throbber
Physicochemical Characterization of Creatine
`N-Methylguanidinium Salts
`
`Brandon T. Gufford
`Kamaraj Sriraghavan
`Nicholas J. Miller
`Donald W. Miller
`Xiaochen Gu
`Jonathan L. Vennerstrom
`Dennis H. Robinson
`
`ABSTRACT. Creatine is widely used as a dietary supplement for body builders to
`enhance athletic performance. As the monohydrate, its low solubility in water and high
`dose lead to water retention and gastrointestinal discomfort. Hence, alternative creatine
`derivatives with enhanced water solubility and potential therapeutic advantages have
`been synthesized. As a zwitterionic compound, creatine can form salts at the N-methyl
`guanidinium or carboxylic acid functional groups. In this study, we determined the
`aqueous solubilities and partition coefficients of six N-methyl guanidinium salts of cre-
`atine compared to those of creatine monohydrate; two of these were new salts, namely,
`creatine mesylate and creatine hydrogen maleate. The aqueous solubilities of the salts
`were significantly more than that of creatine monohydrate with the hydrochloride and
`mesylate being 38 and 30 times more soluble, respectively. The partition coefficients
`of the creatine salts were very low indicating their relatively high polarity. Perme-
`abilities of creatine pyruvate, citrate, and hydrochloride in Caco-2 monolayers were
`compared to that of creatine monohydrate. Aside from the creatine citrate salt form
`that had reduced permeability, there were no significant differences in permeability
`characteristics in Caco-2 monolayers. Typical of an amphoteric compound, creatine
`is least soluble in the pH region near the isoelectric point.
`
`Brandon T. Gufford, Kamaraj Sriraghavan, Jonathan L. Vennerstrom and Dennis H. Robinson
`are affiliated with the Department of Pharmaceutical Sciences, College of Pharmacy, University of
`Nebraska Medical Center, Nebraska Medical Center, Omaha, Nebraska, USA.
`Nicholas J. Miller, Donald W. Miller, and Xiaochen Gu are affiliated with the Department of
`Pharmacology and Therapeutics, University of Manitoba, Winnipeg, Manitoba, Canada.
`Address correspondence to: Dennis H. Robinson, Department of Pharmaceutical Sciences, College
`of Pharmacy, University of Nebraska Medical Center, 986025 Nebraska Medical Center, Omaha, NE,
`USA (E-mail: dhrobins@unmc.edu).
`We thank UNeMed, Omaha, NE for the financial support of this work.
`
`Journal of Dietary Supplements, Vol. 7(3), 2010
`Available online at www.informaworld.com/WJDS
`C(cid:1) 2010 by Informa Healthcare USA, Inc. All rights reserved.
`doi: 10.3109/19390211.2010.491507
`
`240
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`001
`
`Harvest Trading Group - Ex. 1109
`
`

`
`Gufford et al.
`
`241
`
`KEYWORDS. Creatine salts, creatine monohydrate, physicochemical properties,
`solubility, log P, thermal analysis, permeability
`
`INTRODUCTION
`
`Creatine, most commonly in the form of the monohydrate, is widely used as a dietary
`supplement for bodybuilding and for its potential to improve exercise and athletic
`performance (Persky & Brazeau, 2001). Creatine may also be useful in the treatment
`of certain diseases, especially those involving muscular atrophy or fatigue associated
`with impaired energy production (Persky & Brazeau, 2001; Wyss & Kaddurah-Daouk,
`2000). As creatine monohydrate is typically used at relatively high doses (5–25 g/day),
`its water solubility of 16.6 mg/ml (Dash & Sawhney, 2002) dictates that it is administered
`as an oral suspension, leading to water retention and possible gastrointestinal discomfort
`(Persky & Brazeau, 2001).
`It is possible to make salts of the zwitterionic creatine at either the carboxylic acid or
`N-methyl guanidinium functional groups (Figure 1). Due to relatively low aqueous sol-
`ubility of creatine monohydrate, there have been several creatine N-methylguanidinium
`salts introduced in the market as dietary supplements that are claimed to be more
`water-soluble than creatine monohydrate. These include the pyruvate, hydrogen citrate,
`and hydrochloride salt forms. Other creatine N-methylguanidinium salts have been de-
`scribed in patents and/or patent applications. These salt forms include the hydrobromide
`(Rudakova, Pospelova, & Yurkevich, 1967); nitrate (Dhar & Ghosh, 1961); mesylate
`(Blatt et al., 2006); dihydrogen phosphate and hydrogen oxalate (Dhar & Ghosh, 1961);
`malate (Boldt, 2004; Cornelius & Haynes, 2004; Qian, Ye, & Huang, 2005); maleate,
`fumarate, and tartarate (Boldt, 2004); lipoate (Buononato & Festuccia, 2003; Gardiner,
`2000); hydrogen maleate, hydrogen fumarate, hydrogen tartarate, and hydrogen malate
`(Negrisoli & Del Corona, 1996); ascorbate (Pischel, Weiss, Gloxhuber, & Mertschenk,
`1998); and bicarbonate (Kneller, 2008).
`Many of these creatine N-methylguanidinium salts are formed with acids, which are
`insufficiently acidic to directly form a salt with creatine (Figure 1). Thus, it is not clear if
`these “salts” are merely physical mixtures, weak complexes, or “addition salts” (Pischel
`et al., 1998) of creatine monohydrate and the acid in question. The only way to form such
`salts with weak acids is using an ion-exchange between a creatine N-methylguanidinium
`salt and a salt of the acid as described by Qian et al. (2005) in the synthesis of creatine
`maleate using creatine hydrochloride or creatine sulfate with sodium or calcium maleate;
`by Arnold (2001) in the synthesis of creatine pyruvate from creatine hydrochloride and
`sodium pyruvate; and by Gardiner, Heuer, and Molino (2006) in the synthesis of creatine
`citrate from tripotassium citrate and creatine hydrochloride.
`
`FIGURE 1. Creatine acid–base equilibria.
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`002
`
`Harvest Trading Group - Ex. 1109
`
`

`
`242
`
`JOURNAL OF DIETARY SUPPLEMENTS
`
`The purpose of this study was firstly, to synthesize additional salts, namely creatine
`mesylate and creatine hydrogen maleate, and secondly, to determine and compare the
`physicochemical properties of these and other commercially available salts, namely
`creatine hemisulfate, creatine hydrochloride, creatine pyruvate, and “creatine citrus”
`(citrate complex). The aim was to identify a water soluble salt that may have the follow-
`ing potential advantages over creatine monohydrate: (a) increase solubility promoting
`increased bioavailability (b) lower oral dose (c) decrease side effects (d) be able to be
`formulated into a more diverse range of formulations e.g., capsules or a topical product
`for (e) a wider range of therapeutic applications e.g., anti-inflammatory effects.
`
`MATERIALS AND METHODS
`
`Materials
`
`Creatine monohydrate (CreapureTM) and creatine pyruvate were obtained from
`Degussa, creatine citrate, i.e., “creatine citrus” was obtained from Peak Nutrition, and
`creatine hydrochloride (Miller, Vennerstrom, & Faulkner, 2009) was obtained from
`Vireo Systems. With the exception of the 1-octanesulfonic acid sodium salt that was
`obtained from Fluka, all other reagent solvents were obtained from Sigma-Aldrich and
`used as received.
`
`Nuclear Magnetic Resonance (NMR), Elemental Analysis, and Melting Point
`Determination
`
`Proton Nuclear Magnetic Resonance (1H NMR) spectra of each new salt in DMSO-d6
`were recorded on a 500-MHz spectrometer. All chemical shifts are reported in parts
`per million (ppm) and are relative to internal TMS (0 ppm). Elemental analyses were
`determined by M-H-W Laboratories. Melting points were obtained using differential
`scanning calorimetry and were uncorrected.
`
`Creatine Mesylate
`
`A suspension of creatine monohydrate (0.08 mol, 11.93 g) in deionized water
`◦
`(120 ml) warmed to 59
`C was added to a stirred solution of methane sulfonic acid
`◦
`C. The reaction mixture was stirred
`(0.08 mol, 7.69 g) in ethanol (EtOH) (200 ml) at 59
`◦
`C and then allowed to cool to room temperature. The solvents were
`for 10 min at 59
`removed in vacuo affording creatine mesylate (11.19 g, 58%) as a white crystalline solid
`◦
`C; 1H NMRδ 2.40
`that was filtered and washed with cold EtOH: melting point, 179.8
`(s, 3H), 2.95 (s, 3H), 4.16 (s, 2H), 7.44 (s, 4H). Anal. calculated for C5H13N3O5S: C,
`26.43; H, 5.77; N, 18.49; found: C, 26.60; H, 5.70; N, 18.39.
`
`Creatine Hydrogen Maleate
`
`To a stirred solution of maleic acid (0.04 mol, 10.40 g) in EtOH was added a
`◦
`suspension of creatine monohydrate (0.04 mol, 5.98 g) in water (80 ml) at 59
`C. The
`◦
`C before cooling to room temperature. The reaction
`solution was stirred for 10 min at 59
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`003
`
`Harvest Trading Group - Ex. 1109
`
`

`
`Gufford et al.
`
`243
`
`◦
`C for 24 hr after which a creatine monohydrate precipitate
`solution was then cooled to 5
`that got formed was collected using suction filtration. Cold acetonitrile (300 ml) was
`◦
`added to the filtrate and cooled to 0
`C to precipitate additional creatine monohydrate
`that was then removed using suction filtration. This process was repeated thrice until no
`additional precipitate was observed. After removing the solvents in vacuo, crude creatine
`hydrogen maleate (6.18 g, 59%) was filtered and washed with cold acetonitrile. Three
`crystallizations from 80% aqueous EtOH afforded analytically pure creatine hydrogen
`◦
`C; 1H NMRδ 2.95 (s, 3H), 4.14 (s, 2H), 6.03 (s, 2H), 7.33
`maleate: melting point 157.8
`(s, 4H). Anal. calculated for C8H13N3O6: C, 38.87; H, 5.21; N, 17.00; found: C, 38.71;
`H, 5.49; N, 16.85.
`
`HPLC Analysis of Creatine Salts
`
`The HPLC system consisted of a Shimadzu SCL-10A controller, an SIL-10AF au-
`tosampler, dual LC-10AT pumps, an SPD-10A UV-VIS detector set to monitor ab-
`◦
`C, with a
`sorbance at both 210 and 235 nm, and a CTO-10AS column oven set to 30
`T3 column (4.6 × 100 mm, 3 µm, C18). The isocratic mobile phase
`Waters Atlantis R(cid:1)
`consisted of 20% v/v acetonitrile, 5-mM formic acid, and 5-mM 1-octanesulfonic acid
`sodium salt (apparent pH 2.8) at a flow rate of 1.5 ml/min. Simultaneous quantitative
`determination of creatine and creatinine content in test solutions was based on the
`UV absorbance at 210 nm (creatine and creatinine) and 235 nm (creatinine only). The
`UV absorbance of creatinine at 235 nm allows for confirmation of creatinine concen-
`tration at both wavelengths. Calibration curves were generated using stock solutions
`(500 µg/ml) of creatine monohydrate and creatinine diluted in the mobile phase to
`concentrations of 3, 10, 30, and 50 µg/ml. For lower concentrations, calibration curves
`were prepared in mobile phase at concentrations of 0.1, 0.3, 1, and 3 µg/ml. Linear
`regressions gave excellent agreement between concentrations and detector response
`within the experimental concentration ranges (R2 = 0.999).
`
`Saturation Solubility Determination
`
`Using preliminary measurements as a guide, the saturated solubility of each creatine
`salt in deionized water was determined in triplicate by adding increasing amounts to 5
`◦
`C. After
`ml of solvent in screw-capped glass bottles placed in a shaking water bath at 25
`1.5 hr, the saturated solutions were vortexed and 2-ml aliquots removed and centrifuged
`in microcentrifuge tubes at 11,000 rpm for 5 min. Creatine concentrations were analyzed
`by HPLC by diluting 500 µL of supernatant with mobile phase (500 µL). The mean
`± standard deviation of the saturation solubility of each salt were calculated from the
`corresponding standard curves. Extended equilibration times for saturation were per-
`formed with no observable change in creatine concentration. Extending the equilibration
`time of creatine in these acidic saturated solutions did result in cyclization of creatine
`to creatinine as expected. Using the equilibration times outlined above, no significant
`creatinine concentration was observed in any of the saturation solubility experiments.
`
`Determination of the pH-Dependent Saturation Solubility of Creatine Monohydrate
`◦
`C in
`The pH-dependent solubility of creatine monohydrate was determined at 25
`triplicate. Excess creatine monohydrate was added to 20-ml glass vials containing
`10 ml of deionized water. The saturation solubilities of creatine monohydrate at pH
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`004
`
`Harvest Trading Group - Ex. 1109
`
`

`
`244
`
`JOURNAL OF DIETARY SUPPLEMENTS
`
`1.0, 2.0, 3.0, 4.0, 5.0, 6.0, 7.4, and 8.5 were measured; pH was adjusted by adding
`0.1-M HCl or 0.1-M NaOH. The suspensions were equilibrated for 2 hr in a shaking
`◦
`water bath at 25
`C and the final pH recorded. Aliquots were then removed and after
`centrifuging at 3,900 rpm for 10 min, the supernatant was collected, filtered through
`a 0.22-µm hydrophilic AladnTM syringe filter, and diluted in mobile phase and the
`creatine monohydrate concentration determined by HPLC analysis.
`
`Thermal Analysis
`
`Melting points of each creatine salt were obtained by differential scanning calorimetry
`(DSC) using a Shimadzu model DSC-50 with a TA-50WS controller. Samples (∼5 mg)
`◦
`C at a rate of
`were placed in flat aluminum pans, crimped, and heated from 30 to 330
`◦
`10
`C/min in an atmosphere of nitrogen (20 ml/min). To ensure that the true saturation
`solubility, rather than the apparent solubility, i.e., of free creatine, was being measured
`at each pH (Pudipeddi, Serajuddin, Grant, & Stahl, 2002), 3 hr after equilibration,
`1-ml aliquot was removed, lyophilized for 72 hr, and the thermogram of the remaining
`compound obtained and compared to that of the original salt.
`
`Water–Octanol Partition Coefficient Determination
`
`To assay the concentrations of each salt in both the aqueous and octanol phase after
`equilibrium, the creatine salt in the octanol phase was extracted in water as described
`below. Standard solutions of creatine monohydrate in mobile phase were prepared in
`the range of 0.1 to 50 µg/ml by serial dilution of a 500-µg/ml stock solution. Two series
`of standards (0.1, 0.3, 1.0, 3.0 µg/ml and 3.0, 10, 30, 50 µg/ml) were prepared. For the
`partitioning studies, 25 ml of each creatine salt was prepared in octanol–saturated water
`at a nominal creatine concentration of 6.0 mg/ml. Due to very low partition coefficients,
`this relatively high concentration was used to ensure that each salt could be quantified
`in the octanol phase. Triplicate aliquots (5 ml) of each creatine monohydrate or creatine
`salt solution were placed in glass scintillation vials and then an equal volume (5 ml)
`of water-saturated octanol was added. Triplicate 5-ml samples of aqueous solutions
`of each creatine derivative served as controls to test for decomposition. The samples
`◦
`C water bath for 2 hr and then the
`were allowed to equilibrate in a slowly shaking 25
`aqueous phases were assayed after appropriate dilutions with mobile phase. To assay
`the concentration of the creatine salt in the octanol phase, 4 ml of the water-saturated
`octanol phase was removed and placed in scintillation vials and 4 ml of water was added
`to extract the water-soluble salt. The samples were equilibrated for an additional 2 hr
`◦
`C to extract the creatine from the octanol into the
`in a slowly shaking water bath at 25
`aqueous phase. The concentration of each derivative so extracted was then quantified
`after diluting 500-µL aliquots in 500-µL mobile phase. Due to the very small amount of
`creatine partitioning into the octanol phase, larger quantities of creatine (100, 200, 300,
`400 µL of a 10-µg/mL stock solution) were also employed to more accurately assay the
`aqueous phase. The mean of the partition coefficient for each creatine compound was
`calculated and recorded as a log P value.
`
`Caco-2 Monolayer Permeability
`
`The permeability of various creatine salt forms was evaluated using the established
`human intestinal epithelial cell line derived from a human colon carcinoma (Caco-
`2). Although derived from colon epithelial cells, the Caco-2 cell line has many of
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`005
`
`Harvest Trading Group - Ex. 1109
`
`

`
`Gufford et al.
`
`245
`
`the characteristics of absorptive intestinal enterocytes found in small intestine and
`consequently has been used extensively for examining intestinal absorption of drugs
`and nutrients (Batrakova, Li, Miller, & Kabanov, 1999; Chabane, Al Ahmad, Peluso,
`Muller, & Ubeaud, 2009; Tian, Yan, Yang, & Wang, 2009). The Caco-2 cells were seeded
`onto collagen-coated Transwell polycarbonate membrane inserts (12-mm diameter;
`1-µm pore size) (Fisher Scientific, Winnipeg, MB) at a density of 60,000 cells/cm2. The
`cells were grown in DMEM supplemented with 10% fetal bovine serum and maintained
`in 5% CO2 environment. Cells were cultured for a period of 18–21 days, after which
`confluent monolayers were used for permeability studies.
`For the permeability studies, the culture media was removed from the cells and
`replaced with Tyrode’s buffer consisting of 136-mM NaCl, 2.6-mM KCl, 1.8-mM
`CaCl2, 1-mM MgCl2, 0.36-mM NaH2PO4, 5.56-mM D-Glucose, and 5-mM HEPES
`and pH 7.4. Following a 30-min pre-incubation period, Tyrode’s buffer was removed
`from the apical compartment and replaced with 0.5 ml of Tyrode’s buffer containing
`creatine monohydrate, creatine hydrochloride, creatine pyruvate, and creatine citrate
`(10 mM). Samples (50 µL) were removed from the apical (donor) compartment at the
`start and conclusion of the permeability experiment. Samples (100 µL) were removed
`from the receiver compartment at 0, 15, 30, 60, and 90 min. Passage of creatine from the
`donor to the receiver compartments was analyzed using HPLC methods described above.
`◦
`C. Permeability coefficients
`All permeability studies were performed at pH 7.4 and 37
`were determined based on the following equation:
`Papp= dCr/dt(Vd/(A
`where Cr is creatine concentration in the receiver compartment, t is time, Vd is the
`volume in the donor compartment, A is area, and Cd is the concentration in the donor
`compartment at time zero.
`
`Cd)),
`
`∗
`
`RESULTS AND DISCUSSION
`
`Creatine Salt Synthesis
`
`Zwitterionic creatine (pKa 2.79 and 12.1) (Prankerd, 2007) has the potential to form
`N-methylguanidinium salts directly only with acids that are more acidic than its car-
`boxylic acid functional group. Examples of such salts are the hydrochloride, mesylate,
`and hydrogen maleate that were successfully synthesized from hydrochloric acid (pKa
`−6), methane sulfonic acid (pKa −1.2), and maleic acid (pKa1 1.92) (Table 1). How-
`ever, using acids that are more acidic than the acidic carboxyl group of creatine does
`not necessarily guarantee that a salt will be formed as illustrated by our numerous
`attempts to synthesize creatine dihydrogen phosphate by mixing creatine monohydrate
`with phosphoric acid (pKa1 2.15) at various temperatures and cosolvent mixtures with
`only creatine monohydrate being recovered after each attempt. By monitoring freezing
`point depression, Dhar and Ghosh (1961) reported that a stable complex, but not nec-
`essarily a salt, is formed from mixtures of creatine and phosphoric acid. It is important
`to note that creatine citrate salts cannot be synthesized by simply mixing solutions of
`creatine and the insufficiently acidic citric acid (pKa 3.1, 4.8, 6.4). Indeed, using proton
`NMR, no chemical shifts were observed in the spectrum of the “creatine citrus” used in
`this study; the spectrum was simply a superposition of the spectra of creatine monohy-
`drate and citric acid confirming that this product was a physical mixture of creatine and
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`006
`
`Harvest Trading Group - Ex. 1109
`
`

`
`−3.5
`297.1
`8.6
`
`–
`
`1.0
`
`17.1±0.4
`195.0
`87.9
`149.1
`
`−3.2
`141.2
`3.1
`
`3.1/4.8/6.4
`
`3.0
`
`44.8±2.4
`
`194.0
`58.1
`451.1
`
`(Degussa)
`
`Nutrition)
`
`Monohydrate
`
`“Citrus”(Peak
`
`−3.3
`123.5
`2.4
`2.4
`
`4.8
`
`91.6±7.7
`
`191.3
`59.8
`219.2
`
`Pyruvate
`
`−3.8
`157.8
`2.2
`1.9
`
`2.2
`
`37.4±7.3
`201.0
`50.4
`260.2
`
`Maleate
`Hydrogen
`
`CreatineSalt
`
`−3.3
`179.8
`1.1
`−1.2
`29.6
`
`588±8
`191.0
`54.6
`240.3
`
`−3.2
`166.1
`0.3
`−6
`37.9
`
`709±7
`196.0
`78.2
`167.6
`
`−3.5
`255.0
`2.0
`−9.0
`7.1
`
`121±1
`195.0
`72.8
`180.2
`
`25
`Octanol–WaterLogPat
`MeltingPoint(◦C)
`pHatSaturation
`pKaofAcid
`CreatineMonohydrate
`RatioAqSolubilityw.r.t.
`(mg/ml)
`AqueousSolubilityat25◦C
`λmax(nm)
`WeightCreatine(%)
`MW(g/mole)
`
`C
`
`◦
`
`Mesylate
`
`Hydrochloride
`
`Hemisulfate
`
`Property
`
`TABLE1.Physicochemicalpropertiesofcreatinesalts
`
`246
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`007
`
`Harvest Trading Group - Ex. 1109
`
`

`
`Gufford et al.
`
`247
`
`citric acid in a 2:1 molar ratio, and not a true salt. As a further example, we attempted
`to synthesize creatine lipoate (Buononato & Festuccia, 2003) by mixing a 1:1 molar
`ratio of creatine monohydrate and lipoic acid (pKa 5.4) in ethanol at room temperature
`for 12 hr, but recovered only the two starting materials. In summary, without using an
`ion-exchange process, it is not possible to form creatine salts with acids that are less
`acidic than creatine itself.
`
`Physiochemical Properties
`
`Table 1 lists the properties of the salts in order of the pKa values (Prankerd, 2007)
`of the conjugate acid. Specifically, these properties are molecular weight, percentage
`◦
`C, ratio
`weight of creatine in each salt, UV absorption maxima, aqueous solubility at 25
`of the aqueous solubility compared to creatine monohydrate, saturated solution pH, mp,
`and logarithm of the octanol:water partition coefficient.
`
`Aqueous Solubility
`
`Each creatine salt (or complex in the form of “creatine citrus”) was significantly more
`soluble than creatine monohydrate. A contributing factor could be that the pH values of
`all the saturated salt solutions in unbuffered water were strongly acidic in the range of
`0.3 to 3.1, whereas the pH of the saturated creatine monohydrate solution was 8.6. At
`neutral pH, or in biological fluid, creatine exists in the least soluble, electrically neutral,
`zwitterionic form because it has an isoelectric point of 7.4. The hydrochloride and
`mesylate had notably high water solubilities; the former was nearly 40-fold more soluble
`than creatine monohydrate. Although the lack of a common ion effect for mesylate salts
`is a potential advantage over hydrochloride salts, the former have the potential liability
`of contamination with residual genotoxic alkyl mesylate ester impurities (Snodin, 2006).
`The creatine salts formed from carboxylic acids were less soluble than the creatine salts
`formed from methanesulfonic acid and the two mineral acids, although the hemisulfate
`was only 7.1-fold more soluble than creatine monohydrate.
`
`Thermal Analysis
`
`Melting points of creatine monohydrate and salts are listed in Table 1. In addition,
`thermograms of creatine monohydrate and salts are shown in Figure 2. Thermograms
`of creatine monohydrate and salts are also shown after lyophization of the saturation
`solubility tests to confirm that what was measured in solution was the salt and not molec-
`ular creatine or creatinine. With the exception of creatine monohydrate, no significant
`change in the DSC thermograms was observed for the creatine salts before and after
`lyophilization of the saturated solution. This indicates that the solubility measurements
`are true representations of the salt saturation solubilities and not just the solubility of
`free creatine at that particular pH, described as the apparent solubility (Pudipeddi et al.,
`2002). The expected difference observed in creatine monohydrate is the result of the
`◦
`C
`water loss during lyophilization, as indicated by the absence of the peak near 100
`(Dash, Mo, & Pyne, 2002). HPLC analysis further confirms these conclusions as neither
`the 210-nm nor 235-nm trace shows significant creatinine content in any of the measured
`test solutions.
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`008
`
`Harvest Trading Group - Ex. 1109
`
`

`
`248
`
`JOURNAL OF DIETARY SUPPLEMENTS
`
`FIGURE 2. DSC thermograms of (a) creatine monohydrate; (b) creatine mesylate; (c) cre-
`atine hydrogen maleate; (d) creatine hydrochloride; (e) creatine hemisulfate; (f) creatine
`pyruvate; and (g) “creatine citrus.” Solid lines indicate thermal profiles of original samples
`and dashed lines represent thermal profiles of saturated solutions post-lyophilization.
`
`Octanol–water Partition Coefficients
`
`As expected, the octanol–water partition coefficients for the creatine salts were very
`low and essentially equivalent reflecting their high polarity. Due to dissociation of
`salts in aqueous medium, the measured partition coefficients reflect the free creatine
`concentration, and hence, yield similar values (log P – 3.8 to – 3.2) for each creatine
`salt or zwitterionic monohydrate. Although the aqueous solubilities of the creatine salts
`and monohydrate varied considerably, these very low log P values indicate that all had
`very low affinity for the lipophilic octanol phase. It should be noted that because of
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`009
`
`Harvest Trading Group - Ex. 1109
`
`

`
`Gufford et al.
`
`249
`
`the highly hydrophilic nature of creatine, it was necessary to use the extraction method
`to measure the creatine concentrations in octanol phases rather than estimating octanol
`concentrations by differences in the water and control samples.
`
`pH-Dependent Solubility of Creatine Monohydrate
`
`Figure 3 illustrates the solubility of creatine monohydrate from pH 1.0 to 8.5. The
`solubility of creatine monohydrate at pH 1.0 and 4.0 was 52.0 ± 3.3 mg/ml and 15.7 ±
`0.3 mg/ml, respectively. At pH 1.0, the carboxylic acid functional group of creatine is
`largely unionized and creatine exists predominately as the monocation, whereas at pH
`4.0, the carboxylic acid group is largely ionized and the zwitterionic form dominates.
`In the pH range of 4.0 to 9.0, where the zwitterionic form also predominates, the
`solubility remained essentially constant. The pH-dependent saturation solubility profile
`of creatine monohydrate follows expected results with the lowest measured solubility
`being at pH 7.4, the isoelectric point of the creatine zwitterion. The creatine salts would
`be expected to yield comparable profiles when the same acid or base is used for pH
`
`FIGURE 3. pH-dependent solubility of creatine monohydrate.
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`010
`
`Harvest Trading Group - Ex. 1109
`
`

`
`250
`
`JOURNAL OF DIETARY SUPPLEMENTS
`
`adjustment (Pudipeddi et al., 2002). The pH-dependent solubilities of the creatine salts
`were not determined, as all saturated aqueous solutions are strongly acidic and buffer
`concentrations required to maintain a constant pH would make the data meaningless.
`
`Caco-2 Monolayer Permeability
`
`The permeability of several creatine salts was determined in confluent Caco-2 mono-
`layers (Figure 4). For each salt form examined, the flux of creatine across Caco-2
`monolayers was linear over the entire permeability study. While there was a tendency
`for creatine citrate to have lower permeability in Caco-2 monolayers, there were no sig-
`nificant differences in the permeability coefficients for any of the examined creatine salt
`forms (Figure 4). The fact that permeability of all the creatine salt forms was relatively
`low is certainly consistent with the octanol–water partitioning studies. The results of the
`present study are also in agreement with the relatively low permeability of radiolabeled
`creatine monohydrate in Caco-2 monolayers reported previously by Dash, Miller, Han,
`Carnazzo, and Stout (2001). Furthermore, as the permeability studies were performed at
`
`FIGURE 4. Caco-2 monolayer permeability. Creatine hydrochloride (CHCl), creatine mono-
`hydrate (CM), “creatine citrus” (CCit), and creatine pyruvate (CPyr).
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`011
`
`Harvest Trading Group - Ex. 1109
`
`

`
`Gufford et al.
`
`251
`
`a concentration where all the salt forms examined were soluble, significant differences
`in permeability between the various salt forms were not anticipated.
`Permeability in Caco-2 monolayers has been used to predict oral absorption of a
`variety of dietary supplements. For example, Mandagere, Thompson, and Hwang (2002)
`have established and validated a model using Caco-2 monolayer permeability and in
`vitro metabolic stability to predict the oral absorption of drugs. Based on this model,
`the creatine permeability observed in Caco-2 monolayers in the present study (i.e., Papp
`of 10–15 × 10
`−6 cm/s) would predict a 20–50% oral absorption of creatine. Given the
`relatively large doses of creatine dietary supplements that are required for biological
`effects (i.e., 5–20 gm daily dose), identification of creatine salts with improved aqueous
`solubility would be anticipated to improve oral absorption.
`
`CONCLUSIONS
`
`Two new salts, specifically creatine mesylate and creatine hydrogen maleate, were
`synthesized. The physicochemical properties of six creatine salts were determined and
`compared with creatine monohydrate. All salts were significantly more water-soluble
`than the monohydrate and all had very low partition coefficients reflecting their highly
`hydrophilic nature. Saturation solubility was dependent on both pH of the solution at
`saturation and the specific salt derivative. Caco-2 permeability studies further confirm
`the hydrophilic nature of these compounds and with the exception of creatine citrate,
`no significant differences in permeability were observed between the salts and creatine
`monohydrate. Although all salts are similar in permeability when studied at soluble
`concentrations, the large doses of less soluble, commercially available creatine mono-
`hydrate compounds typically used for human consumption result in administration of
`suspensions expected to have reduced overall absorption. Potential advantages of crea-
`tine salts over creatine monohydrate include: enhanced aqueous solubility resulting in
`increased bioavailability, reduced adverse effects, additional formulation options, and
`wider range of therapeutic applications. Increasing water solubility would facilitate the
`administration of creatine salts as a solution potentially creating less GI disturbance and
`improved bioavailability. The enhanced water solubility makes the preparation of com-
`mercial oral and topical dosage forms more attractive than comparable dosage forms
`using creatine monohydrate. It is anticipated that tablet formulations of creatine salts
`would improve dissolution kinetics and oral absorption compared to creatine monohy-
`drate. In addition, the differences in crystalline structure between individual salts may
`allow for direct compression into a tablet. The potential for alternative formulations of
`creatine salts may also allow for a wider range of therapeutic applications e.g., topical
`anti-inflammatory. It is acknowledged that more complete in vivo pharmacokinetic anal-
`ysis of these compounds would be required to fully elucidate their potential advantages
`and disadvantages over currently available creatine supplement formulations.
`
`REFERENCES
`Arnold MJ. Preparation of stable ketals of pyruvic acid and its acid derivatives which are useful as
`dietary supplements. US Patent No. 6177576 B1, 2001.
`Batrakova EV, Li S, Miller DW, Kabanov AV. Pluronic P85 increases permeability of a broad spectrum
`of drugs in polarized BBMEC and Caco-2 cell monolayers. Pharm Res. 1999;16:1366–1372.
`Blatt T, Breitenbach U, Mummert C, Raschke T, Scherner C, Schulz J, Staeb F. Active ingredient
`combinations of glucosyl glycerides and creatine and/or creatinine. Patent No. WO 2006122668A1,
`Germany: Beiersdorf A.-G., 2006.
`
`Downloaded by [Reprints Desk Inc] at 10:50 11 November 2015
`
`012
`
`Harvest Trading Group - Ex. 1109
`
`

`
`252
`
`JOURNAL OF DIETARY SUPPLEMENTS
`
`Boldt M. Preparation of creatine salts of dicarboxylic acids. US Patent 2004133040A1, San Corpo-
`ration, 2004.
`Buononato A, Festuccia A. Process for the preparation of creatine lipoate salt. Patent No. WO
`2003-IT316 20030523, Italy: Licrea S.R.L., 2003.
`Chabane NM, Al Ahmad A, Peluso J, Muller CD, Ubeaud G. Quercetin and naringenin transport
`across human intestinal Caco-2 cells. J Pharm Pharmacol. 2009;61:1473–1483.
`Cornelius DW, Haynes GL. Dicreatine malate with increased bioavailability. US Patent No.
`2004220263A1, Creative Compounds, LLC, 2004.
`Dash AK, Miller DW, Han H-Y, Carnazzo J, Stout JR. Evaluation of creatine transport using Caco-2
`monolayers as an in vitro model for intestinal absorption. J Pharm Sci. 2001;90:1593–1598.
`Dash AK, Mo Y, Pyne A. Solid-state properties of creatine monohydrate. J Pharm Sci.
`2002;91:708–718.
`Dash AK, Sawhney A. A simple LC method with UV detection for the analysis of creatine and
`creatinine and its application to several creatine formulations. Pharmaceutical Biomed Anal.
`2002;29:939–945.
`Dhar NR, Ghosh GP. Complex compounds of acid, base, and salts with nitrogenous and other organic
`substances. Proc Natl Acad Sci India 1961;31A:74–77.
`Gardiner PT. Food supplements and methods comprising lipoic acid and creatine. US Patent No.
`6,136,339, 2000.
`Gardiner PT, Heuer MA, Molino M. Creatine hydroxycitric acid salts and methods for their production
`and use as food supplements e

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