throbber
LWT - Food Science and Technology 126 (2020) 109284
`
`Contents lists available at ScienceDirect
`LWT - Food Science and Technology
`
`journal homepage: www.elsevier.com/locate/lwt
`
`Separation and concentration of ω-3 PUFA-rich phospholipids by hydration
`of krill oil☆
`Casey Showman, Kimberly Barnes, Jacek Jaczynski∗∗, Kristen E. Matak∗
`Division of Animal and Nutritional Sciences, West Virginia University, PO Box 6108, Morgantown, WV, 26505-6108, USA
`
`T
`
`A R T I C L E I N F O
`
`A B S T R A C T
`
`Keywords:
`Krill oil
`Omega-3 fatty acids
`Phospholipids
`Water degumming
`Nutraceuticals
`
`Krill oil (KO) is unique in contrast to fish oil because the ω-3 PUFAs, particularly eicosapentaenoic (EPA) and
`docosahexaenoic (DHA), are almost exclusively esterified on phospholipids (PL) rather than triglycerides (TAG)
`which improves bioavailability and water solubility. Therefore, the primary aim of this study was to use the
`principles of water degumming to separate PL from TAG in KO. Water was mixed with KO in varying ratios,
`centrifuged for separation and freeze-dried. Separation into gum and oil fractions occurred in the 75:25 and
`50:50 KO:H2O samples. TLC-densitometry revealed 67.6 ± 1.97% and 49.73 ± 3.90% PL concentrations
`(p < 0.05) in the lipid recovered from the 75:25 and 50:50 KO:H2O sample, respectively. The 75:25 KO:H2O
`gum fraction did not contain detectable triglycerides (TAG); however, the 50:50 KO:H2O gum fraction had
`23.88 ± 6.36% (p < 0.05). The fatty acid profile of the 75:25 KO:H2O gum fractions presented with more EPA
`and DHA than the oil fraction (p < 0.05). Phosphatidylcholine (PC) with EPA and DHA fatty acids attached
`concurrently at sn-1 and sn-2 (i.e., PC-EPA/EPA, PC-EPA/DHA, and PC-DHA/DHA) and was present in the gum
`fractions of both samples. This approach offers a simple separation technique to separate health-beneficial PL
`from KO.
`
`1. Introduction
`Cardiovascular disease (CVD) is the leading cause of death in the
`United States (Bunea, El Farrah, & Deutsch, 2004). Omega-3 poly-
`unsaturated fatty acids (ω-3 PUFAs) have demonstrated health benefits
`that may protect against it (Bunea et al., 2004). Specific ω-3 PUFAs of
`importance include eicosapentaenoic acid (EPA, 20:5 ω-3) and doc-
`osahexaenoic acid (DHA, 22:6 ω-3). Typical sources of EPA and DHA
`are fatty fish; however, krill also contains significant amounts of these
`PUFAs.
`Antarctic krill (Euphausia superba) are small crustaceans with one of
`the largest biomasses of any multicellular animal species (Nicol, James,
`& Pitcher, 1987). The lipid content, on a dry matter basis, ranges from
`12 to 50% based on maturity, season and location of harvest
`(Kolakowska, 1991; Saether, Ellingsen, & Mohr, 1986). Krill are a un-
`ique species due to the composition of lipid in their bodies. In contrast
`to fish oil, the ω-3 PUFAs, particularly EPA and DHA, in krill oil are
`almost exclusively esterified on phospholipids (PL) rather than trigly-
`cerides (TAG) (Kutzner et al., 2016; Tou, Jaczynski, & Chen, 2007;
`
`Winther, Hoem, Berge, & Reubsaet, 2010). Interestingly, some of these
`ω-3 PUFAs in KO appear uniquely esterified on a single PL molecule
`concurrently occupying sn-1 and sn-2 positions.
`PLs are most often found within cell membranes. Their structure
`consists of a polar, hydrophilic phosphate head and two hydrophobic
`fatty acid (FA) chains. The polarity and hydrophobic interactions allow
`for the formation of lipid bilayers (Gilbert, 2000). The level of satura-
`tion of the FA chain varies. Unsaturated FA allow the bilayer to be more
`fluid in colder temperatures. This is due to cis-double bonds imposing
`stress on the FA chain, which results in a non-linear shape of the FA.
`These molecules increase their stearic hindrance, which inhibits the
`chains from lining up closely (Mathews, van Holde, & Ahern, 2000).
`More saturated FA chains would cause the lipid bilayer to be more rigid
`in the colder temperatures of the krill's natural environment.
`PLs have several functions in the pharmaceutical and food produc-
`tion industries due to their amphiphilic character. PLs are commonly
`called lecithin. One of the most common uses of lecithin in the food
`industry is for stabilizing oil/water emulsions (Ceci, Constenla, &
`Crapiste, 2008). In the pharmaceutical industry, PLs serve as drug
`
`☆ This work was supported by the USDA Hatch program (WVA00722).
`∗ Corresponding author. Division of Animal and Nutritional Sciences, West Virginia University, PO Box 6108, Morgantown, WV, 26505-6108, USA.
`∗∗ Corresponding author. Division of Animal and Nutritional Sciences, West Virginia University, PO Box 6108, Morgantown, WV, 26505-6108, USA
`E-mail addresses: clshowman@mix.wvu.edu (C. Showman), kim.barnes@mail.wvu.edu (K. Barnes), jacek.jaczynski@mail.wvu.edu (J. Jaczynski),
`kristen.matak@mail.wvu.edu (K.E. Matak).
`https://doi.org/10.1016/j.lwt.2020.109284
`Received 19 November 2019; Received in revised form 9 March 2020; Accepted 13 March 2020
`Available online 15 March 2020
`0023-6438/ © 2020 Elsevier Ltd. All rights reserved.
`
`RIMFROST EXHIBIT 1174 Page 0001
`
`

`

`C. Showman, et al.
`
`delivery systems because of their unique biocompatibility and hydro-
`philic properties (Li et al., 2015). Dietary PLs have a high bioavail-
`ability and can have a protective effect on diseases such as coronary
`heart disease (Küllenberg, Taylor, Schneider, & Massing, 2012).
`There are several species of PL present in krill oils which includes
`phosphatidylcholine (PC), phosphatidylinositol
`(PI), phosphatidy-
`lethanolamine (PE), phosphatidylserine (PS), and phosphatidic acid
`(PA). The PL species are separated based on the chemical structure
`attached to the phosphate head. This structure also makes them either
`hydratable or nonhydratable. Nonhydratable PL are commonly sepa-
`rated from oil and neutral lipids through processes such as acid de-
`gumming, whereas hydratable PLs are separated with the addition of
`water (Dijkstra & Segers, 2007). The predominant PL species in krill is
`PC and interestingly, Castro-Gomez, Holgado, Rodriguez-Alacala,
`Montero, and Fontecha (2015) reported that 6.5% of the total PC in KO
`contained EPA/DHA and EPA/EPA concurrently on the sn-1 and sn-2
`positions of one PC molecule. PC is fully hydratable due to its chemical
`structure, swelling capacity, and ordered structure formation (Diez,
`1995; Larsson, 1994). The head group on a PC molecule carries a large
`dipole moment due to the negative charge on the phosphate and po-
`sitive charge on the trimethylamine group causing PC to strongly favor
`water (Dijkstra, 2018).
`Water degumming is an industrial process to separate PLs in crude
`oil and produce commercial lecithin. In the degumming process, the
`amount of water added is typically equal to that of the amount of gum
`to be separated (Braae, 1976). The process includes mixing water with
`crude oil to hydrate the polar lipids causing the oil and gum to separate.
`The resulting products are a refined oil and gum that consists of TAGs
`and PLs, respectively. Crude KO is typically extracted from krill with
`conventional organic and inorganic solvents, supercritical fluids such as
`CO2, as well as solvent-free methods (Bruheim, Tilseth, & Mancinelli,
`2018; Gigliotti, Davenport, Beamer, Tou, & Jaczynski, 2011; Katevas &
`Guerra, 2018; Sampalis & Harland, 2016; Soerensen & Jensen, 2014).
`However, these methods yield crude KO whose composition is variable
`and comprises variety of lipid classes without targeting purification/
`concentration of health-beneficial krill PLs; while some of
`these
`methods are inherently costly.
`To our best knowledge, degumming KO has not been tested for the
`separation of ω-3 PUFA-rich PLs. Therefore, the primary aim of this
`study was to use the principles of water degumming to separate phos-
`pholipids from triglycerides in krill oil by hydration. It is hypothesized
`that water degumming will be an adequate, simple method to separate
`PL-esterified ω-3 PUFAs. The secondary objective of this study was to
`verify the presence and attempt quantification of unique krill PC whose
`sn-1 and sn-2 are concurrently occupied by various permutations of
`EPA/DHA (i.e., PC-EPA/EPA, PC-EPA/DHA, and PC-DHA/DHA).
`2. Materials and methods
`2.1. Sample
`Virgin krill oil (KO) from Antarctic krill (Euphausia superba) was
`purchased from Jedwards International, Inc. (Braintree, MA, U.S.A.).
`Upon arrival, KO was stored in an amber glass bottle, flushed with ni-
`trogen, and stored at 2–5 °C throughout the experiment.
`2.2. Hydration fractionation
`KO was fractionated using deionized distilled water (ddH2O).
`Various ratios of KO and ddH2O were used. Hydration trials included
`100% KO, 75:25 KO:H2O, 50:50 KO:H2O, and 25:75 KO:H2O wt/wt.
`Krill oil and ddH2O were weighed to a final weight of 50 g into glass
`beakers, flooded with nitrogen, and covered with foil. Each ratio was
`stirred at 200 rpm for 24 h at room temperature (~20 °C). After 24 h,
`each sample was placed into a centrifuge tube. Water soluble and in-
`soluble fractions were separated by centrifuging at 5000 g for 5 min at
`
`LWT - Food Science and Technology 126 (2020) 109284
`
`22 °C. Centrifugation generally resulted in two visual fractions, gum
`and oil, that were manually decanted. Special attention was paid to
`prevent fraction contamination during decanting. Separated gum and
`oil fractions were freeze dried (VirTis, Model #35L) to remove H2O
`followed by storage at −80 °C. Each of the four ratios were completed
`in triplicates. Weights were recorded after stirring, separation of each
`fraction (i.e., gum and oil after centrifugation and decanting), and
`freeze-drying.
`2.3. Thin layer chromatography and densitometry
`Thin Layer Chromatography (TLC) was applied to separate lipid
`classes using a method described by Gigliotti et al. (2011). Fractions
`separated as described above were diluted to 5 mg/mL in a 1:1
`chloroform: methanol solvent (vol:vol). Twenty μL of the mixture was
`pipetted onto a silica TLC plate (Merck TLC Silicagel 60 plates with
`60 Å pore size, Darmstadt, Germany). For the mobile phase, the TLC
`plate was developed in a hexane: diethyl ether: acetic acid (80:20:1.5;
`vol:vol:vol) for approximately 1 h until the solution reached 2.5 cm
`from the top of the plate. After drying for 5 min, TLC plates were
`sprayed with 50% sulfuric acid in water solution (1:1 vol:vol) and
`placed in a 120 °C drying oven for 45 min to develop spots indicating
`lipid classes. Standards for mono-, di-, and triglycerides, phospholipids,
`cholesterol (Supelco, Bellefonte, PA), and free fatty acids (oleic acid, Nu
`Chek Prep, Elysian, MN) were also plated and run for identification of
`each lipid class. TLC was applied to all fractions from each of the four
`KO:H2O ratio tested.
`For densitometry analysis, images of the TLC plates were captured
`with Gel Doc XR + gel imaging system (Bio-Rad Gel Doc XR+ #
`170–8170 and ChemiDoc XRS + Imaging Software Version 6,
`California, United States) using transluminating white light and pixel
`density was measured. Densitometry values are expressed as percent
`lipid class (mean ± SD) of the total bands detected within the TLC
`lane.
`2.4. Fatty acid profile analysis
`2.4.1. Extraction
`The Bligh and Dyer (1959) method using chloroform-methanol
`(C:M) mixture (2:1 v/v) was followed to extract total lipids from each
`fraction (i.e., gum and oil). A 0.05 g sample was used from each fraction
`for analysis. Trizma/EDTA buffer (50 mM, pH = 7.4) and a chloroform:
`methanol: glacial acetic acid (C:M:A; 400:200:3 mL) solvent were
`added. Nonadecanoic acid (19:0) was added (125 μL) as an internal
`standard for quantification of fatty acids. Samples were incubated at
`room temperature for 10 min, then centrifuged at 900 g for 10 min at
`10 °C. The separated lower layer was filtered through a pre-rinsed
`Whatman 1-PS filter (Whatman 2200090, Buckinghamshire HP7, 9NA,
`UK). The upper layer was run through extraction methods using a 4:1
`chloroform: methanol solvent, centrifuged to separate, and again the
`lower layer was filtered through the same filter. Filtrates were flushed
`with nitrogen gas at 60 °C to dry.
`2.4.2. Methylation
`The extracted lipid was transmethylated according to Fritshe and
`Johnston (1990). Methylation occurred by adding 4 mL of 4% H2SO4, in
`anhydrous methanol, to the lipid and incubating in a water bath at
`90 °C for 60 min. Deionized distilled water was added to stop the re-
`action after incubation period. Chloroform was added to extract fatty
`acid methyl esters (FAMEs), and mixture was centrifuged at 900 g for
`10 min at 10 °C. The separated lower layer was filtered through an-
`hydrous Na2SO4. The collected layer was dried with nitrogen gas at
`60 °C, diluted with isooctane, and stored at −20 °C until analyzed.
`2.4.3. Fatty acid profile and quantification
`Fatty acid methyl esters were analyzed using a Varian CP-3800 Gas
`2 RIMFROST EXHIBIT 1174 Page 0002
`
`

`

`C. Showman, et al.
`
`Instruments; Walnut Creek,
`Chromatograph (Varian Analytical
`California, U.S.A) equipped with a flame ionization detector (FID;
`Varian Inc., Walnut Creek, California, U.S.A.). A silica capillary column
`(100 m length, 0.25 mm diameter) was used to separate the FAMEs. A
`method of 140 °C held for 5 min followed by a temperature ramp of
`4 °C/min to 220 °C held for 15 min was adopted; totaling to 85 min for
`each FAME separation. Temperatures were held at 270 °C and 300 °C
`for the injector and detector respectively. Identification of FAMEs was
`based on retention times compared to FAME 37 standard (SupelcoTM
`quantitative standard FAME 37; Sigma-Aldrich, St. Louis, Missouri,
`U.S.A.). The Star GC workstation version 6 software (Varian Inc.,
`Walnut Creek, California, U.S.A.) was used to determine the peak area
`and relative amounts of each fatty acid in the samples. Fatty acids are
`expressed as mg FA/g sample based on quantified amounts (mg/g).
`
`2.5. Ultra high-performance liquid chromatography mass spectrometry
`2.5.1. Sample preparation
`Quantification of phosphatidylcholine (PC) in the gum fractions was
`accomplished by reverse phase ultra high-performance liquid chroma-
`tography mass spectrometry (UHPLC-MS). The methodology is similar
`to that developed by Winther et al. (2011). Briefly, each sample and a
`stock solution of PC-free matrix were dissolved in 60:40 (vol:vol) di-
`chloromethane:methanol mixture to a concentration of 1 mg/mL.
`Phosphatidylcholine
`22:6/22:6
`standard
`(Avanti Polar
`Lipids,
`Alabaster, AL) was used as an external calibrator. The standard was
`mixed into PC free matrix solution in a set of dilutions with con-
`centrations of 1, 5, 10, 50, and 100 μg/mL. A calibration curve of
`chromatographic peak area to standard concentration was established
`from set of dilutions to quantify krill oil samples.
`
`2.5.2. HPLC setup and conditions
`Chromatographic separation was performed in reverse phase on a
`Thermo Accela UHPLC with an Agilent Eclipse XDB-C18 column with
`particle diameter of 5 μm. Column dimensions were 2.1 × 100 mm.
`There were two mobile phases used to elucidate the PC. Phase A con-
`sisted of 90% water with 10% formic acid while phase B consisted of
`60% methanol, 40% acetonitrile, and 0.1% formic acid. A linear gra-
`dient of the two phases, similar to that of Winther et al. (2011), was
`used for separation of PC species. Briefly, for 4 min after injection the
`system was kept constant at 65% mobile phase B and then increased to
`100% mobile phase B for 22 min. The mobile phase flow rate was set to
`0.26 mL/min.
`
`2.5.3. Mass spectrometry
`Detection of PC species was obtained using a Thermo Q-Exactive
`with method setup, data acquisition, and quantification done by soft-
`ware Xcalibur 3.0 (Thermo Scientific, USA). The mass spectrometer ran
`at a voltage of 4.00 kV, sheath gas flow rate of 30 (arbitrary units),
`auxiliary gas flow rate of 5, and sweep gas of 5. Capillary temperature
`was set to 200 °C. Mass spectrometry resolution was set at 70,000 (m/z
`200) for full mass scan and 17,500 for tandem mass. Mass scan range
`was set to 100–1000 m/z in profile mode. Parallel reaction monitoring
`analysis was used. Both the precursor ions and the fragment ions were
`selected to monitor the transition using the ratio of respective peak
`areas to quantify each PC species. Masses for each PC species, de-
`termined by previous studies (Castro-Gomez et al., 2015), were used to
`identify the molecules by mass spectrometry.
`The precursor m/z window was set as ± 1. A full scan of precursor
`ions was carried out. Then the selected precursor ions were subjected to
`high-energy collision dissociation (HCD), followed by the selection of
`specific product ions in the Orbitrap Mass Analyzer (Thermo Scientific,
`USA).
`
`LWT - Food Science and Technology 126 (2020) 109284
`
`2.6. Statistical analysis
`Hydration fractionation trials were performed in triplicates (n = 3).
`At least duplicate TLC with densitometry and fatty acid profile mea-
`surements were performed for each triplicate fractionation trial.
`UHPLC-MS was conducted once for each triplicate. SAS JMP Pro ver-
`sion 13 (SAS Institute Inc., North Carolina, USA) was used for all ana-
`lyses. One-way analysis of variance (ANOVA) with post-hoc analysis
`using Tukey-Kramer's test (p < 0.05) was used to determine statistical
`differences.
`3. Results and discussion
`3.1. Fractionation of krill oil (KO) by hydration
`Fractionation into water soluble and insoluble fractions, or “gum”
`and “oil” fractions, respectively, occurred only in the 75:25 KO:H2O
`and 50:50 KO:H2O ratios. The 25:75 KO:H2O ratio formed an oil/water
`dispersion and since no fractionation occurred after centrifugation,
`further results are not reported. When a 50:50 KO:H2O ratio was used to
`separate lipids, the starting weight of krill oil was 25 g. After separa-
`tion, the weight of the oil fraction was 2.50 ± 1.36 g and the weight of
`the gum fraction was 20.53 ± 1.16 g. On the other hand, when a 75:25
`KO:H2O ratio was used the starting weight of the krill oil was 37.5 g.
`After separation, the weight of the oil fraction was 11.78 ± 1.42 g and
`the weight of the gum fraction was 19.83 ± 2.28 g. There was a minor
`but inevitable loss of sample due to the transfer from container to
`container in each step of the process because of the viscosity of KO.
`There were obvious visual differences in gum fractions obtained
`with the different KO:H2O ratios. These differences were not measured,
`but the 50:50 KO:H2O gum fraction had a similar appearance to the
`emulsion created by the 25:75 KO:H2O ratio, while the 75:25 KO:H2O
`gum fraction had a dark red and sludge-like appearance. This difference
`was likely due to the entrapment of oil in structures formed in the
`hydration of the 50:50 KO:H2O sample rather than the precipitation of
`polar lipids.
`Water is a polar solvent that is immiscible in oil; therefore, when
`water interacts with PLs they also become insoluble in the oil when
`hydrated. When there is an optimal ratio of KO to water the PLs form
`aggregates that are oil insoluble and can be separated from the oil (Lei,
`Ma, Kodali, Liang, & Davis, 2003). Lei et al. (2003) demonstrated a
`ternary phase diagram of soybean oil-water-soybean PC to assist in
`determination the appropriate amount of water to be added during
`degumming. They also showed that when an excess of water is added,
`above the PL saturation point, water droplets penetrate liposomes and
`disperse in the oil. This is likely why the 25:75 KO:H2O ratio had no
`separation, as well as the 50:50 KO:H2O ratio was approaching the
`saturation point of the PLs found in KO.
`3.2. Lipid class profiles
`Based on thin layer chromatography (TLC) and densitometry, water
`fractionation of PLs from neutral lipids, such as TAGs, was possible. TLC
`displayed how each major lipid class in KO fractionated after the ad-
`dition of water (Fig. 1). Densitometry gave the relative distribution of
`lipid classes within each fraction. Data are reported in Table 1 and are
`expressed as % lipid class ± standard deviation in that particular
`fraction. The 100% KO contained 19.78 ± 3.43% TAG and
`46.28 ± 0.32% PLs. The relative distribution of PLs is consistent with
`reported ranges between 46 and 64% (Kolakowska, 1991; Castro-
`Gomez et al., 2015, and Xie et al., 2017).
`Undetectable amounts of PLs remained in the 50:50 and 75:25
`KO:H2O oil fraction,
`likely because the PLs in KO are comprised
`of > 90% phosphatidylcholine (PC) which is fully able to absorb water
`(Table 1; Akanbi & Barrow, 2018; Castro-Gomez et al., 2015; Dijkstra &
`Segers, 2007). In addition, the 75:25 KO:H2O ratio gave rise to better
`3 RIMFROST EXHIBIT 1174 Page 0003
`
`

`

`C. Showman, et al.
`
`LWT - Food Science and Technology 126 (2020) 109284
`
`Fig. 1. Thin layer chromatography (TLC) displaying lipid classes for 100% krill oil (KO) and the 25:75 KO:H2O, 50:50 KO:H2O, and 75:25 KO:H2O and fractions
`separated by water degumming principles as well as standards used for identification. Along the x-axis, the lanes represent: MG DAG TAG = mono-, di-, triglyceride
`standard; FFA = free fatty acid (oleic acid) standard, cholesterol standard, PL = phospholipid standard, 100% KO, 25:75 KO:H2O (fractionation for oil and gum did
`not occur), 50:50 KO:H2O oil fraction, 50:50 KO:H2O gum fraction, 75:25 KO:H2O oil fraction, 75:25 KO:H2O gum fraction. Along the y-axis, lipid classes are
`abbreviated as TAG (triglyceride), FFA (free fatty acid), CHOL (cholesterol), DAG (diglycerides), and PL (phospholipids).
`
`separation of lipid classes with more concentrated PL (approximately
`68%) and undetectable amounts of TAG in the gum fraction; whereas
`the 50:50 KO:H2O gum fraction had a less concentrated PL content
`(approximately 50%) due to the presence of about 24% TAGs. The oil
`fractions of both 75:25 KO:H2O and 50:50 KO:H2O ratios contained
`mainly TAGs (> 80%).
`In addition, KO typically contains high and varying levels of free
`fatty acids (FFAs). Since krill are crustaceans, they also contain some
`levels of cholesterol (CHOL). Relatively high levels of FFAs were pre-
`sent
`in both 50:50 (14.51 ± 3.28%) and 75:25 KO:H2O
`(15.41 ± 1.52%) gum fractions. Unrefined oils typically contain some
`levels of FFAs. In this regard, unrefined KO usually contains much
`higher and more variable amounts of FFAs than oils from other sources.
`In addition to causing overall quality deterioration, FFAs are also
`considered impurities because they may promote oil hydrolytic de-
`gradation. FFAs may be removed by a neutralization step during oil
`refining. If neutralization is applied to KO, FFAs can be removed; and
`therefore, krill PL-esterified EPA/DHA can be further purified/con-
`centrated. If one assumes data presented in Table 1 for the 75:25
`
`KO:H2O gum fraction, removal of 15.41 ± 1.52% FFAs by neu-
`tralization could result in approximately 80% of PLs in that fraction
`when compared to 68% without neutralization. This increase in PLs
`would result is a proportional increase of EPA/DHA as well. Although
`the content of FFAs in unrefined KO is typically high as well as highly
`variable, neutralization step of KO would be beneficial in terms of
`higher content of PL-esterified EPA/DHA. Oil neutralization is rela-
`tively simple and a follow up study for neutralized KO is forthcoming.
`
`3.3. Fatty acid profile
`Major fatty acids (FAs) with content greater than 1 mg/g found in
`the 100% KO, 75:25 KO:H2O, and 50:50 KO:H2O ratios and their re-
`spective fractions (gum/oil) are listed in Table 2. These FAs include
`C14:0, C16:0, C16:1, C18:1, C18:3 (ALA), C20:2, C20:5 (EPA), and
`C22:6 (DHA). Results are presented as mg FA in g FAs ± standard
`deviation. There were significant differences in FA profiles for each oil
`and gum fraction (p < 0.05).
`For both the 50:50 KO:H2O and 75:25 KO:H2O samples, the gum
`
`Table 1
`Relative distribution of lipid classes in respective fractions (% lipid class/sample or fraction) of 100% krill oil (KO) and the 50:50 KO:H2O and 75:25 KO:H2O
`fractions separated by water degumming principles, determined by TLC-densitometry.
`Lipid Class
`CHOL
`FFA
`TAG
`19.78 ± 3.43B
`100% Krill Oil
`14.09 ± 0.20AB
`14.19 ± 4.84A
`50:50 KO:H2O
`82.38 ± 0.19A
`Oil
`2.69 ± 0.97D
`4.74 ± 1.28B
`11.87 ± 0.55C
`14.51 ± 3.28A
`23.88 ± 6.36B
`Gum
`75:25 KO:H2O
`80.67 ± 3.70A
`Oil
`2.97 ± 0.55D
`4.58 ± 0.29B
`15.54 ± 1.52A
`n.d.
`Gum
`16.84 ± 0.87A
`TAG-triglycerides, FFA-free fatty acids, CHOL-cholesterol, DAG-diglycerides, PL-phospholipids. n.d. Not detected.
`A,B,C Different letters mean significant differences within a column for a given class among fractions and ratios (p < 0.05).
`4 RIMFROST EXHIBIT 1174 Page 0004
`
`PL
`46.28 ± 0.32B
`n.d.
`49.73 ± 3.90B
`n.d.
`67.62 ± 1.97A
`
`DAG
`5.63 ± 0.89B
`10.19 ± 0.15A
`n.d.
`11.78 ± 2.57A
`n.d.
`
`

`

`C. Showman, et al.
`
`LWT - Food Science and Technology 126 (2020) 109284
`
`Table 2
`Profile of major fatty acid composition (mg fatty acid/g sample or fraction) of 100% krill oil (KO) and the 50:50 KO:H2O and 75:25 KO:H2O and fractions separated
`by water degumming principles.
`75:25 KO:H2O OIL
`50:50 KO:H2O GUM
`50:50 KO:H2O OIL
`100% KO
`7.54 ± 0.62b
`C14:0
`14.21 ± 0.86a
`5.95 ± 0.43bc
`14.57 ± 0.56a
`20.05 ± 0.37c
`21.32 ± 0.19ab
`20.50 ± 0.37bc
`21.14 ± 0.29abc
`C16:0
`6.78 ± 0.49a
`3.16 ± 0.14b
`6.98 ± 0.18a
`3.84 ± 0.14b
`C16:1 (C9)
`12.56 ± 0.61a
`7.43 ± 0.29b
`13.22 ± 0.19a
`8.42 ± 0.20b
`C18:1 (C9)
`5.17 ± 0.13a
`4.20 ± 0.01b
`5.14 ± 0.12a
`4.38 ± 0.09b
`C18:3 (C9,12,15)
`13.94 ± 0.30a
`7.65 ± 0.15c
`14.05 ± 0.19a
`8.91 ± 0.06b
`C20:2 (C11,14)
`14.12 ± 2.01c
`27.60 ± 0.73ab
`12.44 ± 0.32c
`24.48 ± 0.65b
`C20:5 (EPA)
`6.47 ± 0.90c
`16.77 ± 0.50b
`5.76 ± 0.06c
`14.94 ± 0.51b
`C22:6 (DHA)
`a,b,c Different letters mean significant differences within a row for a given FA among fractions and ratios (p < 0.05).
`
`75:25 KO:H2O GUM
`4.21 ± 1.22c
`21.67 ± 0.70a
`2.10 ± 0.54c
`5.82 ± 0.75c
`3.85 ± 0.15c
`5.88 ± 0.73d
`30.98 ± 2.34a
`19.82 ± 1.69a
`
`Fig. 2. Types of fatty acids found in 100% krill oil (KO) as well as oil and gum fractions of the 75:25 KO:H2O ratio trial. (A) Comparison of SFA (saturated fatty acids),
`UFA (unsaturated fatty acids), total ω3 (omega-3 fatty acids), and total ω6 (omega-6) fatty acids found in each sample. 100% KO is included as a reference for the FA
`comparison. Data are presented as mg/g FA in total FAs. (B) Comparison of SFA/UFA and ω6/ω3 ratios based on FA content of KO and separated 75:25 KO:H2O oil
`and gum fractions. a,b,c Different letters on the top of data bars indicate statistical differences (p < 0.05).
`
`fractions contained a greater concentration of EPA and DHA than the oil
`fractions (p < 0.05). Within all fractions and initial KO, the 75:25
`KO:H2O gum fraction had the greatest concentration of EPA and DHA
`(p < 0.05). The 50:50 KO:H2O ratio contained a lower concentration
`
`of EPA and DHA than the 75:25 KO:H2O sample, but only the DHA
`content was statistically different (p < 0.05). There were no differ-
`ences (p > 0.05) in FA profiles in the oil fractions tested. The con-
`centration of EPA and DHA in the fractions are consistent with others
`
`5 RIMFROST EXHIBIT 1174 Page 0005
`
`

`

`C. Showman, et al.
`
`LWT - Food Science and Technology 126 (2020) 109284
`
`Fig. 3. Representative mass spectrum of 75:25 KO:H2O ratio gum fraction. PC species of interest were identified by m/z. Each are labeled in figure. Relative
`intensities are shown on the y-axis with the mass on the x-axis.
`
`Table 3
`Identified phosphatidylcholine (PC) species containing EPA (20:5) and DHA
`(22:6) with the most abundant PC species by mass of molecule.
`PC species
`75:25 KO:H2O gum fraction
`Concentration (μg/mL)
`Mass (m/z)
`Relative Intensity
`54.77 ± 2.90
`780.56
`45.27 ± 1.07
`PC20:5/20:5
`48.98 ± 4.31
`828.56
`40.41 ± 0.86
`PC20:5/22:6
`15.30 ± 1.08
`852.56
`12.63 ± 0.08
`PC22:6/22:6
`121.12 ± 8.71
`878.57
`100
`PC16:0/20:5
`Mass of the PC species is expressed in Daltons. Data are expressed as mean ±
`standard deviation for PC species relative intensity and concentration.
`that reported the FA profiles of the individual lipid classes. Araujo, Zhu,
`Breivik, Hjelle, and Zeng (2014) separated neutral and polar lipids with
`different types of solvents and the fatty acid profiles, based on % nor-
`malized mole units, showed similar results. The FA composition of PL in
`the krill oil was 28.5% EPA and 13.5% DHA whereas there was 9.7%
`EPA and 3.8% DHA in the TAG (Araujo et al., 2014). Other studies
`reported much lower amounts of EPA and DHA bound to TAG with
`ranges of 1–4% for each fatty acid (Castro-Gomez et al., 2015; Gigliotti
`et al., 2011; Xie et al., 2017). The greater EPA and DHA content in the
`oil fraction may be due to diglycerides (DAGs) separated into this
`fraction (Table 1). Castro-Gomez et al. (2015) examined isolated frac-
`tions of KO and reported the DAG-rich fraction had 19.8% EPA and
`6.1% DHA. The content and distribution of these PUFAs in krill oil are
`what make the product interesting. A PL-rich fraction with a high
`content of ω-3 PUFAs can be a marketable food product.
`The 75:25 KO:H2O oil and gum fractions contained different lipid
`classes; and thus, different types of FAs (Fig. 2A). The FA profile of the
`gum fraction presented a greater concentration of ω-3 PUFAs
`(p < 0.05) than the oil fraction (Fig. 2A). The gum fraction was
`composed of approximately 55 mg/g ω-3 FAs while the oil fraction
`contained 26 mg/g when totaling all ω-3 FA present. The gum fraction
`also had approximately four times lower ω-6/ω-3 ratio when compared
`to the oil fraction (Fig. 2B). The ratios for oil and gum fractions are 0.65
`and 0.16, respectively.
`Fatty acid composition is important especially in relation to human
`health. Diets with a higher ω-6/ω-3 ratio are a promoter of cardio-
`vascular and inflammatory diseases (Simopoulos, 2002). Studies in-
`dicate that a lower level of ω-6 FAs in the blood decreases the occur-
`rence of cardiovascular disease (Lands, 2003). The recommended ratio
`of ω-6/ω-3 fatty acids is 5:1, but the Western diet is far below the
`optimal intake (Küllenberg et al., 2012). At the same time, the oil
`
`fraction had a greater saturated fatty acid/unsaturated fatty acid (SFA/
`UFA) ratio than the gum fractions of the 75:25 KO:H2O samples
`(Fig. 2B). The SFA/UFA ratios were 0.58 and 0.38 for the oil and gum
`fractions of the 75:25 KO:H2O, respectively.
`
`3.4. Identification and quantification of phosphatidylcholine (PC) species
`After hydration and fractionation of the PLs, ultra-high performance
`liquid chromatography mass spectrometry (UHPLC-MS) was used to
`identify and quantify the phosphatidylcholine (PC) species that con-
`currently contained EPA and DHA on sn-1 and sn-2 positions in the gum
`fraction of the 75:25 KO:H2O sample. Fig. 3 shows a representative
`mass spectrum in the range where the PC species of interest are iden-
`tified. Each PC molecule is identified with an arrow. Concentration (μg/
`mL) and relative intensity data for each PC species are shown in
`Table 3. Results indicate the presence of PC-20:5/20:5, PC-20:5/22:6,
`and PC-22:6/22:6 which are consistent with previous studies that found
`PC species in KO do have both EPA and DHA incorporated onto the
`same molecule (Castro-Gomez et al., 2015; Le Grandois et al., 2009;
`Winther et al., 2011). Typically, PLs contain a saturated and un-
`saturated FAs attached to sn-1 and sn-2, respectively. Having two highly
`unsaturated and long chain FAs such as EPA and DHA concurrently
`esterified on one PC molecule is unlikely because of stearic hindrance
`that these FAs exhibit (Berdanier, 2008). It was not determined if the
`EPA or DHA was on the sn-2 position; however, PC species with these ω-
`3 PUFAs on the sn-2 position can increase cardiovascular health as well
`as brain function (Parmentier, Mahmoud, Linder, & Fanni, 2007).
`Of the PLs of interest, PC20:5/20:5 was the most abundant, followed
`by PC20:5/22:6 in the gum fraction of the 75:25 KO:H2O sample
`(p < 0.05). Results are consistent with Winther et al. (2011) who
`reported similar order of abundance with only slightly lower relative
`intensities for each of the PC species of interest. A similar method of
`quantification was used in the previously mentioned study, but only
`relative intensity was reported by Winther et al. (2011). In addition to
`relative intensity, Table 3 also lists concentration of the PC species of
`interest.
`PLs have a different digestibility; and therefore, bioavailability
`when compared to TAGs. Due to the amphiphilic nature of PLs, they are
`not dependent on bile salts forming micelles for lipid digestion.
`Intestinal absorption of PLs is much higher than TAGs as well. Up to
`20% of dietary PLs can be absorbed passively in the intestine
`(Zierenberg & Grundy, 1982). Supplementation of PLs, especially ones
`containing EPA and DHA, will affect blood lipid profiles (Küllenberg
`et al., 2012). Incorporating these PLs esterified with EPA/DHA into
`6 RIMFROST EXHIBIT 1174 Page 0006
`
`

`

`C. Showman, et al.
`
`food or pharmaceutical products would enhance dietary intake of
`beneficial FAs. Lecithin, the name for commercial PLs,

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