throbber
Amino Acids (2011) 40:1369–1383
`DOI 10.1007/s00726-011-0874-6
`
`R E V I E W A R T I C L E
`
`Analysis of the efficacy, safety, and regulatory status
`of novel forms of creatine
`
`Ralf Jäger · Martin Purpura · Andrew Shao ·
`Toshitada Inoue · Richard B. Kreider
`
`Received: 10 July 2010 / Accepted: 30 November 2010 / Published online: 22 March 2011
`© The Author(s) 2011. This article is published with open access at Springerlink.com
`
`Abstract Creatine has become one of the most popular
`dietary supplements in the sports nutrition market. The
`form of creatine that has been most extensively studied and
`commonly used in dietary supplements is creatine mono-
`hydrate (CM). Studies have consistently indicated that
`CM supplementation increases muscle creatine and phos-
`phocreatine concentrations by approximately 15–40%,
`enhances anaerobic exercise capacity, and increases train-
`ing volume leading to greater gains in strength, power, and
`muscle mass. A number of potential therapeutic benefits
`have also been suggested in various clinical populations.
`Studies have indicated that CM is not degraded during
`normal digestion and that nearly 99% of orally ingested
`CM is either taken up by muscle or excreted in urine.
`Further, no medically significant side effects have been
`reported in literature. Nevertheless, supplement manufac-
`turers have continually introduced newer forms of creatine
`
`Invited paper presented at the Creatine in Health and Sport 2010
`conference. Submitted to Amino Acids, 15 June 2010.
`
`R. Ja¨ger · M. Purpura
`Increnovo LLC, 2138 E Lafayette Pl, Milwaukee,
`WI 53202, USA
`
`A. Shao
`Council for Responsible Nutrition, 1828 L Street NW,
`Suite 510, Washington, DC 20036, USA
`
`T. Inoue
`Healthy Navi Co., Ltd., 3-18-1-801, Minami-rokugo,
`Ota-ku, Tokyo 144-0045, Japan
`
`R. B. Kreider (&)
`Exercise and Sport Nutrition Lab, Department of Health and
`Kinesiology, Texas A&M University, 158 Read Building,
`TAMU 4243, College Station, TX 77843-4243, USA
`e-mail: rkreider@hlkn.tamu.edu
`
`into the marketplace. These newer forms have been pur-
`ported to have better physical and chemical properties,
`bioavailability, efficacy, and/or safety profiles than CM.
`However, there is little to no evidence that any of the newer
`forms of creatine are more effective and/or safer than CM
`whether ingested alone and/or in combination with other
`nutrients. In addition, whereas the safety, efficacy, and
`regulatory status of CM is clearly defined in almost all
`global markets; the safety, efficacy, and regulatory status of
`other forms of creatine present in today’s marketplace as a
`dietary or food supplement is less clear.
`
`Keywords Creatine · Dietary supplements ·
`Ergogenic aids · Exercise · Performance
`
`Introduction
`
`Creatine (N-(aminoiminomethyl)-N-methyl glycine) is an
`ingredient commonly found in food, mainly in fish and
`meat, and is sold as a dietary supplement in markets around
`the world. Its use as an ergogenic aid and possible treat-
`ment
`for
`certain neuromuscular disorders
`is well
`documented in scientific literature (Buford et al. 2007;
`Kreider et al. 2010). In recent years, the popularity of
`creatine has risen dramatically, especially among athletes.
`In the USA alone, creatine-containing dietary supplements
`make up a large portion of the estimated $2.7 billion in
`annual sales of sports nutrition supplements (NBJ 2009).
`Accompanying this explosive growth in sales has been
`the introduction of different forms of creatine. Creatine
`monohydrate (CM), first marketed in the early 1990s, is the
`form most commonly found in dietary supplement/food
`products and most frequently cited in scientific literature.
`The introduction into the marketplace of alternate forms of
`
`123
`
`001
`
`Harvest Trading Group - Ex. 1130
`
`

`
`1370
`
`R. Ja¨ger et al.
`
`creatine, beginning in the late 1990s, was presumably an
`attempt to differentiate the multitude of creatine-containing
`products available to consumers and improve certain
`attributes such as solubility and efficacy. However, the
`legal and regulatory status of these various forms of crea-
`tine in the USA and other markets around the world is at
`best uncertain. To date, with the exception of Japan, CM is
`the only form of creatine to be officially approved or
`accepted in key markets such as the USA, European Union
`(EU), Canada and South Korea. The continued presence of
`other forms of creatine in the marketplace, especially in the
`US, may be due to a multitude of factors. These include,
`but may not be limited to, a lack of awareness or under-
`standing on the part of marketers of applicable laws and
`regulations, intentional noncompliance with the law, and/or
`inadequate enforcement of the law. The public health
`implications of widespread distribution and use of these
`unauthorized forms of creatine is unknown and warrants
`careful monitoring.
`New forms of creatine are marketed with claims of
`improved physical, chemical, and physiological properties
`in comparison to CM. Claims include improved stability
`when combined with other
`ingredients or
`in liquids,
`improved solubility in water, improved bioavailability, and
`even an increase in performance. This review will evaluate
`the available literature on new forms of creatine and
`compare them to available data on CM in terms of efficacy
`and safety. In addition, the current international regulatory
`status of the various forms of creatine that are commer-
`cially available will be examined.
`
`Methods
`
`This analysis represents a systematic review of the litera-
`ture on the various forms of creatine available in the global
`marketplace as dietary supplements, food supplements, or
`natural health products. For technical and performance
`comparisons,
`literature searches were performed by
`searching the Medline database of the US National Library
`of Medicine of the National Institutes of Health. The search
`strategy involved entering the various creatine search terms
`(Table 1), along with the technical or performance aspect
`of interest (e.g., solubility, stability, bioavailability, per-
`formance). In addition, a patent research was performed by
`searching the database of the World Intellectual Property
`Organization (WIPO),
`the European Patent Office,
`the
`Japan Patent Office, and the United States Patent and
`Trademark Office. Articles were reviewed, analyzed, and
`interpreted, with results of the relevant studies presented
`below.
`For the assessment of the current regulatory status of the
`various forms of creatine, the Web sites of regulatory
`
`123
`
`Table 1 Creatine content of different forms of creatine
`
`Form of creatine
`
`Creatine
`content (%)
`
`Difference
`in CM (%)
`
`Creatine anhydrous
`
`100.0
`
`CM
`
`Creatine ethyl ester
`
`Creatine malate (3:1)
`
`Creatine methyl ester HCl
`
`Creatine citrate (3:1)
`
`Creatine malate (2:1)
`
`Creatine pyruvate
`Creatine α-amino butyrate
`Creatine α-ketoglutarate
`Sodium creatine phosphate
`
`Creatine taurinate
`
`Creatine pyroglutamate
`
`Creatine ketoisocaproate
`
`Creatine orotate (3:1)
`
`Carnitine creatinate
`
`Creatine decanoate
`
`Creatine gluconate
`
`87.9
`
`82.4
`
`74.7
`
`72.2
`
`66
`
`66
`
`60
`
`56.2
`
`53.8
`
`51.4
`
`51.4
`
`50.6
`
`50.4
`
`45.8
`
`44.9
`
`43.4
`
`40.2
`
`+13.8
`0
`−6.3
`−15.0
`−17.9
`−24.9
`−24.9
`−31.7
`−36.0
`−38.8
`−41.5
`−41.6
`−42.4
`−42.7
`−47.9
`−49.0
`−50.7
`−54.3
`
`bodies for key markets around the world were accessed
`(USA: US Food and Drug Administration; Canada: Health
`Canada; EU: European Commission; Japan: Ministry of
`Health, Labor and Welfare; Korea: Korea Food and Drug
`Administration). Information derived from these sites was
`used to determine the legal and regulatory framework
`governing creatine products in these markets and the cur-
`rent regulatory status of the various forms of creatine as
`dietary supplements, food supplements, and natural health
`products.
`
`Physio-chemical properties
`
`Creatine crystallizes from water as monoclinic prisms
`holding one molecule of water of crystallization per mol-
`ecule of creatine. Continued drying of CM results in a loss
`of the water of crystallization at around 100°C, yielding
`anhydrous creatine. Creatine is a weak base with a pkb
`value of 11.02 at 25°C. As a result, creatine can only form
`salts with strong acids, having a pka value of less than 3.98.
`Creatine forms salts by the protonation of its guanidine
`moiety (see Fig. 1). In addition to salt formation, creatine is
`able to act as a complexing agent.
`Creatine salts such as citrate, maleate, fumarate, tartrate
`(Negrisoli and Del Corona 1997), pyruvate (Pischel and
`Weiss 1996), ascorbate (Pischel et al. 1999), and orotate
`(Abraham and Jiang 2005) were first introduced to the
`marketplace as early as the late 1990s. Creatine and acids
`
`002
`
`Harvest Trading Group - Ex. 1130
`
`

`
`Efficacy and safety of novel creatine forms
`
`1371
`
`NH2
`
`H2N
`
`N
`CH3
`
`O
`
`O
`
`+
`
`HA
`
`H2O
`
`A
`
`H2N
`
`NH2
`
`N
`CH3
`
`O
`
`OH
`
`Creatine
`
`Strong Acid
`
`Creatine-Salt
`
`Fig. 1 Creatine salts
`
`with multiple acid moieties such as citric acid can form
`salts as well as complexation products. The first acid
`moiety of citric acid is strong enough (pka = 3.09) to form
`a salt with creatine; however,
`the other two moieties
`(pka2 = 4.75, pka3 = 5.41) should only be able to form
`complexes with creatine. A salt, a salt-complex combina-
`tion, or a simple physical mixture can be differentiated by
`measurement of the enthalpy changes of neutralization,
`which ranges usually in the area of −55 to −66 kJ/mole for
`the salt formation to less than −5 kJ/mole for the change in
`complexation enthalpy to no changes in enthalpy for a
`physical mixture (1995). A “tricreatine citrate” is actually a
`complex of creatine citrate with two additional creatine
`moieties, resulting in a molecule with a ratio of creatine to
`citrate of 3:1.
`In addition to creatine and its salts, derivatives of cre-
`atine such as creatine ester or even creatine alcohols are
`currently marketed as dietary supplements in the USA (see
`Fig. 2). Both ingredients do not contain creatine as such,
`since they have been chemically altered. While it
`is
`assumed that the human body will transfer those molecules
`into creatine upon intake,
`there are no published data
`available to base firm conclusions.
`The amount of creatine in different forms of creatine
`varies. Creatine monohydrate contains 87.9% of creatine,
`whereas the creatine content in other forms of creatine is
`lower with the exception of creatine anhydrous (see
`Table 1). Commercial creatine salts are formed in solution
`or by mechanical processes such as milling or grinding
`under the presence of residual water. Complexes are
`formed by the subsequent replacement of the solvating
`molecules by the new ligands.
`
`Solubility
`
`correlation between solubility and temperature is almost
`linear. One liter of water dissolves 6 g of creatine at 4°C,
`14 g at 20°C, 34 g at 50°C, and 45 g at 60°C. The solubility
`of creatine can also be increased by lowering the pH of the
`solution. This principle is the basis for the improved sol-
`ubility of creatine salts, since creatine salts lower the pH of
`water due to the nature of acid moiety. Creatine monohy-
`drate dissolves at 14 g/L at 20°C resulting in a neutral pH
`of 7. A saturated solution of tricreatine citrate in water has
`a pH of 3.2; whereas a saturated solution of creatine
`pyruvate even has a pH of 2.6 (pyruvic acid is a stronger
`acid than citric acid). The decrease in pH results in an
`increase in solubility: 29 g/L creatine citrate at 20°C, and
`54 g/L creatine pyruvate at 20°C. Normalized by the rel-
`ative amount of creatine per molecule (monohydrate
`87.9%, citrate 66%, pyruvate 60%), creatine citrate
`(19.14 g/L) shows a 1.55-fold and creatine pyruvate
`(32.4 g/L) a 2.63-fold better solubility when compared with
`the monohydrate (12.3 g/L). Whereas the creatine deriva-
`tive creatinol-O-phosphate (5 g/L at 20°C) has inferior
`solubility, dicreatinol sulfate (1,370 g/L at 20°C) shows
`superior solubility when compared with CM, creatine salts,
`or creatine esters (Godfraind et al. 1983; Gastner et al.
`2005).
`
`Stability
`
`Stability in solid form
`
`Creatine monohydrate powder is very stable showing no
`signs of degradation over years, even at elevated temper-
`atures. To detect a potential degradation of creatine, one
`must measure the content of its degradation product, cre-
`atinine (see Fig. 3), which can be quantified by HPLC at
`levels as low as 67 parts per million (ppm). At room
`temperature and even at an increased temperature of 40°C
`(104°F), CM shows no signs of degradation (i.e., creatinine
`levels stay under the quantification limit of 67 ppm) after
`more than 3 years. As Fig. 3 shows, even when stored at
`60°C (140°F), creatinine (106 ppm) was only detected after
`a period of 44 months (Ja¨ger 2003).
`
`Stability in solutions
`
`One major limitation of creatine as an ampholytic amino
`acid is its rather low solubility in water. The solubility of
`creatine in water
`increases with temperature and the
`
`As shown in Fig. 4, in contrast to its stability in a solid
`state, creatine is not stable in aqueous solution due to an
`
`Fig. 2 Chemical structure of
`creatine and creatine derivatives
`
`NH2
`
`H2N
`
`O
`
`N
`CH3
`
`O
`
`NH
`
`H2N
`
`O
`
`N
`CH3
`
`O
`
`CH3
`
`H2N
`
`NH
`
`N
`CH3
`
`O
`O P
`OH
`
`OH
`
`Creatine
`
`Creatine Ethylester
`
`Creatinol-O-Phosphate
`
`123
`
`003
`
`Harvest Trading Group - Ex. 1130
`
`

`
`1372
`
`R. Ja¨ger et al.
`
`Fig. 5 Effect of pH on creatine stability in solution. Adapted from
`Howard and Harris (1999)
`
`ingredient. If creatine is not consumed immediately after it
`has been dissolved in water, it should be stored at a low
`temperature to retard the degradation.
`The degradation of creatine can be reduced or even
`halted by either lowering the pH under 2.5 or increasing the
`pH. A very high pH results in the deprotonation of the acid
`group, thereby slowing down the degradation process by
`making it more difficult for the intramolecular cyclization.
`A very low pH results in the protonation of the amide
`function of the creatine molecule, thereby preventing the
`intramolecular cyclization (see Fig. 6). This effect also
`occurs under the acidic conditions in the stomach, hence
`preventing the breakdown of creatine. The conversion of
`creatine to creatinine in the gastrointestinal tract is minimal
`regardless of transit time (Persky et al. 2003; Harris et al.
`1992b; Deldicque et al. 2008).
`
`Stability of other forms of creatine
`
`Some creatine salts appear to be less stable when compared
`with CM. Tricreatine citrate results in creatinine levels of
`770 ppm at 40°C (104°F) after 28 days of storage. How-
`ever, the addition of carbohydrates has been shown to
`increase stability of some creatine salts (Purpura et al.
`2005). Creatine salts are not expected to have a greater
`stability in solution; however, the pH lowering effect of the
`salt might reduce stability compared to CM in the same
`environment.
`Tallon et al. (Child and Tallon 2007) compared the
`stability of creatine ethyl ester (CEE) head to head with
`CM and found that CEE was actually less stable than CM.
`It was concluded that the addition of the ethyl group to
`creatine actually reduced acid stability and accelerated its
`breakdown to creatinine. The degradation of creatine and
`
`Fig. 3 Stability of creatine monohydrate powder. Adapted from Ja¨ger
`(2003)
`
`intramolecular cyclization (Howard and Harris 1999). The
`rate of creatine degradation in solution is not dependent on
`its concentration, but on pH. Generally, the lower the pH
`and higher the temperature, the faster is the degradation.
`This solid-state and degradation properties have been
`thoroughly investigated as early as the 1920s (Edgar and
`Shiver 1925; Cannon et al. 1990) and more recently by
`Dash et al. (2002), as well as Harris et al. (Howard and
`Harris 1999). These researchers found that whereas crea-
`tine was relatively stable in solution at neutral pH (7.5 or
`6.5), a lowering of pH resulted in an increased rate of
`degradation and after only 3 days of storage at 25°C cre-
`atine degraded significantly: 4% at pH 5.5; 12% at pH 4.5;
`and 21% at pH 3.5 (see Fig. 5). Similarly, Ganguly et al.
`(2003) reported that creatine monohydrate stored at room
`temperature degraded into creatinine within several days
`but
`that refrigerating creatine monohydrate in solution
`slowed degradation. The rapid degradation of creatine in
`solution precludes the manufacture of shelf-stable standard
`acidic beverages containing efficacious amounts of the
`
`O
`
`HN
`
`N
`
`H3C
`Creatinine
`
`- H2O
`
`HN
`
`OH
`
`+ H2O
`
`NH2
`
`HN
`
`N
`
`H3C
`
`O
`
`Creatine
`
`Fig. 4 Degradation of creatine to creatinine
`
`123
`
`004
`
`Harvest Trading Group - Ex. 1130
`
`

`
`Efficacy and safety of novel creatine forms
`
`1373
`
`NH2
`
`O
`
`+ H2O
`
`- H2O
`
`HN
`
`N
`
`H3C
`
`O
`
`O
`
`HN
`
`HN
`
`N
`
`H3C
`
`NH3
`
`HN
`
`N
`
`H3C
`
`- H2O
`
`+ H2O
`
`OH
`
`O
`
`Creatine at very low pH
`
`Creatinine
`
`Creatine at very high pH
`
`Fig. 6 Very low pH prevents
`creatine degradation
`
`O
`
`HN
`
`HN
`
`N
`
`H3C
`
`eninitaerC
`
`Fig. 7 Degradation of creatine
`and creatine ester
`
`NH2
`
`- ROH
`
`+ ROH
`
`R
`
`O O
`
`H2
`N
`
`HN
`
`N
`
`H3C
`
`pH < 7
`
`pH > 7
`
`HN
`
`N
`
`H3C
`
`OR
`
`O
`
`evitavireD-enitaerC
`
`R = OH (Creatine)
`R = OC2H5 (Creatine Ethyl Ester)
`R = OCH3 (Creatine Methyl Ester)
`
`creatine ester involves intramolecular hydrolysis of a car-
`boxyl acid (in case of creatine) or carboxylic ester (in case
`of creatine methyl- or ethyl ester) under acidic conditions
`and the rate of degradation depends on the leaving group
`(see Fig. 7). It is speculated that the methyl ester or ethyl
`ester groups are better leaving groups than hydroxyl or
`water and, therefore, suggesting that the degradation into
`creatinine should be rather accelerated.
`These findings are also in accordance with the recent
`investigations on the stability of CEE at 37ºC in both water
`and phosphate-buffered saline and the in vitro response of
`CEE to incubation in human plasma by H-NMR analysis
`(Giese and Lecher 2009b). The conversion of CEE to
`creatine by the esterases in human plasma was not detec-
`ted, and the only species detected after the incubation
`period was creatinine. It is concluded that CEE is mostly
`converted into creatinine under physiological conditions
`encountered during transit
`through the various tissues,
`suggesting no ergogenic effect is to be expected from
`supplementation of CEE. The high stability of CM is well
`documented, whereas the stability of newer forms of cre-
`atine (salts, ester, etc.) either has not been investigated or
`appears to be inferior. New forms of creatine contain less
`of the active principal creatine in comparison to CM;
`however, creatine salts can offer an advantage over CM
`with regard to solubility.
`
`Bioavailability
`
`The uptake of creatine is simplified in a two-step approach:
`first, uptake into the blood stream; second, uptake into the
`
`target tissue. The term ‘bioavailability’ refers to both the
`intestinal absorption and the use of a substance by the
`body’s cells and tissues. First indications of a potential
`change of creatine bioavailability can be gathered from the
`amount of creatine taken up into the blood plasma after
`oral administration. However, a change in the total amount
`of creatine in the blood plasma cannot be directly extrap-
`olated to a potential increase in desired performance. An
`increased amount of creatine in the plasma could be the
`result of decreased uptake into the target tissue resulting in
`an actual decrease in overall bioavailability. On the other
`hand, an initial rise in plasma creatine levels, followed by a
`reduction in plasma levels, is an indication of increased
`uptake into the target tissue. This has been demonstrated in
`vivo by combining creatine with insulin-stimulating
`ingredients such as high amounts of glucose or protein
`(Bessman and Mohan 1992; Haughland and Chang 1975;
`Rooney et al. 2002). Conclusive proof of an increase in
`relevant bioavailability can only be gained by assessing the
`amount of creatine reaching the target tissue, the muscle,
`measured by muscle biopsy and/or whole body creatine
`retention assessed by measuring the difference between
`creatine intake and urinary excretion.
`Dietary creatine is presumed to have high bioavailability
`since intestinal absorption of CM is already close to 100%
`(Deldicque et al. 2008). However, the response to creatine
`supplementation is heterogeneous, due in part to some non-
`responders, which might be overcome by alternative forms
`of creatine (Greenhaff 1997b; Greenhaff et al. 1993).
`Several studies have examined whether different forms of
`creatine are more effective in terms of promoting muscle
`uptake of creatine than CM. For example, a recent study
`
`123
`
`005
`
`Harvest Trading Group - Ex. 1130
`
`

`
`1374
`
`R. Ja¨ger et al.
`
`examined the effect of the administration of three different
`forms of creatine on plasma creatine concentrations and
`pharmacokinetics. In a balanced cross-over designed study,
`six healthy subjects were assigned to ingest a single dose of
`isomolar amounts of creatine (4.4 g) in the form of CM,
`tricreatine citrate (TCC), or creatine pyruvate (CPY), fol-
`lowed by measurement of the plasma creatine levels (Ja¨ger
`et al. 2007). Mean peak concentrations and area under the
`curve (AUC) were significantly higher with CPY (17 and
`14%, respectively) in comparison to CM. The findings
`suggest
`that different forms of creatine may result
`in
`slightly different kinetics of plasma creatine absorption,
`although differences in velocity constants of absorption
`could not be detected due to the small number of blood
`samples taken during the absorption phase. The small
`differences in kinetics are unlikely to have any clinically
`relevant effects on muscle creatine elevation during periods
`of creatine loading. A follow-up study including muscle
`biopsies would be required to conclude if the bioavail-
`ability of this specific creatine salt was indeed higher
`(Fig. 8).
`Greenwood et al. (2003) investigated how different
`forms of creatine affect whole body creatine retention.
`Sixteen males were assigned to ingest in a single blind
`manner either 5 g of dextrose, 5 g of CM, 5 g of CM plus
`18 g dextrose, or an effervescent creatine supplement
`consisting of 5 g of TCC (66% creatine) plus 18 g dextrose
`four times/day for 3 days. Creatine retention was estimated
`by subtracting total urinary creatine excretion from total
`supplemental creatine intake over the 3-day period. Results
`revealed that average daily creatine retention over the
`
`Plasma creatine
`µmol/l
`
`5g CM
`
`6.7g CC
`
`7.3g CYP
`
`1000
`
`800
`
`600
`
`400
`
`200
`
`0
`
`0
`
`2
`
`4
`Time (h)
`
`6
`
`8
`
`Fig. 8 Comparison of blood plasma levels of different forms of
`creatine. Adapted from Ja¨ger et al. (2007)
`
`123
`
`Fig. 9 Percentage of creatine retained during a 3-day loading period
`(20 g/day). Adapted from Greenwood et al. (2003)
`
`3-day period was 12.2 ± 1.3, 16.1 ± 2.2, and 12.6 ± 2.5 g/
`day for the CM, CM with dextrose, and effervescent TCC
`groups, respectively. This amounted to whole body crea-
`tine retention of 61 ± 15% for the CM group, 80 ± 11%
`for the CM plus dextrose group, and 63 ± 13% for the
`effervescent TCC group. While creatine retention was
`significantly greater in the CM and dextrose group, no
`significant differences were seen between the CM and
`effervescent TCC groups. These findings suggest that while
`consuming a relatively small amount of dextrose with CM
`can increase whole body creatine retention, supplementa-
`tion of TCC in an effervescent form does not augment
`whole body creatine retention more than CM alone (Fig. 9).
`Over the years, there has been significant commercial
`interest in determining whether creatine could be delivered
`in a liquid form. The thought has been since CM is rela-
`tively insoluble that development of a liquid or suspended
`form of creatine may be more convenient to consume, be
`more readily absorbed into the blood stream, and promote a
`greater efficiency in transport of creatine to the muscle.
`Some companies have even claimed that minimal amounts
`of liquid creatine would need to be ingested because of
`enhanced efficiency in transport through the blood and into
`the muscle. A limitation with these theories is that CM is
`not stable for any substantial length of time in liquid.
`Consequently, while researchers have been working on
`ways to suspend creatine within gels and fluids, it has been
`generally considered to be impractical to develop into a
`product due to limitations in shelf-life. In addition, while
`people may prefer the taste of liquid or gel versions of
`creatine, there is no evidence that these delivery forms
`provide a superior performance benefit.
`Kreider et al. (2003b) carefully compared the effects
`of ingesting 20 g/day of CM to recommended doses
`
`006
`
`Harvest Trading Group - Ex. 1130
`
`

`
`Efficacy and safety of novel creatine forms
`
`1375
`
`(2.5 g/day of CM), as well as doses that would purportedly
`provide an equivalent amount of CM per day in liquid form
`(20 g/day) on muscle creatine, phosphocreatine, and total
`creatine levels. Subjects donated muscle biopsies prior to
`and following 5 days of supplementing their diet in a
`randomized and double-blind manner with either 5 mL of
`creatine liquid (purportedly providing 2.5 g of CM), 5 mL
`of a flavored placebo, 8 9 5 mL doses of creatine liquid
`(purportedly providing 20 g/day of CM), or 8 9 5 mL
`doses of a flavored placebo. Another group ingested
`4 9 5 g of CM for 5 days as a non-blinded benchmark
`control. This analysis allowed for a comparison of ingest-
`ing recommended doses of liquid creatine to a placebo, as
`well as seven times the amount recommended by the
`manufacturer that would purportedly provide an equal
`amount of CM. The researchers found that CM supple-
`mentation significantly increased muscle free creatine
`content by 31 ± 28%. However, none of the other groups
`experienced any effect on muscle free creatine, phospho-
`creatine, or total creatine content. Moreover, changes in
`muscle creatine and phosphocreatine levels in response to
`CM supplementation were significantly greater than the
`liquid creatine and placebo groups. These findings indicate
`that
`liquid creatine supplementation has no effect on
`muscle phosphagen levels and therefore may have no
`ergogenic value. While other groups have been attempting
`to develop stable forms of liquid and/or gel forms of cre-
`atine with some success, there are no data available to date
`demonstrating that these types of creatine are absorbed
`more efficiently and/or have greater benefit compared to
`CM (Fig. 10).
`An alternative dissolved form of creatine is colloidal
`CM. CM is dissolved in its own crystal water and dispersed
`into a stable protective matrix containing carbohydrates
`(Kessel et al. 2004). The product is claimed to be the only
`solubilized form of powdered creatine in the market,
`
`making it more bioavailable and stable. However, no evi-
`dence has been published to date to substantiate any
`performance or ergogenic benefit
`from this form of
`creatine.
`Creatine ethyl ester has been purported to be a superior
`form of creatine in comparison to CM. However, prior
`studies have shown that it degrades rather quickly to cre-
`atinine when exposed to low pH levels as would be found
`in the stomach (Giese and Lecher 2009a; Katseres et al.
`2009). Theoretically, this would reduce the bioavailability
`of creatine. To test this hypothesis, Spillane et al. (2009)
`compared the effects of supplementing the diet with a
`placebo, CM, or CEE during 42 days of training. Serum
`creatinine and muscle total creatine content was assessed
`prior to and following 6, 27, and 48 days of supplemen-
`tation and training. The researchers found that serum
`creatinine levels were significantly increased in the CEE
`group after 6, 27, and 48 days of supplementation indi-
`cating less efficient bioavailability. In addition, while CEE
`supplementation promoted a modest increase in muscle
`total creatine content, it was increased to a greater extent in
`the CM group. These findings directly contradict claims
`that CEE is more effective in increasing muscle creatine
`stores. Further, the significantly higher creatinine levels
`observed should raise some potential safety concerns about
`potential safety (Fig. 11).
`Several studies have also evaluated whether co-inges-
`tion of creatine with other nutrients may influence creatine
`retention. Initial work by Green and colleagues (Green
`et al. 1996a, b) demonstrated that co-ingesting creatine
`(5 g) with large amounts of glucose (e.g., 95 g) enhanced
`creatine and carbohydrate storage in muscle. Subsequent
`studies by Steenge et al. (2000) found ingesting creatine
`(5 g) with 47–97 g of carbohydrate and 50 g of protein also
`enhanced creatine retention. The researchers suggested that
`
`Fig. 10 Change in muscle-free creatine content in response to 5 days
`of low or high-dose creatine serum and placebo ingestion compared to
`CM. Adapted from Kreider et al. (2003b)
`
`Fig. 11 Changes in total muscle creatine content in response to
`placebo (PLA), creatine monohydrate (CRT), and creatine ethyl ester
`(CEE) supplementation (Spillane et al. 2009). †Significantly different
`from PLA group. *Significant difference from baseline. Reprinted
`with permission
`
`123
`
`007
`
`Harvest Trading Group - Ex. 1130
`
`

`
`1376
`
`R. Ja¨ger et al.
`
`creatine transport was mediated in part by glucose and
`insulin. As a result, additional research has been under-
`taken to assess the effect of co-ingesting creatine with
`nutrients that may enhance insulin sensitivity on creatine
`retention.
`Several studies have examined whether co-ingesting
`creatine with D-pinitol
`influences whole body creatine
`retention. In the first study (Greenwood et al. 2001), 12
`male subjects with no history of creatine supplementation
`donated 24-h urine samples for 4 days. After an initial
`control day designed to determine normal daily creatine
`excretion rates, subjects were then matched according to
`body mass and randomly assigned to ingest in a single-
`blind manner either a placebo (4 9 5 g doses of dextrose),
`CM (4 9 5 g), CM with low-dose D-pinitol (4 9 5 g CM
`with 2 9 0.5 g of D-pinitol), or CM with high-dose D-pinitol
`(4 9 5 g CM with 4 9 0.5 g D-pinitol) for 3 days. Whole
`body creatine retention was estimated by subtracting total
`urinary creatine excretion from total supplemental creatine
`intake over the 3-day period. Results revealed that whole
`body creatine retention over the 3-day loading period was
`significantly greater in the low-dose D-Pinitol group in
`comparison to the group ingesting CM alone. However, no
`differences were seen between CM alone and CM with the
`higher dose of D-Pinitol (Fig. 12).
`In a follow-up study, Kerksick et al. (2009) examined
`whether co-ingestion of D-pinitol with CM would affect
`training adaptations, body composition, and/or whole-body
`creatine retention in resistance-trained males. In the study,
`24 resistance trained males were randomly assigned in a
`double-blind manner to CM + D-pinitol or CM alone prior
`to beginning a supervised 4-week resistance training pro-
`gram. Subjects ingested a typical loading phase (i.e., 20 g/
`day for 5 days) before ingesting 5 g/day for the remaining
`
`23 days. Results revealed that creatine retention increased
`in both groups as a result of supplementation. However, no
`significant differences were observed between groups in
`training adaptations. Consequently, additional research is
`needed to determine whether D-pinitol supplementation
`enhances creatine uptake and/or affects the ergogenicity of
`creatine supplementation before firm conclusions can be
`drawn.
`Russian tarragon (Artemisia dracunculus) is an etha-
`nolic extract that is often used as a cooking herb. Studies
`have shown that Russian tarragon (RT) appears to have
`antihyperglycemic activity when combined with CM
`ingestion (Ja¨ger et al. 2008a; Wang et al. 2008). Theoret-
`ically, ingesting RT extract prior to creatine loading may
`enhance insulin sensitivity and thereby promote greater
`creatine absorption/retention. To support this hypothesis,
`Ja¨ger et al. (2008a) reported that RT influences plasma
`creatine levels during the ingestion of CM in a similar
`manner to glucose and protein. However, further research
`is needed to evaluate the effects of RT on creatine uptake
`and retention in muscle before conclusions can be drawn
`(Fig. 13).
`In analysis of this literature, it is clear that CM sup-
`plementation promotes significant
`increases in muscle
`creatine levels in most individuals. There is some evidence
`that co-ingestion of CM with various nutrients (e.g., car-
`bohydrate, protein, D-pinitol) may enhance creatine uptake
`to a greater degree. However, there is no evidence that
`effervescent creatine, liquid creatine, and/or CEE promotes
`greater uptake of creatine to the muscle. Rather, there is
`some evidence that some of these forms of creatine may be
`less effective and/or be of greater clinical concern in terms
`of safety.
`
`Creatine µmol/l
`1000
`
`CM plus RT
`
`CM
`
`750
`
`500
`
`250
`
`
`
`00
`
`200
`
`0
`
`-200
`
`-400
`
`
`
`* ** *
`
`
`
`**
`
`A – B @ 30, 60, 90 & 120 min
`
`0
`
`20
`
`40
`
`80
`60
`Time (min)
`
`100
`
`120
`
`Fig. 13 Influence of Russian tarragon on creatine absorption.
`Adapted from Ja¨ger et al. (2008a)
`
`Fig. 12 Change in whole body creatine r

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