Next Article in Journal
Gastrodin Ameliorates Cognitive Dysfunction in Vascular Dementia Rats by Suppressing Ferroptosis via the Regulation of the Nrf2/Keap1-GPx4 Signaling Pathway
Next Article in Special Issue
Application of Electrospun Water-Soluble Synthetic Polymers for Multifunctional Air Filters and Face Masks
Previous Article in Journal
Mechanistic Aspects for the Modulation of Enzyme Reactions on the DNA Scaffold
Previous Article in Special Issue
A Review of Stimuli-Responsive Smart Materials for Wearable Technology in Healthcare: Retrospective, Perspective, and Prospective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stabilization Effects Induced by Trehalose on Creatine Aqueous Solutions Investigated by Infrared Spectroscopy

by
Maria Teresa Caccamo
1,2,* and
Salvatore Magazù
1,2,*
1
Dipartimento di Scienze Matematiche e Informatiche, Scienze Fisiche e Scienze Della Terra, Università degli Studi di Messina, Viale Ferdinando Stagno D’Alcontres 31, 98166 Messina, Italy
2
Consorzio Interuniversitario Scienze Fisiche Applicate (CISFA), Viale Ferdinando Stagno D’Alcontres 31, 98166 Messina, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(19), 6310; https://doi.org/10.3390/molecules27196310
Submission received: 30 August 2022 / Revised: 13 September 2022 / Accepted: 20 September 2022 / Published: 24 September 2022
(This article belongs to the Special Issue Materials for Healthcare)

Abstract

:
Creatine is a very popular amino acid widely utilized in the sports world due to its functions mainly related to muscle building and increasing performance. The present work investigates the behavior of creatine aqueous solutions and of creatine aqueous in the presence of trehalose as a function of time changes by means of Infrared spectroscopy. Infrared spectra have been gathered and studied over time for both the full spectrum and the intramolecular OH-stretching region for the two mixtures. This latter region was studied more specifically using a cutting-edge technique called Spectral Distance (SD). From this analysis of the spectral features of the investigated samples, it emerges that trehalose has a significant stabilizing effect on creatine aqueous solutions.

1. Introduction

Creatine or methylguanidino-acetic acid, whose chemical formula is C4H9N3O2, is an amino acid present in foods of animal origin (e.g., meat and fish) that is produced within our body and is currently available in powder and pills, especially for its widespread employment in the sports world [1,2,3].
The first person to successfully extract this molecule from meat was the French chemist Michel Eugène Chevreul, who discovered it in 1832. Because of this, he gave it the name creatine, which is derived from the Greek word “kreas,” which means meat. [4,5]. A few years later, in 1847, its presence was also confirmed by the German chemist Justus von Liebig, who had also observed a close link between this compound and muscle activity [6,7]. Creatine is in fact distributed throughout the body, where about 95% is in the skeletal muscles and about 5% is in the brain. The human body is able to synthesize it starting from three amino acids called glycine, methionine, and arginine. Creatine is produced in the liver, pancreas, and kidneys and then it is sent primarily to the muscles, brain, and heart [8,9].
Creatine is an essential substance, mainly for providing energy during muscle contraction. In the cell, the oxidation of carbohydrates and fats produces ATP, adenosine triphosphate, a molecule used to provide the energy necessary for the various cellular processes with the formation of ADP, adenosine diphosphate, and AMP, adenosine monophosphate. The main role of creatine, in the form of phosphocreatine, is to donate a phosphate group to ADP by converting it back into ATP, which can be used in a new cycle of reactions [10,11,12].
The amount of ATP present in the cell is very low and is rapidly exhausted. Creatine intervention allows one to maintain vigorous muscle contraction for a longer period of time without having the cell use oxygen [13,14,15].
After many years of studies and research, creatine has been given considerable importance in cellular energy metabolism, to the point of being used commercially as a supplement for athletes to improve sports performance and increase muscle mass [16,17,18,19].
The use of creatine in dietary supplements has become increasingly popular for improving muscle strength and athletic performance in short-term sports that require high physical exertion, such as running, swimming, and track cycling on short courses. Thanks to the action of creatine on muscle mass, it is also used in bodybuilding, rowing, and weightlifting. However, the use of creatine in dietary supplements has different effects depending on the muscle mass, the amount of creatine already present, and the amount introduced into the diet [20,21,22,23].
Creatine is often used in the treatment of many diseases, including fibromyalgia, Huntington’s disease, multiple sclerosis, and congestive heart failure, and the effects of creatine in the treatment of other diseases are being studied. Furthermore, it appears that creatine can help protect the health of patients at risk of ischemic heart disease or stroke and ensure the good development of the fetus during pregnancy, and it has an antioxidant action [24,25].
Trehalose (α-D-glucopyranosyl-α-D-glucopyranoside or alpha, alpha-trehalose, or α-D-glucopyranosyl-α-D-glucopyranoside, dihydrate) is a sugar, and more precisely a disaccharide (or disaccharide). Unlike sucrose, composed of a molecule of glucose linked to a molecule of fructose, trehalose is composed of two molecules of glucose linked together by an α-bond, α-1,1 (or “1,1-α- glycosidic”) particularly stable. Trehalose has the same chemical formula, C12H22O11, as the other two disaccharides, i.e., maltose and sucrose but different geometrical structures. It is a sugar, almost odorless, composed of white or almost white crystals, with a sweet taste [26,27,28,29].
It begins with a funny drug called Trehala (a soft cocoon formed by insects of the genus Larinus, living on different species of Echinops in the Middle East), used to treat coughs and lung pathologies, and presented at the International Exhibition of Paris, in 1855 by François Della Sudda. In 1858, Marcellin Berthelot (1827–1907) isolated trehalose from this atypical drug and reported its physical and chemical properties. In 1876, Müntz established the fact that trehalose, present in the cocoon, has the same chemical structure as the sugar isolated from Claviceps purpurea by Mitscherlich [30].
In 1913, trehalose was discovered for the first time in the plant environment and, more specifically, in Selaginella lepidophylla. Trehalose will subsequently be found in mosses, green algae, ferns, and liverworts and in a very small number of angiosperms, characterized by their resistance to desiccation. Trehalose was occasionally detected in blooming plants, but the levels were often very small and were thought to be microbial or fungal in origin [31,32]. In the plant environment, genes coding for active trehalose-phosphate-synthetase and for trehalose-phosphatases were identified at the end of the 1990s in Arabidopsis thaliana, a Brassicassaceae corresponding to the attractive French name Arabette des dames. Osmolyte, protector against stress, carrier of carbon, trehalose often gives way in the plant environment to sucrose, giving it primacy (the latter is often 100 to 1000 times better represented than trehalose) [33].
In the 1970s, Clegg studied brine shrimp (Artemia salina) and showed the importance of trehalose in the resistance of encysted embryos to desiccation [34,35]. In a state of anhydrobiosis (a slow state of life linked to low water content), cysts can survive for several decades [36].
Trehalose presents physical–chemical properties different from the other homologous disaccharides, which make it used not only in the food industry but also in pharmaceuticals and cosmetics, as well as in “energy” drinks or foods for athletes, as an additive protective agent for baker’s yeasts, or as a rust inhibitor [37,38,39,40,41,42].
The present work focused on explaining the key role that trehalose plays in bioprotective and shelf-life mechanisms [43,44,45,46,47,48].
For this purpose, according to several researchers [49,50,51], the features of the solvent have a major influence on the dynamics of proteins, and as a result, water/bioprotectant molecule combinations have received much attention.
Within this context, the higher glass transition temperature (Tg) of trehalose and its mixtures with water, in comparison to the other disaccharides, is the only explanation for its superior bioprotectant effectiveness, according to Green and Angell [52]. The higher Tg values of the trehalose/H2O mixtures, in comparison to the Tg values of the other disaccharides/H2O mixtures at all concentration values, imply that at a given temperature, the glass transition for trehalose and its mixtures with water always occurs at a higher water content. This trait is crucial for bioprotection.
However, this explanation is not sufficient on its own because there are other related systems, such as dextran ((C6H10O5)x) [51], a linear polysaccharide with α(1–6) glycosidic linkages, which has a greater Tg value but does not exhibit a similar bioprotective effect.
On the other hand, the “water-replacement hypothesis” was developed by Crowe et al. [53] to explain the protective function of trehalose by trehalose’s direct hydrogen bonding with polar headgroups of the lipids, just as water does. In contrast, Crowe et al. hypothesized that a direct interaction between the sugars and the biostructures occurs.
Numerous experimental results from various spectroscopic examinations, together with certain computer studies, unmistakably show that disaccharides and trehalose have the greatest impact on the structural and dynamical features of water [44,45,46].
More specifically, the destruction of the tetrahedral coordination of pure water is demonstrated by neutron diffraction results, which demonstrate for all disaccharides a strong distortion of the peaks linked to the hydrogen-bonded network in the partial radial distribution functions.
The addition of trehalose, in contrast to the other disaccharides, breaks down the tetrahedral intermolecular network of water, which results in the formation of ice by a decrease in temperature, according to Raman scattering data. These findings demonstrate the distinct kosmotropic nature of disaccharides by demonstrating that the intensity of the contact between the disaccharide and the water molecule is greater than that between the water molecules. Additionally, ultrasonic velocity tests provide evidence that, in comparison to the other disaccharides, the trehalose–water system exhibits the highest solute–solvent interaction strength and hydration number across all concentration ranges.
In terms of dynamics, Quasi Elastic Neutron Scattering (QENS) findings on disaccharide solutions show that trehalose, in particular, has a significant impact on water dynamics when disaccharides are present. Additionally, trehalose has a stronger kinetic character in the Angell’s classification system than the other disaccharides, according to viscosity studies on trehalose, maltose, and sucrose aqueous solutions. The low-frequency dynamics of trehalose, maltose, and sucrose water mixtures across the glass transition were further studied using QENS and Inelastic Neutron Scattering (INS). The results of the experiments demonstrate that the trehalose/H2O mixture has a stronger character, as seen by the relaxational to vibrational contribution ratio [54].
The degree of fragility of glass-forming systems has also been given a new operational definition [55], and in this context, the stronger nature of the trehalose/H2O mixture suggests a better attitude toward the ability of maltose and sucrose/H2O mixtures to encapsulate biostructures in a more rigid matrix.
Finally, it is noteworthy to stress that trehalose increases the surface tension and stabilizes the aqueous solution of creatine, preventing evaporation [56,57].
More precisely, creatine aqueous solutions and creatine aqueous solutions in the presence of trehalose have been studied as a function of time by means of infrared spectroscopy [58].
The class of spectroscopy known as infrared absorption (IR) focuses on the electromagnetic spectrum’s infrared region. It can be used, similar to certain other spectroscopic techniques, to identify compounds, establish the composition of a sample, and analyze changes in a composition of a sample with temperature or over time. It is well known that the infrared region of the electromagnetic spectrum is split into three groups: near, mid, and far infrared, named in connection with the visible spectrum [59,60,61]. The far infrared, which is adjacent to the microwave zone and ranges from roughly 400 to 10 cm−1 (1000–30 μm), has low energy and can be used for rotational spectroscopy. It is possible to examine fundamental vibrations and related rovibrational structures in the mid-infrared, which has a wavelength range of roughly 4000 to 400 cm−1 (30–1.4 μm). Harmonic vibrations can be stimulated by the more intense near-infrared, which falls between 14,000 and 4000 cm−1 (1.4–0.8 μm). The names and classifications of these sub-regions are fundamentally conventions. They are not based on precise chemical or electromagnetic properties, tight divisions, or other criteria [62,63].
The fact that molecules have distinctive frequencies at which they rotate or vibrate in relation to distinct energy levels is taken advantage of by infrared spectroscopy (vibrational modes). The atomic masses, the related vibronic coupling, and the form of the potential energy surfaces all influence these resonant frequencies. A molecular vibrational mode must be connected to modifications in the permanent dipole in order for it to be active in the infrared. In particular, within the Born–Oppenheimer and harmonic approximations, the resonant frequencies are determined by the normal modes corresponding to the potential energy surface of the molecular electronic ground state when the molecular Hamiltonian, corresponding to the electronic ground state, can be approximated by a harmonic oscillator in the vicinity of the equilibrium molecular geometry [64,65]. However, the resonance frequencies may initially be connected to the atomic masses of termination and the strength of the bond. Consequently, the vibration frequency can be connected to a specific bond. Since complex compounds frequently contain bonds, it is possible for vibrations to couple together, resulting in infrared absorptions at distinctive frequencies that can be coupled to chemical groups. A beam of infrared light is sent through a sample to produce the sample’s infrared spectrum. The quantity of energy absorbed at each wavelength can be determined by analyzing the transmitted light. A monochromatic beam, a change in wavelength over time, or a Fourier transform instrument that measures all wave measurements simultaneously can all be used to achieve this. Thus, the spectra in transmittance or in absorbance can be created. These features can be analyzed to learn more about the sample’s molecular makeup and how it has changed over time [66,67,68,69].
The Spectral Distance (SD) is a novel method for comparing spectra and characterizing sample changes as a function of time, which is used in the current study, as far as data processing is concerned [70,71].

2. Materials and Experimental Set-Up

Creatine aqueous solutions and creatine aqueous solutions in the presence of trehalose were prepared starting from pure creatine powder and pure trehalose, both purchased from Aldrich-Chemie, and double distilled water. For the mixtures, the investigated concentration values correspond to 90% of creatine and 10% of double distilled water for the binary system and to 90% of creatine, 8% of double distilled water, and 2% of trehalose for the ternary system. Concerning the data acquisition, the Fourier transform IR spectroscopy spectra in attenuated total reflection mode (ATR), were collected on samples brought into contact with the detection device element (crystal) and were performed in air. More specifically, the measurements were carried out on a Bruker alpha spectrometer equipped with an ATR stage with a diamond crystal, controlled by the OpusLab v 7.5 software. Here, the infrared beam is reflected on the internal surface of the crystal and creates an evanescent wave. Part of the energy of this wave is then absorbed, the reflected radiation being sent back to the detector [72]. One of the major advantages of this technique is the possibility of analyzing samples without special preparation. The acquisitions were carried out by performing 32 scans between 4000 and 400 cm−1, with a resolution of 0.4 cm−1.

3. Methods

The Spectral Distance (SD) is mathematically represented by the following expression:
S D = I ( ω , t 0 ) I ( ω , t ) · Δ ω
where I ( ω , t 0 ) is the intensity at frequency ω and at time t = 0 while Δ ω is the instrumental resolution.
When applied to different spectra, it furnishes a unique parameter to estimate the similarity degree between spectra acquired under different conditions or referring to different systems. More specifically, in our case, the SD measures the affinity between the reference spectrum, which is the spectrum acquired at time t = 0, and the spectra collected at different times. Such evaluation has been performed for the spectra of the two investigated systems [73,74,75,76]. Such a method allows one to characterize the system changes in time.

4. Results and Discussion

Figure 1 reports the infrared spectra of creatine aqueous solutions, in the spectral range of 4000–400 cm−1, as a function of time in the range of 0 ÷ 1800 s.
Figure 2 reports the infrared spectra of creatine aqueous solutions in the presence of trehalose, in the spectral range of 4000–400 cm−1, as a function of time in the range of 0 ÷ 3000 s.

4.1. Infrared Spectra Integrated Area

In order to detect behavioral differences between the binary and the ternary investigated systems, the spectra integrated areas were calculated for the whole spectral range (i.e., 4000 to 400 cm−1). Figure 3a (on the left) reports the calculated integrated area values of the whole investigated spectra as a function of time for the creatine aqueous solution, while Figure 3b (on the right) shows the integrated area values of the whole investigated spectra versus time for the creatine aqueous solution in the presence of trehalose.
As it can be seen from the inspection of the figures, the calculated integrated area values fulfill a decaying exponential law reaching a plateau value in the long-time range. In particular, the decaying characteristic time, which is 267 s for the ternary system and 220 s for the binary system, and the plateau value, which is 190 for the ternary system and 100 for the binary system, in the presence of trehalose is higher with respect to the binary system. These results suggest that the presence of trehalose significantly stabilizes in time with the creatine solution. Furthermore, the black line in Figure 3b indicates the comparison of the same time between the binary system and the ternary system.

4.2. OH-Stretching Integrated Area

Since a significant contribution to the total integrated area is derived from the intramolecular OH-stretching region, in the following, we focused the attention on the spectral feature changes as a function of time in the restricted region 2800–3700 cm−1.
Figure 4 reports the intramolecular OH-stretching spectra at different times in the 2800–3700 cm−1 spectral range for the creatine aqueous solution (a) and for the creatine aqueous solution in the presence of trehalose (b).
What emerges from the comparison of the two data sets is that the intramolecular OH-stretching mode revealed a marked dependence on time for the creatine aqueous solution, while the addition of trehalose to the binary solution reduced such a dependence.
In order to detect quantitative differences between the binary and the ternary investigated systems, the spectra integrated areas have been calculated for the spectral range 2800–3700 cm−1. Figure 5a (on the left) reports the calculated integrated area values of the OH-stretching region, i.e., 2800–3700 cm−1, for all the investigated spectra as a function of time for the creatine aqueous solution; while Figure 5b (on the right) shows the integrated area values of the OH-stretching region versus time for the creatine aqueous solution in the presence of trehalose [42,44].
As it can be seen, the calculated integrated area values of the OH-stretching region fulfill a decaying exponential law reaching a plateau value in the long-time range. In particular, the decaying characteristic time, which is 270 s for the ternary system and 220 s for the binary system, and the plateau value, which is 190 for the ternary system and 100 for the binary system, in the presence of trehalose is higher with respect to the binary system. These results are in agreement with the data obtained considering the total spectral range (i.e., 4000 to 400 cm−1) and confirm that the presence of trehalose significantly stabilizes in time with the creatine solution [45,60]. In addition, in this case, the black line in Figure 5b specifies the comparison of the same time between the binary system and the ternary system.

4.3. Spectral Distance Analysis of Normalized Spectra

Figure 6 reports the normalized spectra relative to the intramolecular OH-stretching region, i.e., in the spectral range of 2800–3700 cm−1, for the creatine aqueous solution on the left and for the creatine aqueous solution in the presence of trehalose on the right.
In order to better clarify the role played by trehalose, in the following, we apply the Spectral Distance analysis, taking as the reference spectrum the one at time 0. Figure 7 shows the values of the Spectral Distance calculated, by means of Equation (1), for the creatine aqueous solution on the left and for the creatine aqueous solution in the presence of trehalose on the right [53,64].
As it can be seen, the SD values versus time fulfill an increasing sigmoid law reaching a plateau value in the long-time range only for both systems investigated. However, in the case of the creatine aqueous solution in the presence of trehalose, the incremental rate is much lower with respect to the binary system. This result is in agreement with the previous findings and confirms that the presence of trehalose stabilizes significantly in time with the creatine solution.

5. Conclusions

The goal of the current work is to employ infrared spectroscopy to investigate how creatine aqueous solutions and creatine aqueous solutions, in the presence of trehalose, behave as a function of time.
Both the whole infrared spectra and the intramolecular OH-stretching bands have been collected and examined as a function of time for the two investigated systems, and the latter spectral region was investigated by using the supposed Spectral Distance (SD) approach. Such a procedure, applied to different spectra, furnishes a unique parameter to estimate the similarity degree between spectra acquired under different conditions allowing one to characterize the system changes in time
According to the obtained experimental findings and the performed data analyses, trehalose has a substantial stabilizing impact on creatine aqueous solutions.

Author Contributions

Conceptualization, M.T.C. and S.M.; methodology, M.T.C. and S.M.; validation, M.T.C. and S.M.; formal analysis, M.T.C. and S.M.; investigation, M.T.C. and S.M.; rdata curation, M.T.C. and S.M.; writing—original draft preparation, M.T.C. and S.M.; writing—review and editing, M.T.C. and S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Balsom, P.D.; Söderlund, K.; Ekblom, B. Creatine in humans with special reference to creatine supplementation. Sports Med. 1994, 18, 268–280. [Google Scholar] [CrossRef] [PubMed]
  2. Cooper, R.; Naclerio, F.; Allgrove, J.; Jimenez, A. Creatine supplementation with specific view to exercise/sports performance: An update. J. Int. Soc. Sports Nutr. 2012, 9, 33. [Google Scholar] [CrossRef] [PubMed]
  3. Snow, R.J.; Murphy, R.M. Creatine and the creatine transporter: A review. Mol. Cell. Biochem. 2001, 224, 169–181. [Google Scholar] [CrossRef] [PubMed]
  4. Rawson, E.S.; Dolan, E.; Saunders, B.; Williams, M.E.; Gualano, B. Creatine supplementation in sport, exercise and health. In Dietary Supplementation in Sport and Exercise; Routledge: London, UK, 2019; pp. 141–164. [Google Scholar]
  5. Kramer, H.; Rosas, S.E.; Matsushita, K. Beef Tea, Vitality, Creatinine, and the Estimated GFR. Am. J. Kid. Dis. 2016, 67, 169–172. [Google Scholar] [CrossRef] [PubMed]
  6. Finlay, M.R. Quackery and cookery: Justus von Liebig’s extract of meat and the theory of nutrition in the Victorian age. Bull. His. Med. 1992, 66, 404–418. [Google Scholar]
  7. Wang, C.C.; Lin, S.C.; Hsu, S.C.; Yang, M.T.; Chan, K.H. Effects of Creatine Supplementation on Muscle Strength and Optimal Individual Post-Activation Potentiation Time of the Upper Body in Canoeists. Nutrients 2017, 27, 1169. [Google Scholar] [CrossRef]
  8. Wyss, M.; Kaddurah-Daouk, R. Creatine and creatinine metabolism. Physiol. Rev. 2000, 80, 1107–1213. [Google Scholar] [CrossRef]
  9. Kreider, R.B.; Kalman, D.S.; Antonio, J.; Ziegenfuss, T.N.; Wildman, R.; Collins, R.; Candow, D.G.; Kleiner, S.M.; Almada, A.L.; Lopez, H.L. International Society of Sports Nutrition position stand: Safety and efficacy of creatine supplementation in exercise, sport, and medicine. J. Int. Soc. Sports Nutr. 2017, 14, 1. [Google Scholar] [CrossRef]
  10. Kreider, R.B.; Melton, C.; Rasmussen, C.J.; Greenwood, M.; Lancaster, S.; Cantler, E.C.; Milnor, P.; Almada, A.L. Long-term creatine supplementation does not significantly affect clinical markers of health in athletes. Mol. Cell. Biochem. 2003, 244, 95–104. [Google Scholar] [CrossRef]
  11. Harris, R.C.; Söderlund, K.; Hultman, E. Elevation of creatine in resting and exercised muscle of normal subjects by creatine supplementation. Clin. Sci. 1992, 83, 367–374. [Google Scholar] [CrossRef]
  12. Bertin, M.; Pomponi, S.M.; Kokuhuta, C.; Iwasaki, N.; Suzuki, T.; Ellington, W.R. Origin of the genes for the isoforms of creatine kinase. Gene 2007, 392, 273–282. [Google Scholar] [CrossRef] [PubMed]
  13. Suzuki, T.; Mizuta, C.; Uda, K.; Ishida, K.; Mizuta, K.; Sona, S.; Compaan, D.M.; Ellington, R.W. Evolution and divergence of the genes for cytoplasmic, mitochondrial, and flagellar creatine kinases. J. Mol. Evol. 2004, 59, 218–226. [Google Scholar] [CrossRef] [PubMed]
  14. Sahlin, K.; Harris, R.C. The creatine kinase reaction: A simple reaction with functional complexity. Amino Acids 2011, 40, 1363–1367. [Google Scholar] [CrossRef] [PubMed]
  15. Harris, R. Creatine in health, medicine and sport: An introduction to a meeting held at Downing College, University of Cambridge. 2010. Amino Acids 2011, 40, 1267–1270. [Google Scholar] [CrossRef]
  16. Hultman, E.; Soderlund, K.; Timmons, J.A.; Cederblad, G.; Greenhaff, P.L. Muscle creatine loading in men. J. Appl. Phys. 1996, 81, 232–237. [Google Scholar] [CrossRef]
  17. Kreider, R.B.; Jung, Y.P. Creatine supplementation in exercise, sport, and medicine. J. Exerc. Nutr. Biochem. 2011, 15, 53–69. [Google Scholar] [CrossRef]
  18. Andres, R.H.; Ducray, A.D.; Schlattner, U.; Wallimann, T.; Widmer, H.R. Functions and effects of creatine in the central nervous system. Brain Res. Bull. 2008, 76, 329–343. [Google Scholar] [CrossRef]
  19. Racette, S.B. Creatine supplementation and athletic performance. J. Orthop. Sports Phys. Ther. 2003, 33, 615–621. [Google Scholar] [CrossRef]
  20. Juhn, M.S.; Tarnopolsky, M. Oral creatine supplementation and athletic performance: A critical review. Clin. J. Sport Med. 1998, 8, 286–297. [Google Scholar] [CrossRef]
  21. Brosnan, M.E.; Brosnan, J.T. The role of dietary creatine. Amino Acids 2016, 48, 1785–1791. [Google Scholar] [CrossRef]
  22. Brancaccio, P.; Maffulli, N.; Limongelli, F.M. Creatine kinase monitoring in sport medicine. Brit. Med. Bull. 2007, 81, 209–230. [Google Scholar] [CrossRef] [PubMed]
  23. Wax, B.; Kerksick, C.M.; Jagim, A.R.; Mayo, J.J.; Lyons, B.C.; Kreider, R.B. Creatine for Exercise and Sports Performance, with Recovery Considerations for Healthy Populations. Nutrients 2021, 13, 1915. [Google Scholar] [CrossRef] [PubMed]
  24. Clark, J.F.; Kemp, G.J.; Radda, G.K. The creatine kinase equilibrium, free [ADP] and myosin ATPase in vascular smooth muscle cross-bridges. J. Theor. Biol. 1995, 173, 207–211. [Google Scholar] [CrossRef]
  25. Ronner, P.; Friel, E.; Czerniawski, K.; Fränkle, S. Luminometric assays of ATP, phosphocreatine, and creatine for estimation of free ADP and free AMP. Anal. Biochem. 1999, 275, 208–216. [Google Scholar] [CrossRef]
  26. O’brien, J. Stability of trehalose, sucrose and glucose to nonenzymatic browning in model systems. J. Food Sci. 1996, 61, 679–682. [Google Scholar] [CrossRef]
  27. Babić, J.; Šubarić, D.; Milicevic, B.; Ačkar, D.; Kopjar, M.; Tiban, N.N. Influence of trehalose, glucose, fructose, and sucrose on gelatinisation and retrogradation of corn and tapioca starches. Czech. J. Food Sci. 2009, 27, 151. [Google Scholar] [CrossRef]
  28. Bieganski, R.M.; Fowler, A.; Morgan, J.R.; Toner, M. Stabilization of active recombinant retroviruses in an amorphous dry state with trehalose. Biotechnol. Prog. 1998, 14, 615–620. [Google Scholar] [CrossRef]
  29. Bordat, P.; Lerbret, A.; Demaret, J.P.; Affouard, F.; Descamps, M. Comparative study of trehalose, sucrose and maltose in water solutions by molecular modelling. Europhys. Lett. 2004, 65, 41. [Google Scholar] [CrossRef]
  30. Tillequin, F. Trehala, a meeting point between zoology, botany, chemistry, and biochemistry. Rev. Hist. Pharm. 2009, 57, 163–172. [Google Scholar] [CrossRef]
  31. Adams, R.P.; Kendall, E.; Kartha, K.K. Comparison of free sugars in growing and desiccated plants of Selaginella lepidophylla. Biochem. Syst. Ecol. 1990, 18, 107–110. [Google Scholar] [CrossRef]
  32. Pampurova, S.; Van Dijck, P. The desiccation tolerant secrets of Selaginella lepidophylla: What we have learned so far? Plant Physiol. Biochem. 2014, 80, 285–290. [Google Scholar] [CrossRef] [PubMed]
  33. Lunn, J.E.; Delorge, I.; Figueroa, C.M.; Van Dijck, P.; Stitt, M. Trehalose metabolism in plants. Plant J. 2014, 79, 544–567. [Google Scholar] [CrossRef] [PubMed]
  34. Clegg, J.S. Free glycerol in dormant cysts of the brine shrimp Artemia salina, and its disappearance during development. Biol. Bull. 1962, 123, 295–301. [Google Scholar] [CrossRef]
  35. Clegg, J.S. Metabolic studies of crytobiosis in encysted embryos of Artemia salina. Comp. Biochem. Phys. 1967, 20, 801–809. [Google Scholar] [CrossRef]
  36. Crowe, J.H. Trehalose and anhydrobiosis: The early work of J. S. Clegg. J. Exp. Biol. 2008, 211, 2899–2900. [Google Scholar] [CrossRef]
  37. Feofilova, E.P.; Usov, A.I.; Mysyakina, I.S.; Kochkina, G.A. Trehalose: Chemical structure, biological functions, and practical application. Microbiology 2014, 83, 184–194. [Google Scholar] [CrossRef]
  38. Higashiyama, T. Novel functions and applications of trehalose. Pure Appl. Chem. 2002, 74, 1263–1269. [Google Scholar] [CrossRef]
  39. Burek, M.; Waśkiewicz, S.; Wandzik, I. Trehalose–properties, biosynthesis and applications. Chemik 2015, 3, 9–10. [Google Scholar]
  40. Lombardo, D.; Calandra, P.; Caccamo, M.T.; Magazù, S.; Pasqua, L.; Kiselev, M.A. Interdisciplinary approaches to the study of biological membranes. AIMS Biophys. 2020, 7, 267–290. [Google Scholar] [CrossRef]
  41. Caccamo, M.T.; Cannuli, A. PEG Acoustic Levitation Treatment for Historic Wood Preservation investigated by means of FTIR spectroscopy. Curr. Chem. Biol. 2019, 13, 60–72. [Google Scholar] [CrossRef]
  42. Caccamo, M.T.; Magazù, S. Experimental Investigation on the Bioprotective Role of Trehalose on Glutamine Solutions by Infrared Spectroscopy. Materials 2022, 15, 4329. [Google Scholar] [CrossRef] [PubMed]
  43. Caccamo, M.T.; Mavilia, G.; Mavilia, L.; Lombardo, D.; Magazù, S. Self-assembly Processes in Hydrated Montmorillonite by FTIR Investigations. Materials 2020, 13, 1100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Migliardo, F.; Affouard, F.; Bordat, P.; Descamps, M.; Lerbret, A.; Magazù, S.; Ramirez-Cuesta, A.J.; Telling, M.F.T. A Combined Neutron Scattering and Simulation Study on Bioprotectant Systems. Chem. Phys. 2005, 317, 258–266. [Google Scholar]
  45. Branca, C.; Magazù, S.; Maisano, G.; Migliardo, F. Vibrational and Relaxational Contributions in Disaccharide/H2O Glass Formers. Phy. Rev. B. 2001, 64, 224204. [Google Scholar] [CrossRef]
  46. Magazù, S.; Maisano, G.; Migliardo, P.; Villari, V. Experimental Simulation of Macromolecules in Trehalose Aqueous solutions: A Photon Correlation Spectroscopy Study. J. Chem. Phys. 1999, 111, 9086–9092. [Google Scholar] [CrossRef]
  47. Fenimore, P.W.; Frauenfelder, H.; Magazù, S.; McMahon, B.H.; Mezei, F.; Migliardo, F.; Young, R.D.; Stroe, I. Concepts and problems in protein dynamics. Chem. Phys. 2013, 424, 2–6. [Google Scholar] [CrossRef]
  48. Cannuli, A.; Caccamo, M.T.; Castorina, G.; Colombo, F.; Magazù, S. Laser Techniques on Acoustically Levitated Droplets. EPJ Web Conf. 2018, 167, 05010. [Google Scholar] [CrossRef]
  49. Frauenfelder, H.; Sligar, S.G.; Wolynes, P.G. The energy landscapes and motions of proteins. Science 1991, 254, 1598–1603. [Google Scholar] [CrossRef]
  50. Frauenfelder, H.; McMahon, B. Dynamics and function of proteins: The search for general concepts. Proc. Natl. Acad. Sci. USA 1998, 9995, 4795–4797. [Google Scholar] [CrossRef]
  51. Oliver, A.E.; Crowe, L.M.; Crowe, J.H. Methods for dehydration-tolerance: Depression of the phase transition temperature in dry membranes and carbohydrate vitrification. Seed Sci. Res. 1998, 8, 211–221. [Google Scholar] [CrossRef]
  52. Green, J.L.; Angell, C.A. Phase relations and vitrification in saccharide-water solutions and the trehalose anomaly. J. Phys. Chem. B 1989, 93, 2880–2882. [Google Scholar] [CrossRef]
  53. Crowe, J.H.; Cooper, A.F., Jr. Cryptobiosis. Sci. Am. 1971, 225, 30–36. [Google Scholar] [CrossRef]
  54. Maisano, G.; Majolino, D.; Migliardo, P.; Venuto, S.; Aliotta, F.; Magazú, S. Sound velocity and hydration phenomena in aqueous polymeric solutions. Mol. Phys. 1993, 78, 421–435. [Google Scholar] [CrossRef]
  55. Magazù, S.; Maisano, G.; Migliardo, F.; Mondelli, C. Mean-Square Displacement Relationship in Bioprotectant Systems by Elastic Neutron Scattering. Biophys. J. 2004, 86, 3241–3249. [Google Scholar] [CrossRef]
  56. Lin, T.Y.; Timasheff, S.N. On the role of surface tension in the stabilization of globular proteins. Protein Sci. 1996, 5, 372–381. [Google Scholar] [CrossRef] [PubMed]
  57. Olsson, C.; Zangana, R.; Swenson, J. Stabilization of proteins embedded in sugars and water as studied by dielectric spectroscopy. Phys. Chem. Chem. Phys. 2020, 22, 21197–21207. [Google Scholar] [CrossRef]
  58. Magazù, S.; Calabrò, E.; Caccamo, M.T. Experimental study of thermal restraint in bio-protectant disaccharides by FTIR spectroscopy. Open Biotechnol. 2018, 12, 123–133. [Google Scholar] [CrossRef]
  59. Stuart, B.H. Infrared Spectroscopy: Fundamentals and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2004. [Google Scholar]
  60. Migliardo, F.; Magazù, S.; Caccamo, M.T. Infrared, Raman and INS Studies of Poly-Ethylene Oxide Oligomers. J. Mol. Struc. 2013, 1048, 261–266. [Google Scholar] [CrossRef]
  61. Ferrari, M.; Mottola, L.; Quaresima, V. Principles, techniques, and limitations of near infrared spectroscopy. Can. J. Appl. Phys. 2004, 29, 463–487. [Google Scholar] [CrossRef]
  62. Caccamo, M.T.; Gugliandolo, C.; Zammuto, V.; Magazù, S. Thermal properties of an exopolysaccharide produced by a marine thermotolerant Bacillus licheniformis by ATR-FTIR spectroscopy. Int. J. Biol. Macromol. 2020, 145, 77–83. [Google Scholar] [CrossRef]
  63. Kazarian, S.G.; Chan, K.A. ATR-FTIR spectroscopic imaging: Recent advances and applications to biological systems. Analyst 2013, 138, 1940–1951. [Google Scholar] [CrossRef] [PubMed]
  64. Ahmadi, H.; Vafaee, M.; Maghari, A. Understanding molecular harmonic emission at relatively long intense laser pulses: Beyond the Born-Oppenheimer approximation. Phys. Rev. A 2016, 94, 033415. [Google Scholar] [CrossRef]
  65. Kreibich, T.; Lein, M.; Engel, V.; Gross, E.K.U. Even-harmonic generation due to beyond-born-oppenheimer dynamics. Phys. Rev. Lett. 2001, 87, 103901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Caccamo, M.T.; Magazù, S. Tagging the oligomer-to-polymer crossover on EG and PEGs by infrared and Raman spectroscopies and by wavelet cross-correlation spectral analysis. Vib. Spectr. 2016, 85, 222–227. [Google Scholar] [CrossRef]
  67. Ozaki, Y. Infrared Spectroscopy—Mid-infrared, Near-infrared, and Far-infrared/Terahertz Spectroscopy. Anal. Sci. 2021, 37, 1193–1212. [Google Scholar] [CrossRef] [PubMed]
  68. Caccamo, M.T.; Magazù, S. Ethylene Glycol-Polyethylene Glycol (EG-PEG) Mixtures: IR Spectra Wavelet Cross-Correlation Analysis. Appl. Spectr. 2017, 71, 401–409. [Google Scholar] [CrossRef]
  69. Caccamo, M.T.; Magazù, S. Multiscaling Wavelet Analysis of Infrared and Raman Data on Polyethylene Glycol 1000 Aqueous Solutions. Spectr. Lett. 2017, 50, 130–136. [Google Scholar] [CrossRef]
  70. Caccamo, M.T.; Magazù, S. A Conic Pendulum of Variable Length Analysed by wavelets. In New Trends in Physics Education Research; Nova Science Publishers, Inc.: New York, NY, USA, 2018; pp. 117–131. ISBN 978-1-53613-893-1. [Google Scholar]
  71. Posada, H.; Ferrand, M.; Davrieux, F.; Lashermes, P.; Bertrand, B. Stability across environments of the coffee variety near infrared spectral signature. Heredity 2009, 102, 113–119. [Google Scholar] [CrossRef]
  72. Griffiths, P.; Haseth, J. Fourier Transform Infrared Spectrometry, 2nd ed.; Wiley: Hoboken, NJ, USA, 2006. [Google Scholar]
  73. Caccamo, M.T.; Magazù, S. Thermal restraint on PEG-EG mixtures by FTIR investigations and wavelet cross-correlation analysis. Pol. Test. 2017, 62, 311–318. [Google Scholar] [CrossRef]
  74. Yang, Z.; Xiao, H.; Sui, Q.; Zhang, L.; Jia, L.; Jiang, M.; Zhang, F. Novel methodology to improve the accuracy of oxide determination in cement raw meal by near infrared spectroscopy (NIRS) and cross-validation-absolute-deviation-F-test (CVADF). Anal. Lett. 2020, 53, 2734–2747. [Google Scholar] [CrossRef]
  75. Caccamo, M.T.; Calabró, E.; Cannuli, A.; Magazù, S. Wavelet Study of Meteorological Data Collected by Arduino-Weather Station: Impact on Solar Energy Collection Technology. MATEC Web Conf. 2016, 55, 02004. [Google Scholar] [CrossRef]
  76. Shi, Z.; Cogdill, R.P.; Short, S.M.; Anderson, C.A. Process characterization of powder blending by near-infrared spectroscopy: Blend end-points and beyond. J. Pharm. Biomed. Anal. 2008, 47, 738–745. [Google Scholar] [CrossRef] [PubMed]
Figure 1. IR Spectra of creatine aqueous solutions as a function of time (range: 0 ÷ 1800 s) in the spectral range 4000–400 cm−1.
Figure 1. IR Spectra of creatine aqueous solutions as a function of time (range: 0 ÷ 1800 s) in the spectral range 4000–400 cm−1.
Molecules 27 06310 g001
Figure 2. IR Spectra of creatine aqueous solutions in the presence of trehalose as a function of time (0 ÷ 3000 s) in the spectral range of 4000–400 cm−1.
Figure 2. IR Spectra of creatine aqueous solutions in the presence of trehalose as a function of time (0 ÷ 3000 s) in the spectral range of 4000–400 cm−1.
Molecules 27 06310 g002
Figure 3. Integrated area values of the whole investigated spectra as a function of time for: (a) creatine aqueous solution; (b) creatine aqueous solution in the presence of trehalose.
Figure 3. Integrated area values of the whole investigated spectra as a function of time for: (a) creatine aqueous solution; (b) creatine aqueous solution in the presence of trehalose.
Molecules 27 06310 g003
Figure 4. Infrared spectra at different times in the 2800–3700 cm−1 spectral range, corresponding to the intramolecular OH-stretching region, for the creatine aqueous solution (a) and for the creatine aqueous solution in the presence of trehalose (b).
Figure 4. Infrared spectra at different times in the 2800–3700 cm−1 spectral range, corresponding to the intramolecular OH-stretching region, for the creatine aqueous solution (a) and for the creatine aqueous solution in the presence of trehalose (b).
Molecules 27 06310 g004
Figure 5. Integrated area values of the OH-stretching region, i.e., 2800–3700 cm−1, for all the investigated spectra as a function of time in the spectral range 2800–3700 cm−1 for: (a) creatine aqueous solution; (b) creatine aqueous solution in the presence of trehalose.
Figure 5. Integrated area values of the OH-stretching region, i.e., 2800–3700 cm−1, for all the investigated spectra as a function of time in the spectral range 2800–3700 cm−1 for: (a) creatine aqueous solution; (b) creatine aqueous solution in the presence of trehalose.
Molecules 27 06310 g005
Figure 6. Normalized spectra in the intramolecular OH-stretching region, i.e., in the spectral range of 2800–3700 cm−1, for the creatine aqueous solution (a) and for the creatine aqueous solution in the presence of trehalose on (b).
Figure 6. Normalized spectra in the intramolecular OH-stretching region, i.e., in the spectral range of 2800–3700 cm−1, for the creatine aqueous solution (a) and for the creatine aqueous solution in the presence of trehalose on (b).
Molecules 27 06310 g006
Figure 7. SD calculated values versus time for the creatine aqueous solution on the (left) and for the creatine aqueous solution in the presence of trehalose on the (right).
Figure 7. SD calculated values versus time for the creatine aqueous solution on the (left) and for the creatine aqueous solution in the presence of trehalose on the (right).
Molecules 27 06310 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Caccamo, M.T.; Magazù, S. Stabilization Effects Induced by Trehalose on Creatine Aqueous Solutions Investigated by Infrared Spectroscopy. Molecules 2022, 27, 6310. https://doi.org/10.3390/molecules27196310

AMA Style

Caccamo MT, Magazù S. Stabilization Effects Induced by Trehalose on Creatine Aqueous Solutions Investigated by Infrared Spectroscopy. Molecules. 2022; 27(19):6310. https://doi.org/10.3390/molecules27196310

Chicago/Turabian Style

Caccamo, Maria Teresa, and Salvatore Magazù. 2022. "Stabilization Effects Induced by Trehalose on Creatine Aqueous Solutions Investigated by Infrared Spectroscopy" Molecules 27, no. 19: 6310. https://doi.org/10.3390/molecules27196310

Article Metrics

Back to TopTop