could be measured by HTCO in the residual water.


The bulk of the remaining DOC reportedly occurred in the >4000-amu size fractions, indicating again that high-molecular-weight DOC is particularly resistant to
persulfate oxidation.
In the same paper, Sugimura and Suzuki reported application of the new HTCO
method in an extensive survey of surface seawaters and three depth profiles from
the western North Pacific. In general, these field results (Fig. 8) strongly reinforced the implications of the previously described laboratory tests. DOC values
in the range of 180-490 luM were measured for surface seawaters south of Japan,
with values above 275 /xM being limited to five low-salinity (<33) surface waters
collected off the mouth of the Changjiang (Yangtze) River in the East China Sea
(Fig. 8). In comparison, DOC concentrations previously measured by the conventional persulfate method in surface waters of the open ocean (e.g.. Fig. 2) rarely
exceeded 100 fiM (WilHams, 1971, 1975). Commensurately high (-35-45 fiM)
DON concentrations were measured for these same surface waters using the Ptcatalyzed HTCO combustion method of Suzuki et al. (1985). The average atomic
C/N ratio for all 37 surface seawaters for which DOC and DON data are listed
in Table VII of Sugimura and Suzuki (1988) is 7.2 1.6. If the five carbon-rich
(C/N = 9.5-12) water samples collected off the Changjiang (Fig. 8) are deleted,
the remaining 32 surface waters have an average C/N ratio of 6.6 zb 0.6. This ratio
has a very small (< 10%) standard deviation and is the same in value as the atomic
C/N of average marine plankton (Redfield et al, 1963).

18

John I. Hedges

However, other patterns in these same data sets were curious. In particular,
the previously mentioned strong correlation between DON and DOC values for
the 32 open-ocean surface waters generates a best-fit line (r^ = 0.63) with an
intercept of 25.9 1.6 /xM of DON at zero DOC (Fig. 8). The slope of this line
(0.0554 0.0078) corresponds to an atomic C/N of 18. In contrast, Jackson and
Williams (1985) found little correlation between seawater DOC and DON contents
measured throughout the ocean by persulfate oxidation and noted that most of
their individual samples gave DOC/DON ratios higher than Redfield values. They
observed, however, that it is the slope of the DOC/DON correlation line for a
sample suite, and not the individual ratios, which indicates the C/N ratio within
DOM sample sets. Their seawater samples converged on minimal persulfate-based
DOC and DON concentrations near 30 and 1.2 /xM, respectively, indicating a
predominant (--80% of DOC) background of "inert" DOM with a C/N near 25
occurring in the deep ocean. The slope of Sugimura and Suzuki's data set for
surface waters (Fig. 8) actually indicates DOM components with a much higher
C/N ratio than Redfield value (18 vs 7). The intercept obtained by extrapolating this
data trend to the DON axis indicates a huge background (^^25 /xM) of additional
nitrogenous material deriving systematically from either the water samples or
the DON measurement procedure. The close cluster of the measured individual
C/N ratios about the Redfield value thus appears highly fortuitous. Elevated C/N
values for the five surface waters exhibiting lower salinities would be expected
because riverine DOM is typically nitrogen poor with a global average C/N near
25 (Meybeck, 1982).
It was the deep ocean profiles of DOC versus DON, DIN, and AOU, however, that appeared to give the greatest evidence that the new Pt-catalyzed
HTCO method provided "oceanographically consistent" results. The three profiles
(0-2000 m) published by Sugimura and Suzuki (1988) were all very similar in form
and extent (e.g.. Fig. 9). All show sharp increases in AOU and DIN with depth to
values near 250 and 40 /xM, respectively. DOC and DON profiles were inverse to
those for the inorganic species, decreasing from surface values near 300 and 40 /xM,
respectively, to concentrations at 2000 m of approximately 50 and 5 /xM. Although
no data were tabulated for any of these profiles, the nominal DOC/DON ratios for
all samples can be seen to closely approximate Redfield values near 7. Molecular
weight analysis by gel filtration of waters from two of the three profiles again
indicated that most of the DON and DOC occurred in molecules >4000 amu,
which were reported to make up ~90% of surface DON and DOC. The relationship that most captured the attention of the oceanographic community, however,
was the reported near-linear inverse relationship between DOC and AOU measurements for all three profiles (Fig. 10). The slope of this trend was very near 1.0,
which led the authors to conclude that "the AOU in water can be explained by the
in situ decomposition of DOC." Based on such evidence, Sugimura and Suzuki
inferred that previous findings of little or no relationship between AOU and DOC

Why Dissolved Organics Matter

19
AOU, DIG, ^iM

(D
0-

100

200

i-|

100^

DOcNi

DON v i
if

k k

jc 1

M 1

^o^^lh:^

E
Q.
0
Q

1--^

1 ^^
\f f

50 _ 19 4 AOU
^ f

-1

300

__L.,

300^
1 TDN
500-]
DIN \

1000-1

2000 J
0

10

20

30

40

50

DON, DIN, TDN, |iM


Figure 9 DON, DOC, DIN, and AOU depth profiles in the western Pacific Ocean at 1901' N, 13400'
E (after Fig. 1 la of Sugimura and Suzuki (1988), with TDN determined graphically as DIN + DON).

1 1

^^^^^mf^^m^^^^^

xJfc "

050r N, 13444'E |-|

3001

X 1955'N, 13000'E 1

I^ ^
II

^ 2000'N, 13659'E | |

*
X

200

O
O

1
1

1
J
1

ioor

[l

^
A^ ^
X

1
1
1

^ i ^ J i ^

100
AOU,

200

300

RM

Figure 10 DOC versus AOU from all three depth profiles in the western Pacific Ocean published by
Sugimura and Suzuki (1988).

20

John L Hedges

(e.g., Figs. 3 and 4) were the result of incomplete persulfate- and UV-based DOC
measurements.
As has been pointed out (Williams and Druffel, 1988; Farrington, 1992), however, there were reasons to question the Sugimura and Suzuki findings and conclusion. One conceptual problem pertains to the likelihood of a nearly linear relationship of DOC versus AOU at a molar slope near 1 (Fig. 10). This is because the
initial concentration of DOC in seawater at the time it sinks from the surface is not
physically controlled within a narrow range (Toggweiler, 1989a), as is the initial
concentration of oxygen gas (by its solubility). In fact, the DOC concentrations in
local surface seawaters reported by Sugimura and Suzuki in the same 1988 paper
varied by over 100 /xM (Fig. 8), even when the five high values measured for low
salinity samples in the East China Sea are excluded. Moreover, waters collected
over the upper 2000 m of the water colunm in the western North Pacific do not all
have the same geographic origin or time of sinking and thus also do not represent
a common degradation history that would closely couple DOC/AOU trends down
the depth profile (Toggweiler, 1989a).
Also at issue is whether DOM is likely alone to support respiration throughout the top two kilometers of the western North Pacific Ocean. Large concentrations of dissolved organic phosphorous and nitrogen that should accompany
the higher concentrations of DOC measured by Pt-catalyzed HTCO instruments
in the surface ocean were not measured by other researchers (Karl et al, 1993;
WilUams and Druffel, 1988). Why the "excess" DOM missed by relatively severe
wet chemical oxidation should nevertheless be readily utilized by microorganisms
in the upper thermocline also was unclear (Toggweiler, 1989b). By this time it
had become widely recognized from sediment trap studies that sinking particles
are a major sink for O2 and a source of dissolved nutrients, in the interior ocean
(Knauer et al, 1979; Martin et al, 1987). Appreciable in situ degradation of sinking particles, however, would decrease the slope of DOC/AOU relationship to
values considerably less than one (Fig. 4). In addition, any surface seawater that
downwelled containing 40 /xM DON and >300 /xM DOC would have more than
enough reducing potential (>380 /xM) in DOM alone to remove all the oxygen
it could possibly dissolve (~350 /xM O2, at 0C and salinity of 35; Broecker
and Peng, 1982). Such waters would be highly prone to anoxia if in addition
they hosted any in situ respiration of sinking organic particles. Sinking bioactive
particles also would be expected to cause a pronounced downward displacement
in TDN profiles, which is not evident in Fig. 9. These inconsistencies, however,
were difficult to test at the time because the exact composition of the platinized
alumina support (Sumitomo Chemical Industry Co. Ltd.), which appeared to be
critical to the efficient performance of the Sugimura and Suzuki HTCO DOC
and DON analyzers, was proprietary information. Another hurdle for comparative measurements was the authors' report that reliable DOC results could be
obtained only when seawater samples were filtered and immediately analyzed

Why Dissolved Organics Matter

21

aboard ship. In addition, neither reference seawater samples nor DOM-free water
for blank testing were widely available to the oceanographic community at
that time.
Not surprisingly, the "revolutionary" (Toggweiller, 1989a)findingsof Sugimura
and Suzuki generated a rush by other research groups to acquire and apply
Pt-catalyzed HTCO units for DOC and DON analyses. Not only did such instruments appear to measure DOC and DON more efficiently than predecessors
employing wet chemical oxidation, they also were faster, easier, more readily automated and required much smaller sample volumes (~ 100 /xL) than were necessary
(5 mL) for the standard persulfate method. By 1991, Pt-catalyzed HTCO analyzers for DOC and DON were available from at least six different companies and
were in use by more than 20 research groups in the oceanographic community
(Hedges et al, 1993). Data both supporting and refuting the "New" DOC/DON
hypothesis flooded the oceanographic literature. Theories for why this previously
unknown component of seawater DOM was measurable only by Pt-HTCO analyzers proliferated, as did papers attempting to characterize these materials and
investigate their biogeochemical importance (see Hedges and Lee, 1993). Unfortunately, the results of these follow-ups were mixed, including two papers (Benner
et al, 1992; Ogawa and Ogura, 1992) that cast doubt on most of Suzuki's size
distribution and composition findings. Such divergent observations in the early
1990s caused confusion, consternation, and cautious support by national agencies
for DOM research.
The growing furor led to an NSF/NOAA/DOE-sponsored workshop on the
"Measurement of Dissolved Organic Carbon and Nitrogen in Natural Waters" that
was held in Seattle in July 1991 (Hedges and Lee, 1993). To assess measurement uniformity within the oceanographic conmiunity, ampoulated samples of
surface, mid-depth, and deep-ocean waters collected at the Hawaii ALOHA station (and from a Hawaiian river) were distributed before the workshop to invited
participants. A total of 13 independent DON measurements, and 34 independent
DOC analyses of the sample suite were made. The results were not encouraging
(Hedges et al., 1993). The DON measurements varied by an average of 30% of
the mean value for the samples and were not related to any known aspect of the
analyzers or their use. The corresponding DOC concentrations varied by an average of 40% of the mean values, with HTCO instruments generally measuring
higher concentrations than were obtained for the same samples by more conventional wet-chemical techniques (Fig. 11). The DOC differences, however, largely
disappeared when the mean value for each analyst's four individual samples was
subtracted from the corresponding individual measurements (Hedges et al, 1993).
Such a pattern following mean subtraction would be expected if the major sources
of difference among DOC concentrations measured by the participating labs were
traceable to large background signals intrinsic to each instrument and its method of
operation.

22

John I. Hedges

350|
300

" Oj Minimum Zone

250|
^ 200|
U
Surface Ocean

lOOl
5o| Deep Ocean
0

Means from Each Set of Analyses


Figure 11 Trends in measured DOC among reference seawater samples analyzed for the Seattle
DOC/DON Workshop (data from Hedges et ai, 1993). Individual data sets are listed in order of
decreasing measured concentration for the surface sample. Analyses by wet chemical oxidation methods
are indicated by shaded backgrounds.

Supporting evidence for large DOC blanks intrinsic to HTCO-based instruments came from additional workshop reports. In particular, a critical evaluation
by Benner and Strom (1993) of the analytical blank associated with DOC measurements using HTCO instruments showed that the platinized-alumina packing
used in the combustion columns of almost all of the tested commercially-available
analyzers (Shimadzu, Aldrich and Sumitomo) generated large initial blanks equivalent to 50 to >200 /xM of DOC in a 200-/xL injection. A sample of the type of
support distributed by Sumitomo and used in the Sugimura and Suzuki HTCO
system gave an initial instrument blank value equivalent to 90 /xM of DOC, which
could only be reduced to 27 di 5 /xM by 100 sequential injections of reoxidized
water. DON measurements (by persulfate or HTCO) approaching the 40 /xM levels
routinely reported by Suzuki et al (1985) and Sugimura and Suzuki (1988) were
not later measured (e.g. Walsh, 1989; Hansell, 1993; Koike and Tupas, 1993; Karl
et al, 1993). Subsequent to these findings, and to a report by Tanoue (1992) of
much lower DOC and DON concentrations obtained in the western North Pacific
with an improved HTCO analyzer, Suzuki (1993) retracted the data presented in
both his 1985 and 1988 papers. The reasons for the unusually high and closely correlated DON and DOC concentrations appearing in the two retracted papers were
not completely clear, although inappropriate attention to instrument blanks was
apparently a major problem (Suzuki, 1993). Essentially none of the concentration,

Why Dissolved Organics Matter

23

elemental composition or size distribution results published in the two Suzuki


papers has been subsequently confirmed.

VI. WHY DISSOLVED ORGANICS MATTER


In view of the huge community response to "New" DOC and the subsequent
spate of research on DOM that is described in the following chapters, it is interesting to evaluate why this high level of interest and productivity has been sustained
through what could have been a discouraging setback brought about by the Suzuki
retractions. In the case of DOM research, much of the reason for this continued
level of research activity is traceable to major advances in parallel fields. One of
these allied developments has been steadily increasing interest in the global carbon
cycle, especially as it relates to greenhouse gases such as CO2 and associated climate change. The birth of this movement can be traced to Svante Arrhenius (1896),
who pointed out that humans are increasing atmospheric CO2 concentrations by
burning fossil fuels. Arrhenius made the remarkable estimate that a doubling of
atmospheric CO2 concentration would lead to a 5-6C increase in the average
temperature of the Earth's surface. Over 50 years later, the reality of increasing
atmospheric CO2 concentrations was demonstrated by direct measurements (e.g..
Keeling, 1973; Keeling et al, 1995) and the magnitude of the Arrhenius temperature projection was supported by numeric global climate models (Houghton et al,
1996).
Both the great size and potential dynamics of the ocean DOM pool have
brought it within the focus of global cycle research (Williams and Druffel,
1988; Toggweiler, 1989b, Hedges, 1992). Because the amounts of carbon in
oceanic DOM (--700 x 10^^ g) and atmospheric CO2 (^750 x 10^^ g) are similar
(Siegenthaler and Sarmiento, 1993), net oxidation of only 1 % of the seawater DOM
pool within 1 year would be sufficient to generate a CO2 flux larger than that produced annually by fossil fuel combustion. Concentration differences of this magnitude would be extremely difficult to identify due to the current limits of analytical
precision and the heterogeneous distributions of DOC in the ocean. It is not surprising, therefore, that the Sugimura and Suzuki 1988 report of roughly twice as much
total DOC in the ocean as was previously measured gained the immediate attention
of the oceanographic conmiunity. In addition to greatly raising the global stakes
for budgeting actively cycling organic carbon, this report placed three to six times
more DOM in the surface ocean where the bulk of this uncharacterized material
appeared to be biologically active on time scales of years to decades (Toggweiler,
1989b). This apparent increase in the organic acid component of seawater was also
of substantial interest as a potential explanation for the discrepancy between total CO2 measurements in seawater by potentiometric versus manometric methods

24

John I. Hedges

(Bradshaw and Brewer, 1988). Quantitative constraints on organic matter cycling


are particularly difficult in the physically and biologically active upper surface
ocean (Quay, 1997), where DOC versus POC exports are difficult to distinguish and
DOM photodegradation can accompany photosynthesis. Given the susceptibility
of DOC to photolysis (Kieber et ai, 1989; Vodacek et al, 1997) and subsequently
biodegradation (Benner and Biddanda, 1998), as well as the rapid increases in
UV irradiation of the deeply mixed hub for global thermohaline circulation in the
Southern Ocean (Solomon, 1999), an assumption that the contemporary oceanic
DOM pool is at steady state seems questionable.
Another development that has greatly increased interest in DOM distributions
and dynamics over the past three decades has been growing recognition that dissolved organic substrates are important intermediates in rapid cycling of bioactive
elements within the ocean (Pomeroy, 1974; Azam and Hodson, 1977). This "microbial loop" from DOM to bacteria, to protists and zooplankton became evident
from measurements of heterotrophic bacterial production that typically demanded
20-40% of the local average carbon fixation rate (Azam and Fuhrman, 1984). The
only means of supplying such a large a flux of nutrients is by rapid cycling of
DOM released by a variety of processes including phytoplankton exudation, viral
lysis, and protozoan and zooplankton grazing (Jumars et ai, 1989; Nagata, 2000).
Given a global net primary production of ~50 x 10^^ g C/year, the microbial
loop would appear to pass the DOM equivalent of at least 10-20 x 10^^ g C/year.
If applied uniformly throughout the ocean, this respiration flux could turn over
the entire marine DOC pool in less than 100 years, compared to its ^"^C-based
"age" of thousands of years. The reason for this discrepancy, of course, is that a
small fraction of seawater DOC is recycled biologically in the surface ocean at
an exceedingly rapid rate. The "survivor" molecules left behind to accumulate in
the DOM pool must nevertheless be subjected to continuous and severe bacterial
pressure. Thus, any chink in their molecular armor, such as might be imparted
by photolysis (Benner and Biddanda, 1998), abiotic chemical oxidation (Sunda
and Kieber, 1994), or physical transformation into gels (Chin et ai, 1998) might
lead to rapid and efficient remineralization to CO2. Conversely, the chemical and
conformational characteristics of those organic substances that can withstand such
concerted attacks for thousands of years in the ocean DOM pool may carry the
molecular Rosetta stone for deciphering the degradation mechanisms responsible
for recycling the other 99.9% of global primary production (Hedges, 1992). Thus,
we have much to learn from both the fast- and the slow-cycling components of
ocean DOM.
There are many other reasons for continued and growing interest in the forms
and reactions of seawater DOM. For example, some molecules dissolved in seawater strongly complex trace metal ions, greatly affecting their bioavailability and
toxicity (Buffle, 1988; Kozelka and Bruland, 1998). Dissolved organic molecules
also can affect the surface properties of minerals (Stumm, 1992), act as aquatic

Why Dissolved Organics Matter

25

telemediators (Gauthier and Aubert, 1981), and change the spectral properties of
seawater (Whitehead and Vemet, 2000). Recent demonstrations that the lignin
components of seawater vary in composition and concentration with geographic
source (Opsahl and Benner, 1997; Opsahl et ai, 1999) and photodegradation history (Opsahl and Benner, 1998) point toward a future where dissolved organic
molecules will provide detailed information about the origins and physical histories
of their parent waters. Ultimately, however, the biogeochemical usefulness of any
class of chemical tracers is limited by its structural diversity (Blumer, 1976) and
the range of its sources, input functions, and chemical reactivities (Middelburg,
1989). It is clear, therefore, that the information content of organic molecules,
which also carry imbedded stable isotopic signatures and radiochemical clocks, is
unsurpassed by any other seawater component. The 10^^ diverse organic molecules
dissolved in every milliliter of seawater are the only constituents whose stored information approaches the richness needed to understand where that water has been
and what has happened within it over time. The future of oceanographic research
belongs in large part to those who can learn to read these molecular messages.

VII. WHAT DID WE LEARN?


Setbacks can be useful learning experiences. The following chapters are testimony that the oceanographic community not only persisted through the "New"
DOC experience, but also gained substantially from it in numerous ways over the
past decade. First of all, HT(C)0 (now often used without oxidation catalyst) analyzers were fundamentally a great idea and are becoming the instruments of choice
for analyses of DOC and DON in laboratories and aboard ships. With appropriate
attention to blanks and sample handling, the precision and accuracy of DOC analysis by HT(C)0 have improved to the point that meaningful comparisons among
deep-ocean waters and within surface ocean time series have become feasible. In
addition, the minimal volume requirements of these analyzers have opened the
door to multiple analyses of limited volumes of water from such sources as sediments and experimental incubations. It seems feasible that HT(C)0 analyzers will
soon be employed for rapid analysis of the stable carbon isotope composition of
DOC in individual samples or used in batch mode to obtain sufficient carbon for
^"^C analysis by accelerator mass spectrometry.
Second, we (should) also have come to appreciate the immense importance of
carefully developing and rigorously testing new analytical methods. Although fresh
concepts will continue to be the primary means of advancement in oceanographic
research, it is clear from this retrospective that new perceptions almost always ride
the back of improved methods for DOM isolation (e.g., dipped prisms, hydrophobic sorption, and tangential-flow ultrafiltration) and characterization (e.g., stable
isotopes, NMR and sensitive molecular analyses). Along with the power to make

26

John I. Hedges

such major advances, however, comes the responsibiUty to adequately test and
describe new analytical methods, pointing out their weaknesses as well as their
advantages. Editors and reviewers of papers describing new analytical methods
also carry the burden of protecting the scientific community (and authors) from
published oversights that can propagate for years to great disadvantage. Given
the reality that it is often the makers of analytical tools, rather than the wielders, who pace modem scientific advances, this rare skill seems underappreciated
overall. Many journals covering the aquatic sciences in fact exclude or strongly
discourage "analytical papers," even when the research they describe clearly has
been developed specifically to attack biogeochemical research problems. Funding
agencies and peer reviewers are often reluctant to fund strictly analytical projects or
proposals that involve innovative measurement techniques. Analytical chemistry
departments in many universities are presently being dismantled or recombined
into topical entities that no longer emphasize or adequately train students in the
basics of sound analytical methods. Although the current trend by oceanographers
toward broad interests and general skills is healthy, a critical mass of analytically
oriented investigators with sufficient chemical understanding to imagine the potential pitfalls involved with measurements of trace organic substances in an ocean
of salt will always be required.
Finally, the Seattle DOC/DON Workshop and subsequent efforts on the part of
Jon Sharp and many other oceanographers demonstrated the tremendous logistical
advantage of readily available reference samples and the power of communitywide efforts focused on a shared challenge. The crucial demonstration of a problem in DOC and DON analyses in the early 1990s (Fig. 11) came directly from
painstaking analyses by the more than 25 different research groups that participated
voluntarily in the Seattle Workshop. Fittingly, the diagnostic offsets in this data set
also indicated an experimental path toward resolution that proved to be fruitful.
The DOC/DON community also is notable as one of the few oceanographic guilds
with an organic orientation that successfully has estabhshed a system for providing
widely available reference samples (of low-DOC and deep-ocean water) for blank
testing and comparisons among (and within) individual labs. Notably, both actions
have involved close collaborations in planning and execution within the marine
scientific community and their funding organizations. That these combined efforts
have borne such extensive scientific fruit over the past decade (see following chapters) bodes well for future research on the fascinating topic of dissolved organic
molecules in the ocean.

ACKNOWLEDGMENTS
I thank the editors of this book for their encouragement and guidance. This manuscript benefited
greatly from reviews by Ron Benner, Ellen Druffel, John Farrington, Michael Peterson, Jon Sharp, and
Kenia Whitehead.

Why Dissolved Organics Matter

27

REFERENCES
Anderson, L. A. (1995). On the hydrogen and oxygen content of marine phytoplankton. Deep-Sea Res.
42,1675-1680.
Arrhenius, S. (1896). On the influence of carbonic acid in the air upon the temperature of the ground.
Phil Mag. 41, 237-276.
Azam, R, and Fuhrman, J. A. (1984). Measurement of bacterioplankton growth in the sea and its
regulation by environmental conditions. In "Heterotrophic Activity in the Sea" (J. E. Hobbie and
P. J. LeB. WiUiams, Eds.), pp. 179-196. Plenum Press, New York.
Azam, P., and Hodson, R. E. (1977). Size distribution and activity of marine microheterotrophs. Limnol.
Oceanogr. 22,492-501.
Bada, J. L., and Lee, C (1977). Decomposition and alteration of organic compounds dissolved in
seawater. Mar. Chem. 5,523-534.
Baier, R. E. (1972). Organic films on natural bodies of water: Their retrieval, identification and modes
of elimination. /. Geophys. Res. 77,5062-5075.
Baier, R. E., Goupil, D. W., Perlmutter, S., and King, R. (1974). Dominant chemical composition of
sea surface films, natural slicks and foams. /. Rech. Atmos. 8, 571-600.
Barber, R. T. (1968). Dissolved organic carbon from deep waters resists microbial oxidation. Nature
220,274-275.
Benner, R., and Biddanda, B. (1998). Photochemical transformations of surface and deep marine
dissolved organic mater: Effects on bacterial growth. Limnol. Oceanogr. 43,1373-1378.
Benner, R., Pakulski, J. D., McCarthy, M., Hedges, J. I., and Hatcher, R G. (1992). Bulk chemical
characteristics of dissolved organic matter in the ocean. Science 255,1561-1564.
Benner, R., and Strom, M. (1993). A critical evaluation of the analytical blank associated with DOC
measurements by high-temperature catalytic oxidation. Mar. Chem. 41,153-160.
Blumer, M. (1976). Polycyclic aromatic compounds in nature. Sci. Am. 234, 3 4 ^ 5 .
Bradshaw, A. L., and Brewer, P. G. (1988). High precision measurements of alkalinity and total carbon
dioxide in seawater by potentiometric titration. 1. Presence of unknown protolyte(s)? Mar Chem.
23, 69-86.
Broecker, W. S., and Peng, T-H. (1982). "Tracers in the Sea." Columbia University, Palasades, NY.
Buffle, J. (1988). "Complexation Reactions in Aquatic Systems: An Analytical Approach." Ellis
Horwood, Chichester.
Carlson, D. J., Brann, M. L., Mague, T. H., and Mayer, L. M. (1985). Molecular weight distribution
of dissolved organic materials in seawater determined by ultrafiltration: A re-examination. Mar
Chem. 55,155-171.
Chanu, J. (1959). Extraction de la substance jaune dans les eaux cotieres. Rev. Opt. Theor Instrum. 38,
569-572.
Chen, R. P., and Bada, J. L. (1989). Seawater and porewater fluorescence in the Santa Barbara Basin.
Geophys. Res. Lett. 16, 687-690.
Chin, W-C, Orellana, M. V., and Verdugo, P. (1998). Spontaneous assembly of marine dissolved
organic matter into polymer gels. Nature 391, 568-572.
Craig, H. (1971). The deep metabohsm: Oxygen consumption in abyssal ocean water. J. Geophys. Res.
76,5078-5086.
Degens, E. T. (1970). The molecular nature of nitrogenous compounds in sea water and recent marine
sediments. In "Organic Matter in Natural Waters" (D. W. Hood, Ed.), pp. 77-106. Institute of
Marine Science, Fairbanks, Alaska.
Donard, O. F. X., Lamotte, M., Belin, C , and Ewald, M. (1989). High-sensitivity fluorescence spectroscopy of Mediterranean waters using a conventional or pulsed laser excitation source. Mar
Chem. 27,111-136.

28

John I. Hedges

Druffel, E. R. M., Williams, P. M., Robertson, K., Griffin, S., Jull, A. J. T., Donahue, D., Toolin, L.,
and Linick, T. W. (1989). Radiocarbon in dissolved organic and inorganic carbon from the central
North Pacific. Radiocarbon 31, 523-532.
Duursma, E. K. (1961). Dissolved organic carbon, nitrogen and phosphorus in the sea. Neth. J. Sea
Res. 1, 1-147.
Duursma, E. K. (1963). The production of dissolved organic matter in the sea, as related to the primary
production of organic matter. Neth. J. Sea Res. 2, 85-94.
Duursma, E. K. (1965). The dissolved organic constituents of seawater. In "Chemical Oceanography"
(J. P Riley and G. Skirrow, Eds.), Vol.1, pp. 433-475. Academic Press, London.
Ewing, G. (1950). Slicks, surface films and internal waves. /. Mar. Res. 9,161-187.
Farrington, J. (1992). Macromolecular organic matter working group report. Mar Chem. 39, 39-50.
Flaig, W. (1964). Effects of microorganisms in the transformation of Ugnin to humic substances.
Geochim. Cosmochim. Acta 28, 1523-1535.
Fogg, G. E., and Boalch, G. T. (1958). Extracellular products of pure cultures of a brown alga. Nature
181,789-791.
Fox, L. E. (1983). The removal of dissolved humic acid during estuarine mixing. Estuatine Coastal
Shelf Sci.U,A2>\-4A0.
Fu, T, and Pocklington, R. (1983). Quantitative adsorption of organic matter from seawater on sohd
matrices. Mar Chem. 13, 255-264.
Garrasi, C., Degens, E. T, and Mopper, K. (1979). Amino acid composition of sea water obtained
without desalting. Mar Chem. 8,71-85.
Garrett, W. D. (1965). Collection of slick forming material from the surface of the sea. Limnol.
Oceanogr 10,602-605.
Garrett, W. D. (1967). The organic chemical composition of the ocean surface. Deep-Sea Res. 14,
221-227.
Garrett, W. D. (1970). Organic chemistry of natural sea surface films. In "Organic Matter in Natural
Waters" (D. W Hood, Ed.), pp. 469^77. Institute of Marine Science, Fairbanks, AK.
Garrett, W. D., and Barger, W R. (1974). Sampling and determining the concentration of film-forming
organic constituents of the air-water interface. Nav. Res. Lab. Memo. Rep. 2852.
Gauthier, M. J., and Aubert, M. (1981). Chemical telemediators in the marine environment. In "Marine
Organic Chemistry" (E. K. Duursma and R. Dawson, Eds.), pp. 225-257. Elsevier, Amsterdam.
Gillam, A. H., and Wilson, M. A. (1985). Pyrolysis-GC-MS and NMR studies on dissolved seawater
humic substances and isolates of a marine diatom. Org. Geochem. 8, 15-25.
Gordon, D. C , and Sutcliffe, W. H. (1973). A new dry combustion method for the simultaneous
determination of total organic carbon and nitrogen in seawater. Mar Chem. 1,231-244.
Hansen, D. A. (1993). Results and observations from the measurement of DOC and DON in seawater
using a high-temperature catalytic oxidation technique. Mar Chem. 41,195-202.
Harvey, G. R. (1966). Microlayer collection from the sea surface: A new method and initial results.
Limnol. Oceanogr 11, 608-614.
Harvey, G. R., Boran, D. A., Chesal, L. A., and Tokar, J. M. (1983). The structure of marine fulvic and
humic acids. Mar Chem. 12, 119-132.
Harvey, G. R., and Burzell, L. A. (1972). A simple microlayer method for small samples. Limnol.
Oceanogr 17, 156-157.
Hayase, K., Tsubota, H., and Sunada, I. (1988). Vertical distribution of fluorescent organic matter in
the North Pacific. Mar Chem. 25, 373-381.
Hayase, K., Yamamoto, M., Nakazawa, I., and Tsubota, H. (1987). Behavior of natural fluorescence in
Sagami Bay and Tokyo Bay, JapanVertical and lateral distributions. Mar Chem. 20, 265-276.
Hedges, J. I. (1992). Global biogeochemical cycles: progress and problems. Mar Chem. 39, 67-93.
Hedges, J. I., Bergamaschi, B. A., and Benner, R. (1993). Comparative analyses of DOC and DON in
natural waters. Mar Chem. 41, 121-134.

Why Dissolved Organics Matter

29

Hedges, J. L, and Lee, C. (1993). Measurement of dissolved organic carbon and nitrogen in natural
waters. Mar. Chem. 41, 1-290.
Hodge, J. E. (1953). Chemistry of browning reactions in model systems. Agnc. Food Chem. 1,828-943.
Hood, D. W. (1970). "Organic Matter in Natural Waters." Institute of Marine Science, Fairbanks, AK.
Holm-Hansen, O., Strickland, J. D. H., and Williams, P. M. (1966). A detailed analysis of biologically
important substances in a profile off Southern California. Limnol. Oceanogr. 11, 548-561.
Houghton, J. T., Meira Filho, L.G., Callander, B. A., Harris, N., Kattenberg, A., and Maskell, K. (1996).
"Climate Change 1995: The Science of Climate Change." Cambridge University Press, New York.
Hunter, K. A., and Liss P. S. (1981). Organic sea surface films. In "Marine Organic Chemistry"
(E. K. Duursma and R. Dawson, Eds.), pp. 259-298. Elsevier, Amsterdam.
Ittekkot, V. (1982). Variations of dissolved organic matter during a plankton bloom: Qualitative aspects
based on sugar and amino acid analyses. Mar. Chem. 11,143-158.
Ittekkot, v., Brockmann, U., Michaelis, W, and Degens, E. T. (1981). Dissolved free and combined
carbohydrates during a phytoplankton bloom in the northern North Sea. Mar Ecol. Prog. Sen 4,
299-305.
Jaanasch, H. W (1967). Growth of marine bacteria at limiting concentrations of organic carbon in
seawater. Limnol. Oceanogr 12,264-271.
Jackson, G. A., and WiUiams, P. M. (1985). Importance of dissolved organic nitrogen and phosphorus
to biological nutrient cycling. Deep-Sea Res. 2>1, 223-235.
Jarvis, N. L. (1967). Adsorption of surface active material at the air-sea interface. Limnol Oceanogr
12,213-221.
Jeffrey, L. M., and Hood, D. W. (1958). Organic matter in seawater; An evaluation of various methods
for isolation. /. Mar Res. 17,247-271.
Johnston, R. (1955). Biologically active compounds in the sea. /. Mar Biol. Assoc UK 34,185-195.
Jumars, P. A., Penry, D. L., Baross, J. A., Perry, M. J., and Frost, B. W (1989). Closing the microbial
loop: dissolved carbon pathway to heterotrophic bacteria from incomplete ingestion, digestion and
absorption in animals. Deep-Sea Res. 36,483^95.
Kalle, K. (1937). Nahrstoff Untersuchungen als hydrographisches Hilfsmittel zur Unterscheiding von
Wasserkorpen. Annu. Hydrogr Berlin 65, 276-282.
Kalle, K. (1949). Fluoreszenz und Gelbstoff im Bottnischen und Finnischen Meerbusen. Dtsch.
Hydrogr Z. 2,111-124.
Kalle, K. (1966). The problem of gelbstoff in the sea. Oceanogr Mar Biol. Ann. Rev. 4, 91-104.
Karl, D. M., Tien, G., Dore, J., and Winn, C. D. (1993). Total dissolved nitrogen and phosphorus
concentration at US-JGOFS Station ALOHA: Redfield reconciliation. Mar Chem. 41, 203-208.
Kay, H. (1954). Untersuchungen zur Menge und Verteilung der organischen Substanz im Meerwasser.
Keil. Meeresf. 10,202-213.
Keeling, C. D. (1973). Industrial production of carbon dioxide from fossil fuels and limestone. Tellus
25,174-198.
Keeling, C. D., Whorf, T. R, Whalen, M., and van der PUcht, J. (1995). Interannual extremes in the
rate of rise of atmospheric carbon dioxide since 1980. Nature 375, 666-670.
Kerr, R. A., and Quinn, J. G. (1975). Chemical studies on the dissolved organic matter in seawater.
Isolation and fractionation. Deep-Sea Res. 22,107-116.
Kieber, D. J., McDaniel, J., and Mopper, K. (1989). Photochemical source of biological substrates in
sea water: Implications for carbon cycling. Nature. 341,637-639.
Kieber, D. J., and Mopper, K. (1987). Photochemical formation of glyoxylic and pyruvic acids in
seawater. Mar Chem. 21,135-149.
Koike, I., and Tupas, L. (1993). Total dissolved nitrogen in the North Pacific assessed by a hightemperature combustion method. Mar Chem. 41,209-214.
Knauer, G. A., Martin, J. H., and Bruland, K. W. (1979). Fluxes of particulate carbon, nitrogen and
phosphorus in the upper water column of the northeast Pacific. Deep-Sea Res. 26, 97-108.

30

John I. Hedges

Kozelka, P. B., and Bruland, K. W. (1998). Chemical speciation of dissolved Cu, Zn, Cd, Pb, in
Narragansett Bay, Rhode Island. Mar. Chem. 60,267-282.
Kramer, C. J. M. (1979). Degradation by sunlight of dissolved fluorescing substances in the upper
layers of the eastern Atlantic Ocean. Neth. J. Sea Res, 13, 325-329.
Krogh, A. (1934a). Conditions of life in the ocean. Ecol. Monogr. 4,421-429.
Krogh, A. (1934b). Conditions of Ufe at great depths in the ocean. Ecol. Monogr. 4,430^39.
Krogh, A., and Keys, A. (1934). Methods for the determination of dissolved organic carbon and nitrogen
in sea water. Biol. Bull. 67, 132-144.
Laane, R. W. P. M. (1982). Influence of pH on the fluorescence of dissolved organic matter. Mar Chem.
11,395-401.
Laane, R. W. P. M. (1984). Comment on the structure of marine fulvic and humic acids. Mar Chem.
15, 85-87.
Lee, C , and Bada, J. L. (1975). Amino acids in Equatorial Pacific Ocean water. Earth Planet. Sci. Lett.
26,61-68.
Lee, C , and Bada, J. L. (1977). Dissolved amino acids in the Equatorial Pacific, Sargasso Sea and
Biscayne Bay. Limnol. Oceanogr 22, 502-510.
Liebezeit, G., Bolter, M., Brown, I. F., and Dawson, R. (1980). Dissolved free amino acids and carbohydrates at pycnocline boundaries in the Sargasso Sea and related microbial activity. Oceanol
Acta 3, 357-362.
Liss, P. S. (1975). The chemistry of the sea surface microlayer. In "Chemical Oceanography"
(J. P. Riley and G. Skirrow, Eds.), Vol. 2, pp. 193-243. Academic Press, London.
Little, B., and Jacobs, J. (1985). A comparison of two techniques for the isolation of adsorbed dissolved
organic material from seawater. Org. Geochem. 8, 27-33.
Maclntyre, F. (1966). Bubbles: A boundary layer microtome for micron-thick samples of a water
surface. J. Phys. Chem. 72, 589-592.
Mackinnon, M. D. (1978). A dry oxidation method for analysis of the TOC in seawater. Mar Chem.
7, 17-37.
Maillard, L. C. (1913). Formation de matieres humiques par action de polypeptides sur sucres. CR
Acad. Sci. 156, 148-149.
Mantoura, R. F. C , and Riley, J. P. (1975). The analytical concentration of humic substances from
natural waters. Anal. Chem. Acta 76, 97-106.
Mantoura, R. F. C , and Woodward, E. M. S. (1983). Conservative behavior of riverine dissolved organic
carbon in the Severn Estuary: Chemical and geochemical implications. Geochim. Cosmochim. Acta
47,1293-1309.
Martin, J. H., Knauer, G. A., Karl, D. M., and Broenkow, W. W. (1987). VERTEX: Carbon cycling in
the northeast Pacific. Deep-Sea Res. 34, 267-285.
Maurer, L. G. (1976). Organic polymers in seawater: Changes with depth in the Gulf of Mexico.
Deep-Sea Res. 23, 1059-1064.
McCave, I. N. (1975). Vertical flux of particles in the ocean. Deep-Sea Res. 22,491-502.
Menzel, D. W. (1964). The distribution of dissolved organic carbon in the western Indian Ocean.
Deep-Sea Res. 11,757-765.
Menzel, D. W., and Ryther, J. H. (1968). Organic carbon and the oxygen minimum in the South Atlantic
Ocean. Deep-Sea Res. 15, 327-337.
Menzel, D. W., and Ryther, J. H. (1970). Distribution and cycling of organic matter in the oceans. In
"Organic Matter in Natural Waters" (D. W Hood, Ed.), pp. 31-54. Institute of Marine Science,
Alaska.
Menzel, D. W, and Vaccaro, R. F (1964). The measurement of dissolved and particulate organic carbon
in the sea. Limnol. Oeanogr 9, 138-142.
Meybeck, M. (1982). Carbon, nitrogen and phosphorus transport by world rivers. Am. J. Sci. 282,
401^50.

Why Dissolved Organics Matter


Meyers-Schulte, K. J., and Hedges, J. I. (1986). Molecular evidence for a terrestrial component of
organic matter dissolved in ocean water. Nature 321,61-63.
Middelburg, J. J. (1989). A simple rate model for organic matter decomposition in marine sediments.
Geochim. Cosmochim. Acta 53, 1577-1581.
Momzikoff, A., Santus, R., and Giraud, M. (1983). A study of the photosensitizing properties of
seawater. Mar. Chem. 12, 1-14.
Mopper, K:, Dawson, R., Liebezeit, G., and Hansen, H-R (1980). Borate complex ion exchange chromatography with fluorimetric detection for determination of saccharides. Anal. Chem. 52, 20182022.
Mopper, K., and Stahovec, W. L. (1986). Sources and sinks of low molecular weight organic carbonyl
compounds in seawater. Mar Chem. 19, 305-321.
Morris, R. J. (1974). Lipid composition of surface films and zooplankton from the eastern Mediterranean. Mar Pollut. Bull. 5, 105-109.
Nagata, T. (2000). Production mechanisms of dissolved organic matter. In "Microbial Ecology of the
Oceans" (D. L. Kirchman, Ed.), pp. 121-152. Wiley-Liss, New York.
Ogawa, H., Fukuda, R., and Koike, I. (1999). Vertical distributions of dissolved organic carbon and
nitrogen in the Southern Ocean. Deep-Sea Res. 746, 1809-1826.
Ogawa, H., and Ogura, N. (1992). Comparison of two methods for measuring dissolved organic carbon
in sea water. Nature 356, 696-698.
Ogura, N. (1975). Further studies on decomposition of dissolved organic matter in coastal seawater.
Mar Biol. 31, lOl-ni.
Ogura, N. (1977). High molecular weight organic matter in seawater. Mar Chem. 5, 535-549.
Ogura, N. (1970). The relation between dissolved organic carbon and apparent oxygen utilization in
the Western North Pacific. Deep-Sea Res. 17, 221-231.
Opsahl, S., and Benner, R. (1997). Distribution and cycling of terrigenous dissolved organic matter in
the ocean. Nature 386,480^82.
Opsahl, S., and Benner, R. (1998). Photochemical reactivity of dissolved lignin in river and ocean
waters. Limnol. Oceanogr 43,1297-1304.
Opsahl, S., Benner, R., and Amon, R. M. W. (1999). Major flux of terrigenous dissolved organic matter
through the Arctic Ocean. Limnol. Oceanogr 44, 2017-2023.
Plunkett, M. A., and Rakestraw, N. W. (1955). Dissolved organic matter in the sea. Deep-Sea Res. 3,
12-14.
Pomeroy, L. R. (1974). The ocean's food web, a changing paradigm. Bioscience 24,499-504.
Putter, A. (1909). Die Emahrung der Wassertiere und der Stoffhaushalt der Gewasser. Jena,
1909.
Quay, P. (1997). Was a carbon balance measured in the equatorial Pacific during JGOFS? Deep-Sea
Res. 44,1765-1781.
Raben, E. (1910). 1st organischen Kohlenstoff in nennenswerter Menge im Meerwasser gelost vorhanden? Wiss. Meeresunt. Abt. Kiel NF11, 111-117.
Redfield, A. C , Ketchum, B. H., and Richards, F. A. (1963). The influence of organisms on the
composition of sea water. In "The Sea" (M. N. Hill, Ed.), pp. 26-77. Interscience, New York.
Riley, J. P., and Taylor, D. (1969). The analytical concentration of traces of dissolved organic materials
from sea water with Amberlite XAD-1 resin. Anal. Chim. Acta 46, 307-309.
Sakugawa, H., and Handa, N. (1985). Isolation and chemical characterization of dissolved and particulate polysaccharides in Mikawa Bay. Geochim. Cosmochim. Acta 49, 1185-1193.
Sakugawa, H., Handa, N., and Ohta, K. (1985). Isolation and characterization of low molecular weight
carbohydrates dissolved in seawater. Mar Chem. 17, 341-362.
Sharp, J. H. (1973a). Size classes of organic carbon in seawater. Limnol. Oceanogr 18,441-447.
Sharp, J. H. (1973b). Total organic carbon in seawaterComparison of measurements using persulfate
oxidation, and high temperature combustion. Mar Chem. 1, 211-229.

31

32

John I. Hedges

Sieburth, J. McN. (1969). Studies on algal substances in the sea. III. The production of extracellular
organic matter by littoral marine algae. J. Exp. Mar. Biol. Ecol. 3, 290-309.
Sieburth, J. McN., and Conover, J. T. (1965). SUcks associated with the Tricodesmium blooms in the
Sargasso Sea. Nature 295, 830-831.
Sieburth, J. McN., and Jensen, A. (1968). Studies on algal substances in the sea. I. Gelbstoff (humic
material), in terrestrial and marine water. /. Exp. Mar. Biol. Ecol. 2,174-189.
Sieburth, J. McN., and Jensen, A. (1969). Studies on algal substances in the sea. II. The formation of
Gelbstoff (humic material) by exudates of phaeophyta. /. Exp. Mar Biol. Ecol. 3,275-289.
Siegel, A. (1971). Metal-organic interactions in the marine environment. In "Organic Compounds in
Aquatic Envu-onments" (S. D. Faust and J. V. Hunter, Eds.), pp. 203-223. Dekker, New York.
Siegenthaler, U., and Sarmiento, J. L. (1993). Atmospheric carbon dioxide and the ocean. Nature 365,
119-125.
Skopinstev, B. A. (1966). Some considerations on the distribution and state of organic matter in ocean
water. Oceanology 6, 361-368.
Slowley, J. F., Jeffrey, L. M., and Hood, D. W. (1959). Characterization of the ethyl acetate extractable
organic material of sea water. Int. Ocean. Cong. Prepr 935-937.
Solomon, S. (1999). Stratospheric ozone depletion: A review of concepts and history. Rev. Geophys.
yi, 215-3,16.
Starikova, N. D. (1970). Vertical distribution patterns of dissolved organic carbon in sea water and
interstitial solutions. Oceanology 10, 796-807.
Stuermer, D. H., and Harvey, G. R. (1974). Humic substances from seawater. Nature 250,480-481.
Stuermer, D. H., and Harvey, G. R. (1977). The isolation of humic substances and alcohol-soluble
organic matter from seawater. Deep-Sea Res. 24, 303-309.
Stuermer, D. H., and Harvey, G. R. (1978). Structural studies on marine humus: A new reduction
sequence for carbon skeleton determination. Mar Chem. 6, 55-70.
Stuermer, D. H., and Payne, J. R. (1976). Investigations of seawater and terrestrial humic substances
with carbon-13 and proton nuclear magnetic resonance. Geochim. Cosmochim. Acta 40, 11091114.
Stunrni, W. (1992). "Chemistry of the Solid-Water Interface." Wiley, New York.
Sugimura, Y, and Suzuki, Y (1983). Amino acids dissolved in the western North Pacific water. Pap.
Meteorol. Geophys. 34, 267-289.
Sugimura, Y, and Suzuki, Y (1988). A high temperature catalytic oxidation method for the determination of non-volatile dissolved organic carbon in seawater by direct injection of a liquid sample.
Mareorol. Chem. 24,105-131.
Sunda, W. G., and Kieber, D. K. (1994). Oxidation of humic substances by manganese oxides yields
low-molecular-weight organic substrates. Nature 367,62-64.
Suzuki, Y (1993). On the measurement of DOC and DON in seawater. Mar Chem. 41, 287-288.
Suzuki, Y, Sugimura, Y, and Itoh, T. (1985). A catalytic oxidation method for the determination of
total nitrogen dissolved in seawater. Mar Chem. 16, 83-97.
Tanoue, E. (1992). Vertical distribution of dissolved organic carbon in the North Pacific as determined
by the high temperature catalytic oxidation method. Earth Planet. Sci. Lett. I l l , 201-216.
Toggweiler, J. R. (1989a). Is the downward dissolved orgaic matter (DOM) flux important in carbon
transport? In "Productivity in the Ocean: Present and Past" (W. H. Berger, V. S. Smetacek, and
G. Wafer, Eds.), pp. 65-83. Wiley, New York.
Toggweiler, J. R. (1989b). Deep-sea carbon, a burning issue. Nature 334,468.
Vallentyne, J. R. (1957). The molecular nature of organic matter in lakes and oceans, with lesser
reference to sewage and terrestrial soils. J. Fish. Res. Bd. Can. 14, 33-82.
Van Hall, C. E., Safranki, J., and Stenger, V. A. (1963). Rapid combustion method for the determination
of organic substances in aqueous solutions. Anal. Chem. 35, 315-319.

Why Dissolved Organics Matter

33

Vodacek, A., Blough, N. V., DeGrandpre, M. D., Peltzer, E. T., and Nelson, R. K. (1997). Seasonal
variation of CDOM and DOC in the Middle Atlantic Bight: Terrestrial inputs and photooxidation.
Limnol. Oceanogr. 42, 674-686.
Waksman, S. A., and Carey, C. L. (1935). Decomposition of organic matter in sea water by bacteria.
/. Bacteriol 29,531-543.
Walsh, T. W. (1989). Total dissolved nitrogen in seawater: A new high-temperature combustion method
and a comparison with photo-oxidation. Mar. Chem. 26, 295-311.
Whitehead, K., and Vemet, M. (2000). Influence of mycosporine-like amino acids (MAAs) on the
UV absorption of particulate and dissolved organic matter in La JoUa Bay. Limnol Oceanogr. 45,
1788-1796.
Willey, J. D., and Atkinson, L. P. (1982). Natural fluorescence as a tracer for distinguishing between
piedmont and coastal plain river water in the nearshore waters of Georgia and North Carolina.
Estuarine Coastal Shelf ScL 14,49-59.
Wilhams, P. J. le B. (1975). Biological and chemical aspects of dissolved organic matter in sea water.
In "Chemical Oceanography" (J. P. Riley and G. Skirrow, Eds.), Vol. 2, pp. 301-363. Academic
Press, New York.
Wilhams, P. M. (1968a). Organic and inorganic constituents of the Amazon River. Nature 218,937-938.
Wilhams, P. M. (1968b). Stable carbon isotopes in the dissolved organic matter of the sea. Nature 219,
152-153.
Williams, P. M. (1969a). The wet oxidation of organic matter in seawater. Limnol. Oceanogr 14,
292-297.
Williams, P. M. (1969b). The determination of dissolved organic carbon in seawater: A comparison of
two methods. Limnol Oceanogr 14, 297-298.
Wilhams, P. M. (1971). The distribution and cycling of organic matter in the ocean. In "Organic
Compounds in the Aquatic Environment" (S. D. Faust and J. V. Hunter, Eds.), pp. 145-163. Dekker,
New York.
Williams, P. M., and Carlucci, A. F. (1976). Bacterial utihsation of organic matter in the deep sea.
A^fl^Mr^ 262, 810-811.
Wilhams, P. M., and Druffel, E. R. M. (1987). Radiocarbon is dissolved organic matter in the central
North Pacific Ocean. Nature 339,246-248.
Wilhams, P. M., and Druffel, E. R. M. (1988). Dissolved organic matter in the ocean: comments on a
controversy. Oceanography 1, 14-17.
Williams, P. M., and Gordon, L. I. (1970). Carbon-13:carbon-12 ratios in dissolved and particulate
organic matter in the sea. Deep-Sea Res. 17,19-27.
Williams, P. M., Oeschger, H., and Kinney, P. (1969). Natural radiocarbon activity of dissolved organic
carbon in the Northeast Pacific Ocean. Nature 224, 256-258.
Wilson, D. P., and Armstrong, F. A. (1952). Further experiments on biological differences between
natural sea waters. /. Mar Biol Assn. UK 31, 335-349.
Wilson, M. A., Gillam, A. H., and Collins, P. J. (1983). Analysis of the structure of dissolved marine
humic substances and their phytoplanktonic precursors by ^H and ^^C nuclear magnetic resonance.
Chem. Geol 40, 187-201.
Wilson, M. A., Philp, R. P, Gillam, A. H., Gilbert, T. D., and Tate, K. R. (1983). Comparison of the structures of humic substances from aquatic and terrestrial sources by pyrolysis gas chromatographymass spectrometry. Geochim. Cosmochim. Acta 47,497-502.
Zepp, R. G., Schlotzhauer, P. F., and Sink, R. M. (1985). Photosensitized transformations involving
electronic energy transfer in natural waters: Role of humic substances. Environ. ScL Technol 19,
74-81.

You might also like