throbber
Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`Interannual and between species comparison of the lipids, fatty acids
`and sterols of Antarctic krill from the US AMLR Elephant Island
`survey area
`
`a,b
`c
`c,d
`a,
`Charles F. Phleger *, Matthew M. Nelson , Ben D. Mooney , Peter D. Nichols
`
`a D
`
`Department of Biology, San Diego State University, San Diego, CA 92182, USA
`epartment of Zoology, University of Tasmania, Hobart, Tasmania 7001, Australia
`c
`CSIRO Marine Research, Hobart, Tasmania 7001, Australia
`d
`Antarctic CRC, Hobart, Tasmania 7001, Australia
`
`b
`
`Received 15 July 2001; received in revised form 26 November 2001; accepted 15 December 2001
`
`Abstract
`
`Antarctic euphausiids, Euphausia superba, E. tricantha, E. frigida and Thysanoessa macrura were collected near
`Elephant N Island c during 1997 and 1998. Total lipid was highest in E. superba small juveniles (16 mg g wet mass),
`y1
`ranging from 12 to 15 mg in other euphausiids. Polar lipid (56–81% of total lipid) and triacylglycerol (12–38%) were
`the major lipids with wax esters (6%) only present in E. tricantha. Cholesterol was the major sterol (80–100% of total
`sterols) with desmosterol second in abundance (1–18%). 1997 T. macrura and E. superba contained a more diverse
`sterol profile, including 24-nordehydrocholesterol (0.1–1.7%), trans-dehydrocholesterol (1.1–1.5%), brassicasterol (0.5–
`1.7%), 24-methylenecholesterol (0.1–0.4%) and two stanols (0.1–0.2%). Monounsaturated fatty acids included primarily
`18:1(ny9)c (7–21%), 18:1(ny7)c (3–13%) and 16:1(ny7)c (2–7%). The main saturated fatty acids in krill were
`16:0 (18–29%), 14:0 (2–15%) and 18:0 (1–13%). Highest eicosapentaenoic acid wEPA, 20:5(ny3)x and docosahexaenoic
`acid wDHA, 22:6(ny3)x occurred in E. superba (EPA, 15–21%; DHA, 9–14%), and were less abundant in other krill.
`E. superba is a good source of EPA and DHA for consideration of direct or indirect use as a food item for human
`consumption. Lower levels of 18:4(ny3) in E. tricantha, E. frigida and T. macrura (0.4–0.7% of total fatty acids) are
`more consistent with a carnivorous or omnivorous diet as compared with herbivorous E. superba (3.7–9.4%). The
`polyunsaturated fatty acid (PUFA) 18:5(ny3) and the very-long chain (VLC-PUFA), C and C PUFA, were not
`26
`28
`present in 1997 samples, but were detected at low levels in most 1998 euphausiids. Interannual differences in these
`biomarkers suggest greater importance of dinoflagellates or some other phytoplankton group in the Elephant Island area
`during 1998. The data have enabled between year comparisons of trophodynamic interactions of krill collected in the
`Elephant Island region, and will be of use to groups using signature lipid methodology. 䊚 2002 Elsevier Science Inc.
`All rights reserved.
`
`Keywords: Antarctica; Euphausia superba; Fatty acid; Krill; Sterol; Triacylglycerol
`
`1. Introduction
`
`Antarctic krill (Euphausia superba Dana) pro-
`vide 30–90% of the diet for marine carnivores in
`
`*Corresponding author. Tel.: q1-760-632-9447; fax: q1-
`619-594-5676.
`E-mail address:
`phleger@sunstroke.sdsu.edu (C.F. Phleger).
`
`the Southern Ocean and have an estimated standing
`stock biomass of about 500 million metric tons
`(Ross and Quetin, 1988). The global fishery for
`krill peaked prior to 1990 at about 500 thousand
`tons per year. The current fishery is about 100
`thousand tons per year. This present low level is
`due primarily to lack of demand (Nicol and Endo,
`1999). The success of Antarctic krill reflects their
`
`1096-4959/02/$ - see front matter 䊚 2002 Elsevier Science Inc. All rights reserved.
`PII: S 1 0 9 6 - 4 9 5 9 Ž 0 2 . 0 0 0 2 1 - 0
`
`RIMFROST EXHIBIT 1173 Page 0001
`
`

`

`734
`
`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`ability to adapt to major differences in seasonal
`food supply. Krill are mainly herbivorous and feed
`on phytoplankton in the summer. In winter, krill
`feed on ice algae and probably bacteria and marine
`detritus, as well as depleting body protein for
`energy (Virtue et al., 1996). High recruitment and
`early spawning occur during years where there is
`a high pack ice concentration of long duration
`(Siegel and Loeb, 1995).
`The Southern Ocean is a complex ecosystem
`including planktonic herbivores (krill, salps, cope-
`pods) fed upon by squid, seals, baleen whales,
`birds and fish (Quetin and Ross 1991; Loeb et al.,
`1997). Lipid biomarkers,
`including lipid class,
`sterol and fatty acid spectra, have been used
`increasingly to help understand marine trophodyn-
`amics (e.g. Nichols et al., 1984; Sargent et al.,
`1987). Recent studies using the lipid signature
`approach have helped to clarify aspects of Antarc-
`tic ecology not visible by conventional techniques
`(Falk-Petersen et al., 1999; Phleger et al., 1999,
`2000; Nelson et al., 2000, 2001).
`There is a growing literature on Antarctic krill
`(E. superba) lipids (Saether et al., 1985; Quetin
`and Ross, 1991; Pond et al., 1995; Virtue, 1995;
`Hagen et al., 1996; Virtue et al., 1996; Mayzaud
`et al., 1998; Cripps et al., 1999). Fatty acids have
`been used extensively as bioindicators in E. super-
`ba (e.g. Virtue et al., 1993b), with sterols used to
`a lesser extent. More limited detail is available for
`E. tricantha and E. frigida (Phleger et al., 1998)
`and Thysanoessa macrura (Reinhardt and Van
`Vleet, 1986; Kattner et al., 1996; Falk-Petersen et
`al., 1999). However, few studies have examined
`interannual changes in krill lipid composition for
`animals collected from specific locations.
`The purpose of this study is therefore to examine
`comparatively lipid classes, specific sterols and
`fatty acid biomarkers of Antarctic krill E. superba
`and other less-studied euphausiids, including E.
`tricantha, E. frigida and T. macrura. The availa-
`bility of krill collected in both 1997 and 1998 in
`the oceanographic region near Elephant Island was
`possible as the area has been intensively surveyed
`for zooplankton by the United States Antarctic
`Marine Living Resources (US AMLR) Program.
`It is noted for high biological productivity and rich
`krill populations that experienced major fluctua-
`tions in density depending on sea ice extent and
`temperature (Loeb et al., 1997; Brierly et al.,
`1999). According to Loeb et al. (1998), 1993 and
`1998 were ‘salp years’ with Salpa thompsoni
`
`numerically dominant (56–89% of total zooplank-
`ton), post-larval Thysanoessa macrura second in
`abundance (8–14%) followed by post-larval krill
`and copepods. The years 1995 and 1996 were
`‘copepod years’ with copepods (presumably Metri-
`dia gerlachei) dominant taxa. Larval T. macrura
`ranked fourth and second in abundance during
`February–March 1995 and 1996, respectively. Post
`larval T. macrura and chaetognaths switched in
`order of abundance during these years. In contrast,
`S. thompsoni ranked sixth and eighth and contrib-
`uted -1.5% of
`total zooplankton. February–
`March 1994 and March 1997 appear
`to be
`‘transition periods (years)’ between ‘copepod
`years’ and ‘salp years’. During ‘transition years’,
`copepods were numerically dominant, followed in
`order by S. thompsoni, post-larval T. macrura and
`Euphausia frigida (Loeb et al., 1998). ‘Salp years’
`appear to correlate with the 4–5-year period of the
`Antarctic Circumpolar wave, which propagates
`changes in sea surface temperatures and wind
`stress direction (White and Peterson, 1996). An
`objective of our study is to utilize the lipid com-
`position data to help clarify trophodynamics,
`including examining possible interannual changes
`in feeding.
`
`2. Materials and methods
`
`2.1. Sample description
`
`Krill were obtained as part of the AMLR Field
`Study conducted annually in the Elephant Island
`region of the Antarctic Peninsula located between
`60–62.58S and 53–598W. (Loeb et al., 1997;
`Martin, 1997y8). Specimens were collected by
`Isaacs–Kidd midwater trawl fitted with a 505 mm
`mesh plankton net from the RyV Yuzhmorgeolo-
`giya during January and February, 1997 and 1998.
`The net was obliquely towed to 170 m depth for
`approximately 30 min at a speed of 2 knots, or to
`10 m above the bottom in shallower waters.
`Samples were frozen in liquid nitrogen immediate-
`ly after sorting on board ship. They were then
`transported frozen (dry ice) by air to CSIRO
`Marine Research, in Hobart, Tasmania, where they
`were maintained at y70 8C prior to analysis. 1997
`krill were 1.00–1.34 g fresh mass for E. superba
`(five and six pooled individuals for each sample,
`respectively), 0.01 g for E. tricantha (one only)
`and 0.20–0.70 g for T. macrura (20 and 70 pooled
`for each sample, respectively). Krill for 1998 were
`
`RIMFROST EXHIBIT 1173 Page 0002
`
`

`

`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`735
`
`1.29–1.39 g fresh mass for E. superba adults (two
`pooled for each sample), 0.28–0.38 g for E.
`superba juveniles (large), 0.22–0.31 g for E.
`superba juveniles (small) (two pooled for each
`sample), 0.70–0.86 g for E. tricantha (three to
`four pooled for each sample) and 0.24–0.28 g for
`E. frigida (six to eight pooled for each sample).
`
`2.2. Lipid extraction and analysis
`
`Oil (lipid) analyses were conducted using meth-
`ods described in Nichols et al. (1998a,b,c) and
`Phleger et al. (1998, 1999). Briefly, samples were
`extracted using a single phase Bligh and Dyer
`(1959) procedure. Oil yield was determined grav-
`imetrically. An aliquot of the oil was analyzed by
`TLC-FID to determine lipid class composition
`(Volkman and Nichols, 1991). Fatty acid and sterol
`profiles were obtained by capillary GC and GC-
`MS analysis
`following trans-methylation and
`saponification of aliquots of the oil.
`
`2.3. Statistical analyses
`
`y1
`Fatty acid profiles (mg g wet mass) of indi-
`vidual samples were compared by cluster analysis
`using Pearson’s correlation coefficient and average
`linkage (Fig. 1). Pearson’s correlation coefficient
`and non-metric multidimensional scaling (MDS)
`were also used to compare FA profiles (full pro-
`files) in two dimensions, using the Kruskal Loss
`Function. All multivariate analyses were conducted
`using SYSTAT 9 (SYSTAT, Inc., Evanston, IL,
`USA).
`
`3. Results
`
`3.1. Lipid classes
`
`Polar lipids were the major lipid class in all
`euphausiids in 1998 (56–81% of total lipid; Table
`1). Triacylglycerols (TAG) were the second most
`abundant lipid class (22–38% in Euphausia super-
`ba, 16% in E. tricantha and 12% in E. frigida).
`E. tricantha was the only euphausiid with wax
`esters (WE) (6% of total lipid). WE were below
`detection (-0.5%) in the other species. In all
`animals, sterols (ST) accounted for 4–7% of total
`lipids, with low free fatty acids (FFA) (1–3%).
`Total lipid was highest in E. superba small juve-
`y1
`niles (15.9 mg g
`wet mass) and ranged from
`
`y1
`
`in all other euphausiids (Table
`
`12.4–14.6 mg g
`1).
`
`3.2. Sterols
`
`Cholesterol was the major ST in all krill (80–
`100% of total ST) with highest values in E.
`tricantha (94–100%) and E. frigida (97%) (Table
`2). Desmosterol was the second most abundant ST
`(0–18% of total ST), and was highest in 1997 E.
`superba (18%), but markedly less in 1998 E.
`superba (2–4%). Although there was 6% desmos-
`terol in 1997 E. tricantha, none was detected in
`1998 E. tricantha (Table 2). Desmosterol was also
`not detected in E. frigida. The 1997 Thysanoessa
`macrura and E. superba samples were the only
`samples with ST other than cholesterol and des-
`mosterol. These included primarily 24-nordehydro-
`cholesterol (0.1–1.7%),
`trans-dehydrocholesterol
`(1.1–1.5%), brassicasterol (0.5–1.7%) and 24-
`methylenecholesterol (0.1–0.4%). Low levels of
`stanols (0.1–0.2% cholestanol and brassicastanol)
`were only detected in 1997 T. macrura and E.
`superba, but not in the 1998 krill (Table 2).
`
`3.3. Fatty acids
`
`Polyunsaturated fatty acids (PUFA) were 39–
`45% of the total FA in all E. superba samples, but
`were somewhat lower in 1998 E. tricantha (30%)
`and 1998 E. frigida (31%), with 11% total PUFA
`frigida (Table 3). Eicosapentaenoic
`in 1997 E.
`acid wEPA, 20:5(ny3)x and docosahexaenoic acid
`wDHA, 22:6(ny3)x were the two major PUFA in
`all samples (Table 3). Highest EPA and DHA
`values were detected in E. superba (15–21% and
`9–14%, respectively). These two PUFA were gen-
`erally lower in abundance in E. frigida (4–18%
`EPA; 5–9% DHA) than in E. tricantha (12–18%
`EPA; 14–16% DHA). Ratios of EPAyDHA for
`1997 and 1998 E. superba were similar (1.4–1.6;
`Table 3, Fig. 2) whereas this ratio was somewhat
`lower for 1997 E. tricantha (1.2 vs. 0.9 for 1998).
`In contrast, the EPAyDHA ratio for E. frigida was
`lowest (0.7) in 1997 and highest (2.0) in 1998 of
`all samples analyzed.
`Arachidonic acid wAA, 20:4(ny6)x was 1% of
`total FA in 1997 and 1998 E. tricantha, and only
`0.4–0.6% in 1997 and 1998 E. frigida. In contrast,
`AA was not detected in E. superba and T. macrura.
`Although levels of the PUFA 18:4(ny3) were 4–
`9% in E. superba from both years, it was only
`
`RIMFROST EXHIBIT 1173 Page 0003
`
`

`

`736
`
`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`Table 1
`Percentage lipid class composition of 1998 Antarctic euphausiidsa
`
`E. superba (adult)
`E. superba (juv-large)
`E. superba (juv-small)
`E. tricantha
`E. frigida
`
`n
`
`4
`2
`2
`3
`3
`
`Wax ester
`
`Triacylglycerol
`
`Free fatty acid
`
`Sterol
`
`Polar lipid
`
`y1
`Lipid as mg g wet mass
`
`Lipidyindividual (mg g
`
`)y1
`
`–
`–
`–
`5.8"1.8
`–
`
`26.0"7.4
`38.4"3.4
`22.1"0.3
`16.2"7.4
`11.6"2.6
`
`1.3"0.3
`1.1"0.5
`1.8"0.7
`3.2"1.0
`1.8"0.3
`
`6.1"0.9
`4.1"0.8
`4.0"0.2
`7.4"1.1
`5.8"0.6
`
`66.6"6.3
`56.4"2.1
`72.0"0.6
`67.5"6.4
`80.8"3.3
`
`14.5"4.3
`14.6"4.3
`15.9"2.7
`12.4"3.5
`13.9"0.9
`
`7.2"2.1
`14.6"4.3
`8.0"1.3
`3.9"1.6
`2.1"0.2
`
`a
`
`Presented as mean"S.D.; –, below detection.
`
`RIMFROST EXHIBIT 1173 Page 0004
`
`

`

`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`737
`
`Table 2
`Percentage sterol composition of 1997 and 1998 Antarctic euphausiidsa
`
`Sterol
`
`24-Nordehydrocholesterol
`24-Nordehydrocholestanol
`Occelasterol
`trans-Dehydrocholesterol
`Cholesterol
`Cholestanol
`Desmosterol
`Brassicasterol
`Brassicastanol
`24-Methylenecholesterol
`Other
`
`1997
`
`E. superba
`(ns2)
`0.1"0.2
`–
`–
`1.1"0.4
`80.0"2.5
`0.1"0.1
`18.2"1.8
`0.5"0.0
`–
`0.1"0.1
`–
`
`E. tricantha
`
`–
`–
`–
`–
`94.3
`–
`5.7
`–
`–
`–
`–
`
`T. macrura
`(ns2)
`1.7"2.1
`tr
`0.1"0.1
`1.5"1.0
`81.2"15.3
`0.1"0.1
`6.5"1.3
`1.7"1.8
`0.2"0.2
`0.4"0.2
`6.7
`
`1998
`
`E. superba
`(adult, ns4)
`–
`–
`–
`–
`92.8"5.0
`–
`1.7"1.3
`–
`–
`–
`5.5
`
`a
`
`Presented as mean"S.D.; –, below detection; tr, trace (below integration).
`
`E. superba
`(juv-large, ns2)
`–
`–
`–
`–
`86.9"0.9
`–
`3.8"0.5
`–
`–
`–
`9.3
`
`E. superba
`(juv-small)
`
`–
`–
`–
`–
`88.6
`–
`2.9
`–
`–
`–
`8.5
`
`E. tricantha
`(ns3)
`–
`–
`–
`–
`100.0"0.0
`–
`–
`–
`–
`–
`–
`
`E. frigida
`(ns3)
`–
`–
`–
`–
`96.8"2.8
`–
`–
`–
`–
`–
`3.2
`
`RIMFROST EXHIBIT 1173 Page 0005
`
`

`

`738
`
`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`Table 3
`Percentage fatty acid composition of 1997 and 1998 Antarctic euphausiidsa
`
`Fatty acid
`
`1997
`
`1998
`
`E. tricantha
`
`E. frigida
`
`a
`
`?
`
`E. superba
`T. macrura
`E. superba
`E. superba
`E. superba
`E. tricantha
`E. frigida
`(ns4)
`(ns3)
`(adult, ns4)
`(juv-large, ns2)
`(juv-small, ns2)
`(ns3)
`(ns3)
`–
`–
`2.8
`–
`–
`–
`–
`–
`–
`12:0
`5.9"4.0
`15.3"43.8
`8.1"1.6
`8.6"0.5
`6.1"1.4
`5.1"1.2
`7.2"0.8
`2.3
`9.5
`14:0
`21.1"5.2
`22.9"0.7
`22.0"0.5
`22.4"1.1
`23.2"1.1
`17.6"0.7
`25.6"3.1
`25.1
`29.1
`16:0
`0.8"0.1
`4.0"5.9
`0.8"0.2
`0.8"0.0
`1.1"0.4
`0.8"0.1
`0.9"0.1
`1.5
`13.3
`18:0
`27.8"9.3
`42.2"11.4
`30.9"2.3
`31.8"1.6
`30.4"2.9
`23.5"2.0
`33.7"4.0
`28.9
`54.7
`Sum SFA
`16:1(ny7)c
`4.5"1.1
`3.1"0.3
`4.3"0.7
`3.5"1.3
`2.9"0.5
`5.6"0.9
`7.4"0.3
`2.3
`1.5
`16:1(ny5)c
`0.7"0.3
`1.0"0.3
`0.8"0.1
`0.5"0.2
`0.5"0.1
`0.6"0.1
`0.3"0.0
`0.2
`0.1
`18:1(ny9)c
`6.7"2.6
`9.6"0.5
`8.0"1.0
`7.7"1.6
`7.6"0.7
`18.9"1.7
`10.5"0.1
`11.9
`20.5
`18:1(ny7)c
`8.0"2.9
`6.4"1.4
`8.9"0.4
`7.7"0.8
`7.2"0.8
`9.6"0.7
`9.6"0.1
`13.0
`2.9
`20:1(ny9)c
`0.5"0.2
`0.8"0.4
`0.3"0.1
`0.1"0.0
`0.2"0.0
`2.2"0.6
`0.7"0.1
`1.0
`0.7
`20.4"7.1
`20.9"2.9
`22.3"2.3
`19.5"3.7
`18.4"2.1
`36.9"4.0
`28.5"0.6
`28.4
`25.7
`Sum MUFA
`C PUFA
`0.9"0.7
`0.2"0.3
`0.7"0.3
`1.5"0.8
`1.0"0.3
`0.3"0.0
`0.2
`0.1
`–
`16
`18:5(ny3)
`0.6"0.2
`1.2"0.5
`1.0"0.1
`0.2"0.1
`0.5"0.6
`–
`–
`–
`–
`18:4(ny3)
`5.2"4.6
`0.7"0.3
`3.7"1.2
`9.4"1.6
`8.9"0.5
`0.5"0.1
`0.6"0.1
`0.4
`–
`18:2(ny6)
`3.2"0.1
`2.1"0.8
`3.8"0.3
`2.4"0.3
`3.2"0.5
`2.1"0.1
`1.9"0.0
`1.7
`1.6
`18:3(ny6)
`1.3"0.7
`0.3"0.2
`1.7"0.4
`3.5"0.7
`3.7"0.3
`0.5"0.0
`0.8"0.0
`0.3
`0.3
`20:4(ny6)
`1.1"0.3
`0.6"0.1
`–
`1.2
`0.4
`–
`–
`–
`–
`20:5(ny3)
`20.5"5.7
`17.2"2.0
`17.1"1.5
`15.2"1.6
`15.0"1.1
`11.9"1.8
`17.6"2.9
`18.0
`3.6
`22:6(ny3)
`14.2"7.2
`10.2"2.9
`11.8"2.2
`9.3"0.5
`10.5"0.0
`13.5"2.5
`8.9"1.8
`15.6
`5.0
`45.3"19.0
`30.7"6.6
`39.4"6.1
`42.5"6.0
`43.3"2.8
`29.8"4.9
`31.2"5.5
`37.3
`11.0
`Sum PUFA
`6.5
`5.4
`8.6
`6.2
`7.4
`6.2
`7.9
`9.8
`6.6
`Other
`)y1
`11.7"5.1
`17.7"6.0
`9.2"2.5
`10.5"3.2
`12.7"0.6
`9.3"3.0
`9.6"1.1
`Total (mg g
`9.7
`7.0
`)y1
`2.2"0.6
`2.9"0.7
`1.5"0.3
`1.6"0.3
`1.9"0.1
`1.1"0.3
`1.7"0.5
`EPA (mg g
`1.7
`0.2
`)y1
`1.5"0.5
`1.7"0.2
`1.0"0.1
`1.0"0.2
`1.3"0.1
`1.2"0.2
`0.9"0.3
`DHA (mg g
`1.5
`0.3
`Ratio EPAyDHA
`1.44
`1.15
`0.72
`1.68
`1.44
`1.63
`1.43
`0.88
`1.99
`Presented as mean"S.D.; SFA, saturated fatty acids; MUFA, monosaturated fatty acids; PUFA, polysaturated fatty acids; EPA, eicosapentaenoic acid w20:5(ny3)x; DHA,
`docosahexaenoic acid w22:6(ny3)x; other components present at -1%: 12:0, 13:0, 15:0, 17:0, 20:0, 22:0, i14:0, i15:0, a15:0, i16:0, i17:0, a17:0, i19:0, 4,8,12-TMTD, 14:1(ny
`7)c, 14:1(ny5)c, 15:1, 16:1(ny7)t, 17:1, 18:1, 18:1(ny7)t, 18:1(ny5), 19:1, 20:1(ny11y13), 20:1(ny7)c, 22:1(ny11)c, 22:1(ny9)c, 22:1(ny7)c, C PUFA, 20:3(ny6),
`20:4(ny3), 20:2, 20:2(ny6), C PUFA, 22:4(ny6), 22:5(ny3), 22:2(ny6), C PUFA, C PUFA.
`
`21
`
`26
`
`28
`
`RIMFROST EXHIBIT 1173 Page 0006
`
`

`

`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`739
`
`the exception of 1997 E. frigida (0.024) and one
`1997 E. superba sample (0.032; Fig. 1). Separation
`between species was obtained with some overlap
`occurring.
`
`4. Discussion
`
`4.1. Lipid content and class composition
`
`Lipid content varied between krill species and
`between juvenile size classes in E. superba. Large
`y1
`juveniles of E. superba (14.6 mg g wet mass)
`showed the highest TAG (38.4% of total lipid;
`Table 1). Lipid content was 12.4–15.9 mg gy1
`wet mass in all other euphausiid species, which
`also had relatively less TAG (11.6–26.0%; Table
`1). TAG is a short-term energy reserve lipid (Lee
`et al., 1971), and a better fuel than PL. Higher
`TAG levels probably reflect greater feeding activ-
`ity in E. superba small juveniles. All krill (Table
`1) were collected in February, during the Antarctic
`summer, a time when phytoplankton production is
`at its maximum. The lipid level, lipid class and
`fatty acid composition of Thysanoessa species
`from Norwegian waters have been shown to
`change dramatically during this intense growth
`phase (Falk-Petersen, 1981). E. superba adults
`from the Weddell Sea have a large seasonal accu-
`mulation of lipid (average 28.2% lipid as% dry
`mass), whereas after winter, lipid values fall to
`10.5% (Hagen et al., 1996). Phosphatidylcholine
`appears to also be an energy storage lipid in E.
`superba (Hagen et al., 1996; Mayzaud, 1997).
`Polar lipid was high in all krill species from the
`AMLR study area (56–81% of total lipid; Table
`1). A similar PL level was found for Thysanoessa
`inermis (Falk-Petersen et al., 1982). Low lipid
`levels (high PL and low neutral
`lipid) usually
`indicate a low energy status. Individual phospho-
`lipids were not determined in this study. It has
`also been suggested that proteins contributed most
`energy during long term starvation in krill (Ikeda
`and Dixon, 1982). WE, a long-term energy reserve
`molecule were only detected in E. tricantha (6%
`of total lipid; Table 1). Moderate levels of WE
`(5–8%) were also reported form E. tricantha from
`the AMLR study area in 1996 (Phleger et al.,
`1998). Thysanoessa macrura, not analyzed in this
`study, has also been reported to have high WE
`(Hagen et al., 1996; Kattner et al., 1996; Falk-
`Petersen et al., 1999; Phleger and Nichols, unpub-
`lished data).
`
`RIMFROST EXHIBIT 1173 Page 0007
`
`Fig. 1. Dendrogram of cluster analysis comparing Antarctic
`euphausiids using fatty acids (performed using all fatty acids
`analysed), (individual samples; mgyg wet mass).
`
`present at lower levels (0–0.6%) in other krill
`species (Table 3; Fig. 3). The PUFA 18:5(ny3),
`not detected in 1997 samples, comprised 0.2–1.2%
`in all 1998 samples.
`The C and C very-long chain-polyunsaturat-
`26
`28
`ed fatty acids (VLC-PUFA) were identified by GC
`retention time data, with confirmation by GC–MS.
`The mass spectra of these components are similar
`to those of the other more common marine-derived
`PUFA with prominent ions at myz 79 and 91. The
`VLC-PUFA were not present in 1997 samples, but
`were detected as minor components (trace–0.1%)
`in 1998 samples of all krill except E. frigida.
`Monounsaturated fatty acids (MUFA) were 18–
`37% of the total FA (Table 3). MUFA included
`primarily 18:1(ny9)c (7–21%), 18:1(ny7)c (3–
`13%) and 16:1(ny7)c (2–7%) (Table 3). Highest
`levels of 18:1(ny9)c were found in 1997 E.
`frigida (21%) and 1998 E. tricantha (19%; Table
`3, Fig. 2). Although low levels (-1%) of
`20:1(ny9) and 20:1(ny7) were present in most
`krill species, 2.2% of 20:1(ny9) was detected in
`1998 E. tricantha (Table 3). The principal saturat-
`ed FA in all samples were 16:0 (18–29%), 14:0
`(2–15%) and 18:0 (1–13%) (Table 3).
`y1
`The fatty acid profiles (mg g
`wet mass) of
`individual samples were compared by cluster anal-
`ysis and MDS. All euphausiids were joined at a
`distance of F0.013 in the cluster analysis, with
`
`

`

`740
`
`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`Fig. 2. Ratio 18:1(n-7) to 18:1(n-9) versus ratio eicosapentaenoic acid (EPA) to docosahexaenoic acid (DHA) in Antarctic euphausiids (mean values). High 18:1(n-17)y18:1(n-
`9) ratios are consistent with an herbivorous diet. High EPAyDHA ratios reflect a diatom diet, while low ratios reflect a flagellate diet. (d), our data; (s), data from Virtue et al.
`(1993a,b), Virtue (1995).
`
`RIMFROST EXHIBIT 1173 Page 0008
`
`

`

`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`741
`
`Fig. 3. Percent 18:4 versus ratio 16:1 to 16:0 in Antarctic euphausiids (mean values). Low levels of 18:4 are consistent with an omnivorous diet. High 16:1y16:0 reflect a diatom
`diet. (d), our data (s), data from Virtue et al. (1993a,b), Virtue (1995).
`
`RIMFROST EXHIBIT 1173 Page 0009
`
`

`

`742
`
`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`4.2. Sterols
`
`The principle ST of all euphausiids analyzed
`was cholesterol (80–100%, of total ST) with lesser
`amounts of desmosterol (0–18%) (Table 2). Crus-
`taceans are incapable of de novo ST synthesis and
`depend on diet and phytosterol dealkylation (Goad,
`1978). Desmosterol is produced as an intermediate
`from phytosterol dealkylation and is also found in
`marine microalgae Nitzschia closterium (100%
`desmosterol) and Rhizoselenia setigera (94% des-
`mosterol) (Barrett et al., 1995). The waters where
`some of these krill were collected, known as
`Bransfield Strait waters, are dominated by micro-
`planktonic diatoms, such as Chaetoceros, Nitzschia
`and Rhizoselenia (Villafane et al., 1995). Desmos-

`terol present in krill may be from dealkylation of
`dietary phytosterols or, alternatively, directly sour-
`ced from the diet. Testing of the latter hypothesis
`will require examination of the ST content of
`phytoplankton from this region.
`The presence of ST other than cholesterol and
`desmosterol in 1997 E. superba and T. macrura
`(not in 1998 samples of these species) may indi-
`cate different food sources between these two
`years. 24-Nordehydrocholesterol and trans-dehy-
`drocholesterol (0.1–1.7 and 1.1–1.5%, respective-
`ly; Table 2) may be intermediates in cholesterol
`synthesis or derived from animal sources as well
`as phytoplankton. 24-Nordehydrocholesterol was
`3–12% of the total ST, and trans-dehydrocholes-
`terol 2–24% in a number of Cnidaria and Cteno-
`phora species collected from the AMLR study area
`during 1997 and 1998 (Nelson et al., 2000).
`Similar
`levels of 24-nordehycholesterol were
`observed in 1996 Cnidaria from the same study
`area (Phleger et al., 1998). trans-Dehydrocholes-
`terol was the major ST in the Antarctic ice diatom
`Nitzschia cylindrus (Nichols et al., 1986), and the
`structurally
`similar
`27-nor-24-methylcholest-
`5,22E-3b-ol (occelasterol) has also been reported
`as the major ST in an Antarctic dinoflagellate
`(Thompson, unpublished data). Care is needed, by
`examination of other signature lipids and by other
`procedures, before the exact algal source of these
`minor ST present in krill can be ascertained. The
`low levels of brassicasterol observed in 1997 E.
`superba and T. macrura also may have originated
`from dietary prymnesiophytes, such as Phaeocys-
`tis, and diatoms, where it is a major ST (Nichols
`et al., 1991; Tsitsa-Tzardis et al., 1993). The low
`24-methylenecholesterol (0.1–0.4%; Table 2) may
`
`have originated from sources such as the diatom
`is the major ST (Tsitsa-
`Chaetoceros, where it
`Tzardis et al., 1993). High 24-methylenecholester-
`ol was detected in the solitary salp, Salpa
`thompsoni (52% of total ST) and 5–12% 24-
`methylenecholesterol was reported in aggregate
`salps of the same species (Phleger et al., 2000).
`This indicates different
`feeding preferences or
`sources for these salps as compared to the krill
`reported in this study.
`Low levels of stanols were only detected in
`1997 E. superba and T. macrura but not in 1998
`species (Table 2). A similar difference
`krill
`between these years was also observed in salps (S.
`thompsoni) and their commensal hyperiid amphi-
`pods (Vibilia antarctica, Cyllopus lucasii and C.
`magellanicus) from the AMLR study area. Four
`stanols were detected in 1997 salps, and in their
`commensal
`amphipods;
`the
`stanols
`neither
`occurred in the 1998 salps nor in their amphipods
`(Phleger et al., 2000). The presence of stanols
`reflects diet. For example, in the marine dinoflag-
`ellate Scrippsiella sp., cholestanol comprised 24%
`of total stanols (Mansour et al., 1999) and a
`species of Gymnodinium isolated from Australian
`waters contained 24% dinostanol (Nichols et al.,
`1984). Dinostanol was not detected in the krill
`analyzed in this study. The fact that stanols were
`only detected in some 1997 krill species is proba-
`bly indicative of changes in the importance of the
`dietary sources of these components or products
`of bacterial degradation.
`
`4.3. Fatty acids
`
`Separation between species was represented in
`cluster analysis (Fig. 1). A scatterplot of MDS
`y1
`using FA profiles (mg g wet mass) of individual
`samples showed similar separation, although the
`1998 E. superba grouped tighter than 1997 E.
`superba (data not shown). The separations used
`the full suite of FA analyzed; similar separation
`also occurs when the FA suite is reduced (e.g.
`selection of components )1% of total FA). All E.
`superba were closely joined in the cluster analysis,
`except one 1997 E. superba sample (Fig. 1). Since
`1998 was considered as a ‘salp year’ and 1997
`was a ‘transition year’ (Phleger et al., 2000), 1997
`E. superba may receive a more diverse diet, which
`is reflected in the more varied FA profile. Although
`the composition of zooplankton for these years is
`known (Loeb et al., 1998), the composition of the
`
`RIMFROST EXHIBIT 1173 Page 0010
`
`

`

`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`743
`
`diet of E. superba (mostly phytoplankton) was not
`determined during the AMLR field studies and is
`therefore not known.
`For 1997 T. macrura, cluster analysis joined the
`animals at a distance of 0.009 (Fig. 1). Addition-
`ally, the small and large 1998 E. superba juveniles
`group together separated from the adults. 1998 E.
`tricantha and E. frigida are distinct and separate,
`however, 1997 samples do not cluster near their
`1998 counterparts (Fig. 1). These differences may
`also reflect variations in diet, as a separation was
`also evident for plots of selected biomarkers (Figs.
`2 and 3, and see below).
`Our data for the PUFA eicosapentaenoic acid
`wEPA, 20:5(ny3)x and docosahexaenoic acid
`wDHA, 22:6(ny3)x (Table 3) in E. superba from
`the Elephant Island AMLR study area are similar
`to those reported by Virtue et al. (1993a, 1997)
`for E. superba from the Prydz Bay region of
`Antarctica. These long chain PUFA, EPA and
`DHA, are dominant in the phospholipid fraction
`of krill where they appear to function in membrane
`structure (Virtue et al., 1993b). Higher proportions
`of EPA were present in phosphatidylcholine of E.
`superba compared with phosphatidyethanolamine
`(Mayzaud, 1997). A dietary source of these PUFA
`is considered essential in krill and other crustacean
`species, and phytoplankton, especially diatoms and
`other phaeophytes, are rich in PUFA which they
`synthesize (Brown et al., 1989). Higher EPAy
`DHA ratios in krill may be due to metabolic rate
`variations of these two PUFA. However, they more
`probably reflect differences in the diet. For exam-
`ple, diatoms contain high EPA (Falk-Petersen et
`al., 1998), whilst dinoflagellates and other flagel-
`lates have elevated DHA (Falk-Petersen et al.,
`2000). E. superba had EPAyDHA ratios of 1.4–
`1.6 for 1997 and 1998 adults and juveniles (Fig.
`2; Table 3). Major diatom FA include EPA,
`16:1(ny7) and C PUFA (Volkman et al., 1989;
`16
`Dunstan et al., 1994). The ratios of EPAyDHA
`were less in 1997 E. frigida (0.7) and 1998 E.
`tricantha (0.9) than in E. superba (Fig. 2; Table
`3). These may reflect a different dietary input
`since E.
`frigida is carnivorous, E.
`tricantha is
`omnivorous, and E. superba is herbivorous in
`earlier life stages and omnivorous as an adult.
`Higher levels of oleic acid w18:1(ny9)cx were
`frigida (21%) and 1998 E.
`found in 1997 E.
`tricantha (19%) with lower levels in E. superba
`w7–8% of 18:1(ny9)cx. Higher levels of oleic
`acid are also more consistent with a carnivorous
`
`diet (Nelson et al., 2000). Supporting evidence is
`provided by the ratios of 18:1(ny7)cy18:1(ny
`9)c which were mostly lower (0.1–1.1 for 1997
`and 0.5–0.9 for 1998; Fig. 2; Table 3) for these
`species, whereas the ratios for E. superba of
`18:1(ny7)cy18:1(ny9)c were higher (0.9–1.2
`for both years). In addition, relatively low levels
`of 20:1(ny9)c (-1% of total FA) were detected
`in all krill species except 1998 E. tricantha (2.2%)
`and -1% 20:1(ny7)c (Table 3). These FA are
`found in high levels in copepods and other crus-
`taceans (Kattner et al., 1994). Levels of 4–10%
`of 20:1(ny9)c and 20:1(ny7)c were found in
`1996 E. tricantha from the same AMLR study
`area (Phleger et al., 1998). E.
`tricantha is a
`mesobathypelagic euphausiid which is predomi-
`nantly carnivorous and the carnivory is reflected
`in these higher 20:1(ny9)c values (Table 3).
`Lower 16:1(ny7)c values (2–3%) were noted
`in 1997 E. tricantha, E. frigida and T. macrura,
`which indicates less importance of diatoms in the
`diet (Cripps et al., 1999). The 16:1(ny7)c values
`for E. superba were somewhat higher (5% for
`1997, and 3–4% 1998) as were 1998 values for
`the other krill species (6–7%; Table 3). 16:1y16:0
`ratios for all krill were in the range 0.05–0.57
`(Fig. 3; Table 3). Higher values have been pro-
`posed to indicate a diatom-containing diet (Volk-
`man et al., 1989). In E. superba, conversion of
`16:1(ny7)c to 18:1(ny7) is occurring.
`The PUFA 18:4(ny3), which was 4–9% of
`total FA in E. superba from 1997 and 1998 (Table
`3), is a major FA in the prymnesiophyte Isochrysis
`sp. (T-ISO) and in the cryptomonad Chroomonas
`salina (Volkman et al., 1989). The increased levels
`of 18:4(ny3) in particulate matter samples and
`the copepod Paralabidocera antarctica from coast-
`al waters, at Davis Station in Eastern Antarctica,
`was suggested to be indicative of Cryptomonas
`spp. (Swadling et al., 2000). In composition, the
`proportions of the various PUFA and other FA in
`the krill are suggestive of a phytoplankton diet,
`specifically diatom and cryptomonads for E. super-
`ba. The lower levels of 18:4(ny3) (0–0.6%) in
`other krill species, including E. tricantha, E. fri-
`gida and T. macrura, is more consistent with a
`carnivorous or omnivorous diet (Fig. 3). Another
`source of 18:4(ny3) could be from 18:5(ny3),
`which is rapidly metabolized to 18:4(ny3) in
`cultured fish cells (Ghioni et al., 2001).
`The PUFA octadecapentaenoic acid w18:5(ny
`3)x, present at 0.2–1.2% of total FA in all 1998
`
`RIMFROST EXHIBIT 1173 Page 0011
`
`

`

`744
`
`C.F. Phleger et al. / Comparative Biochemistry and Physiology Part B 131 (2002) 733–747
`
`krill, was not detected in any 1997 samples (Table
`3). This unusual PUFA was first recognized as an
`important biomarker in marine zooplankton by
`Mayzaud et al. (1976). It was also only detected
`in 1998 AMLR Antarctic salps and their commen-
`sal amphipods, AMLR gelatinous zooplankton,
`and AMLR free amphipods, but neither in 1997
`nor 1996 zooplankton (Phleger et al., 1998, 2000;
`Nelson et al., 2000, 2001). 18:5(ny3) is a major
`FA in dinoflagellates (Nichols et al., 1984) and
`coccolithophorids (Volkman et al., 1981), and is
`synt

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket