Next Article in Journal
Cast-In-Situ, Large-Sized Monolithic Silica Xerogel Prepared in Aqueous System
Previous Article in Journal
A Review on the Weight-Loss Effects of Oxidized Tea Polyphenols
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Validated 1H and 13C Nuclear Magnetic Resonance Methods for the Quantitative Determination of Glycerol in Drug Injections

1
Institute of Traditional Chinese Medicine Research, Tianjin University of Traditional Chinese Medicine, Tianjin 300193, China
2
State Key Laboratory of Bioactive Substance and Function of Natural Medicines, Institute of Materia Medica, Chinese Academy of Medical Sciences & Peking Union Medical College, Beijing 100006, China
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(5), 1177; https://doi.org/10.3390/molecules23051177
Submission received: 29 March 2018 / Revised: 5 May 2018 / Accepted: 7 May 2018 / Published: 15 May 2018
(This article belongs to the Section Analytical Chemistry)

Abstract

:
In the current study, we employed high-resolution proton and carbon nuclear magnetic resonance spectroscopy (1H and 13C NMR) for quantitative analysis of glycerol in drug injections without any complex pre-treatment or derivatization on samples. The established methods were validated with good specificity, linearity, accuracy, precision, stability, and repeatability. Our results revealed that the contents of glycerol were convenient to calculate directly via the integration ratios of peak areas with an internal standard in 1H NMR spectra, while the integration of peak heights were proper for 13C NMR in combination with an external calibration of glycerol. The developed methods were both successfully applied in drug injections. Quantitative NMR methods showed an extensive prospect for glycerol determination in various liquid samples.

Graphical Abstract

1. Introduction

Glycerol is a simple polyol compound with three hydroxyl groups that are responsible for its solubility in water, hygroscopic, and hyperosmolar nature. As a potent osmotic dehydrating agent, glycerol has been used in medicinal and pharmaceutical preparations with additional effects on stroke, head injury, brain edema and glaucoma to reduce elevated tissue pressure [1,2,3,4]. Hematuria was the only side effect reported in a small proportion of patients treated with glycerol, which was not explicitly recorded in all studies [5]. Although side effects of glycerol are infrequent and seem to be negligible [6], a large amount of it can produce an ethanol-like anesthesia effect, and lead to high glucose and triacylglycerol in blood [7]. Appropriate analytical methods are therefore needed in order to control the officially required specifications of glycerol.
Chemical titration after reaction with sodium periodate as oxidizing agent has been widely adopted in the pharmacopoeia of China, Japan, America, and Europe. However, any compound containing a couple of neighbor hydroxyls or phenolic hydroxyl is able to participate in a redox reaction, resulting in unsatisfactory specificity and accuracy of this method. Analytical chromatography hyphenated with different detections has also been developed for glycerol determination in drug injections. Physical restrictions of glycerol—such as lack of chromophores, high boiling point, and non-volatility—make it difficult to detect by using conventional gas chromatography (GC) and high performance liquid chromatography-ultraviolet (HPLC-UV) techniques, unless assisted with the process of pre-column derivatization [8,9]. Since every single analyte requires its individual elaborative and time-consuming sample derivatization, efficient alternative screening approaches would be preferable. HPLC equipped with an evaporating light scattering detection and amino column has been described as a useful alternative method to determine glycerol [10]; however, amino column is not durable and has a rapid decrease of column efficiency.
Nuclear magnetic resonance (NMR) spectroscopy is an essential analytical tool used to unambiguously identify known and novel compounds. It has an inherent advantage that the intensity of resonance signal is directly proportional to the number of nuclei [11], which provides a possibility to simultaneously qualify and quantify several molecules in natural samples [12] such as foods [13], plants or herbal remedies [14], and biofluids [15]. Additionally, NMR technique is non-destructive and does not require any complex sample pretreatment. To the extent of our knowledge, NMR method has not been explored for glycerol quantification in the literature. The aim of our study was to develop a rapid and accurate alternative technique to analyze glycerol for compliance with legal requirements. Hence, we investigated for the first time the use of proton and carbon nuclear magnetic resonance spectroscopy (1H and 13C NMR) for quantitative analyses of glycerol with some advantages in relation of these chemical and chromatographic methods.

2. Results and Discussion

2.1. Signal Assignments and the Specificity

Proton NMR spectra of homemade and injection samples in D2O showed the presence of methine proton at δH 3.77 (1H, m) and two methylene groups at δH 3.63 (2H, dd, J = 11.7, 6.5 Hz) and δH 3.54 (2H, dd, J = 11.7, 4.4 Hz), corresponding to proton signals of glycerol. Methyl signal at δH 0.00 and ethylene signal at δH 6.40 were assigned to the chemical shift reference TSP-d4 and the internal standard maleic acid, respectively (Table 1).
Enlarged spectra in range of δ 3.05–4.00 (Figure 1) indicated the proton signals of glycerol were partially affected by some signals of fructose in the glycerol fructose and sodium chloride injection (GFS) and of glucose in the Shenxiong glucose injection (SXG) injections. Several analogues, such as 1,2-propanediol and 1,3-propanediol, have similar chromatographic behaviors to glycerol, which make them difficult to separate. In 1H NMR spectra (Figure 2), the proton signals of 1,3-propanediol and glycerol were totally separated without any overlap, and only one of methylene signals of glycerol was overlapped with that of 1,2-propanediol in the region of δH 3.50–3.56.
Since the range of 13C spectral width is 20 times larger than that of 1H, the problem of overlap in 1H NMR spectrum is expected to be reduced in 13C NMR spectrum. We thereby recorded the carbon NMR spectra of the samples as well, in which the methine carbon signal at δC 73.1 and two methylene signals at δC 62.5 were assigned to glycerol and two ethylene signals at δC 132.6 and 170.8 were to maleic acid. Moreover, carbon signals of glycerol and the internal standard (maleic acid) were all separate with the signals of other ingredients in the homemade samples, injections, and the analogue mixtures (Figure 3 and Figure 4), which revealed a better specificity of 13C NMR spectroscopy than that of 1H NMR.

2.2. Options for Pulse Sequences

Due to the presence of a very strong solvent signal of H2O in drug injections and residual proton signal in deuterated H2O, we adopted a solvent-suppression pulse sequence (zgcppr) to acquire 1H NMR data with pre-saturation at the H2O/HDO solvent frequency before 90° hard pulse. For providing maximum intensity of proton signals, the length of 90° pulse is correspondingly calibrated for each sample, which might vary from sample to sample [16].
Pulse sequence with the effectiveness of solvent suppression can cause some loss of peaks close to the water peak. A particular advantage of 13C NMR experiment is the absence of water resonance, and hence, solvent suppression is no longer required. Other factors such as heteronuclear coupling and nuclear Overhauser effect (NOE) enhancement for 13C nuclei need to be considered [17]. To ameliorate these problems, we employed an inverse gated-decoupling pulse sequence (zgig) to record 13C NMR spectra. In this experiment, 1H decoupling is active during 13C acquisition, whereas it is switched off during the relaxation delay to suppress the NOE-effect, and thus the acquired spectrum can be integrated.

2.3. Options for Acquisition and Post-Processing Conditions

Most of the acquisition parameters are robust and can be varied within wide ranges. For instance, repetition time to acquire a single-scan spectrum depends on the longitudinal relaxation time T1 of interest signals. Theoretically, five times of the longest T1 are chosen to measure 99.3% of the equilibrium magnetization. When multiple scans are performed, it will take a long time to record spectrum. On the other hand, it has been shown [18] that for steady-state magnetization, the sensitivity is maximized by setting the repetition rate equal to 1.3T1. This leads to a signal-to-noise (S/N) ratio that is approximately 1.4 times greater than that obtained using a recycle time of 5T1 for a given period of data collection. Thus, several reports have also employed shorter relaxation time to yield comparable quantification results. The proton T1 values of glycerol and maleic acid were determined by inversion-recovery experiment, and the results were shown in Table 2. We subsequently set the relaxation delay on 2.35T1, at which point magnetization had recovered about 90% and the S/N ratio was also acceptable. Considering the time-consuming acquisition of 13C NMR spectra, an empirical value of 10 s was employed as the relaxation delay time in the experiments. To check the relative errors caused by these settings, glycerol samples with two known concentrations were measured and processed with different methods (Table 3).
The interest peaks of glycerol and internal standard were integrated by areas in 1H NMR spectra, and glycerol concentration was calculated through direct proportion to the integral area of internal standard or by an external calibration obtained in linearity examination. The relative errors of internal standard calculation method at two concentrations were 0.000 and 0.018, while those of external calibration calculation method were 0.073 and 0.037, indicating that the internal standard calculation method and peak area integration was proper to 1H NMR analysis.
Integrations based on peak area and peak height in combination with two calculation methods were employed to determine the content of glycerol in 13C NMR spectra. The results showed that the minimum relative errors at two concentrations were 0.004 and 0.014, revealing that peak height integration and external calibration calculation method were alternative for 13C NMR analysis.

2.4. Method Validation

Using the selected processing conditions, the quantitative NMR methods were validated in sequence of precision, repeatability, stability, accuracy, and linearity (Table 4). The precision was evaluated by six replicate measurements on the testing sample and the relative standard deviation (RSD) values of precision were found to be 0.36% and 0.40% for 1H and 13C NMR methods, respectively. The repeatability was assessed by analyzing six different solutions independently prepared as the testing sample and the RSD values were 0.55% and 1.48% for 1H and 13C NMR methods, respectively. The stability was evaluated by analyzing the same sample solution at an interval of every 2 h and the RSD values were 0.35% and 0.96% for 1H and 13C NMR methods, respectively. Recovery tests were performed to determine the accuracy of quantitative NMR method, and the results showed the average recoveries of glycerol in 1H and 13C NMR experiments were 95.8% and 101.8% with RSD values of 0.68% and 0.98%, respectively.
Six solutions of glycerol in different concentrations were prepared and analyzed in triplicate to determine the linearity of NMR methods. The linear regression equations (correlation coefficients) were y = 20.912x − 0.435 (r2 = 1.0000) and y = 0.1968x + 0.0147 (r2 = 0.9977) for the established 1H and 13C NMR methods, respectively (Figure 5). In addition, LOD and LOQ for glycerol were determined to be 0.015 and 0.045 mM for 1H NMR method, as well as 0.16 and 0.48 mM for 13C NMR method. These results indicated good precision, repeatability, stability, accuracy, and linearity of developed 1H and 13C NMR methods, along with lower LOD and LOQ of 1H NMR method than those of 13C NMR method.

2.5. NMR Quantification for Glycerol in Injection Samples

The contents of glycerol in homemade and injection samples were determined by the proposed NMR methods as well as sodium periodate titration (SPR) method. Each sample was analyzed in triplicate and the results were summarized in Figure 6. According to the known amounts of glycerol in homemade samples, the relative errors of three methods were calculated and the results suggested that chemical titration method was less accurate than NMR methods (Figure 6a).
The average glycerol contents in SXG, GFS, XZL, and EIE injections were estimated to be 13.82 ± 0.04 mg/mL, 102.41 ± 0.40 mg/mL, 133.09 ± 0.27 mg/mL, and 26.86 ± 0.21 mg/mL by employing 1H NMR spectra, and to be 14.10 ± 0.20 mg/mL, 94.86 ± 0.50 mg/mL, 133.54 ± 0.38 mg/mL, and 26.03 ± 0.51 mg/mL by 13C NMR spectra (Figure 6b). It was found that glycerol contents determined by the titration method were significantly higher (p < 0.05) than those measured by NMR techniques, which were far beyond the relative errors of SPR method. One possible explanation for the over-estimation of chemical titration results may be a result of having several reactable compounds in these injections. Periodate oxidation is a selective oxidation reaction that can act with the presence of O-dihydroxy or O-trihydroxy moiety in the molecular structures. The large amount of glucose in SXG and fructose in GFS probably caused a strong discrepancy between SRT and NMR values, while a relatively small amount of tannic acid in XZL and propylene glycol in EIE showed less influence on the discrepancy between these method results. Moreover, the values of glycerol content in GFS by 1H NMR method were found to be higher than those of 13C NMR method, which were probably associated with the overlapping signals of glycerol and fructose in the range of δ 3.45–4.05.

3. Materials and Methods

3.1. Reagents and Materials

Glycerol (99.5%), 3-(trimethylsilyl)propionic-2,2,3,3-d4 acid sodium salt (TSP-d4, 98 atom% D), maleic acid (99.94%), sodium l-lactate (99.0%), deuterium oxide (D2O, 99.9 atom% D), 1,2-propanediol (99.5%), and 1,3-propanediol (98%) were purchased from Sigma-Aldrich (Steinheim, Germany). Shenxiong glucose injection (SXG, batch no. 20150682) was from Guizhou Jingfeng Injection Co., Ltd. (Guizhou, China). Xiaozhiling injection (XZL, batch no. 15050505) was from Jilin Jian Yisheng Pharmaceutical Co., Ltd. (Jilin, China). Glycerol fructose and sodium chloride injection (GFS, batch no. 1504242) was from Nanjing Chia Tai Tianqing Pharmaceutical (Jiangsu, China). Etomidate injectable emulsion (EIE, batch no. 20150803) was from Jiangsu Nhwa Pharmaceutical Co., Ltd. (Jiangsu, China).
NMR I solution consisted of 0.59 mM TSP-d4 and 22.67 mM maleic acid in D2O, and NMR II solution consisted of 543.11 mM maleic acid in D2O.

3.2. NMR Measurement

NMR spectra were recorded on a 600 MHz Bruker AVIII HD spectrometer equipped with a 5 mm BBO H&F cryogenic probe. Standard one-dimensional composite pulse sequencing (zgcppr) was used to acquire 1H NMR spectra with the following instrumental settings: number of scans = 16; temperature = 298 K; relaxation delay = 16 s; pulse width = 11.5 µs; acquisition time = 1.7039 s; receiver gain = 28; spectral width = 9615.4 Hz; offset = 4125 Hz. 13C NMR spectra were acquired by the utilization of the inverse gated-decoupling pulse sequence (zgig) and the acquisition parameters were set as follows: number of scans = 32; temperature = 298 K; relaxation delay = 10 s; pulse width = 6.8250 µs; acquisition time = 3.9716 s; spectral width = 36,057.7 Hz; offset = 4125 Hz. All spectra were manually phased and automatically baseline corrected. Spin-lattice relaxation time (T1) values of protons in glycerol and maleic acid were measured using a classical inversion recovery pulse sequence with 10 relaxation delays (τ) ranging from 0.01 to 20 s.

3.3. Quantification

Resonance assignments were based on chemical shifts and spectral databases. Two integral ways, based on peak areas and peak heights, were adopted in conjunction with internal standard or external calibration. Appropriate processing procedures were picked by the relative errors of predicted values compared to the actual mass of glycerol in NMR I and II solutions. The linearity was evaluated by six various contents of glycerol in a range of 5.48–175.4 mM and 27.18–869.6 mM for 1H and 13C detected experiments, respectively. The limit of detection (LOD) and limit of quantification (LOQ) for glycerol were calculated based on the standard deviation of y-intercept of the regression line and the slope of the calibration curve [11]. Moderate amounts of 1,2-propanediol with glycerol and 1,3-propanediol with glycerol were dissolved in NMR I and II solutions, respectively, which were used to evaluate the specificities of 1H and 13C NMR methods.
Mixtures of glycerol and sodium lactate were homemade as testing samples for method validation with a glycerol concentration of 21.55 mM in 1H NMR experiment and 218.1 mM in 13C NMR experiment. The precision of two NMR methods was evaluated by continuously analyzing one testing sample for six times on the same day. One sample was analyzed to determine stability in 0, 2, 4, 6, 8, and 12 h on the same day. The repeatability was determined by analyzing six replicates of testing samples. The amounts of 0.55 mg and 5.05 mg glycerol were added into homemade samples for 1H and 13C NMR recovery experiments, respectively. The recovery was calculated by (glycerol mass found in glycerol-added testing samples − glycerol mass found in original testing samples)/glycerol mass added × 100%.
For glycerol determination by 1H NMR, 100 µL of drug injections were diluted by 2 mL of NMR I solution. For 13C NMR analysis, 100 µL of Xiaozhiling injection or glycerol fructose and sodium chloride injection was diluted by 300 µL of ultrapure water and 100 µL of NMR II solution. A volume of 400 µL Shenxiong glucose injection or etomidate injectable emulsion was diluted by 100 µL of NMR II solution. Each mixture was homogenized and 550 µL was transferred into 5 mm NMR tubes.

4. Conclusions

Based on quantitative NMR analysis, a reliable method for determination of glycerol in injection has been validated by using maleic acid as internal standard and D2O as the NMR solvent. The results of accuracy, linearity, precision, stability, and repeatability emphasize that 1H and 13C NMR can be used for quantitative determinations of glycerol in injections. Along with the continuous improvement and development of NMR technology, the established methods are probably an alternative for various applications of glycerol in solutions, such as pharmaceutical quality control, food additives, and biofuels.

Author Contributions

J.L. designed and performed the experiments, analyzed the data and prepared the manuscript; P.W. helped perform the experiments; Q.W. analyzed the data; Y.W. established the test conditions; M.J. helped prepare the manuscript and provided discussion.

Funding

This work was financially supported by the Natural Science Foundation of China (no. 81573547), Tianjin Technology Innovation System and Platform Development Plan (15PTCYSY00030) and CAMS Innovation Fund for Medical Sciences (2016-I2M-3-010).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bayer, A.J.; Pathy, M.S.J. Newcombe R Double-blind randomised trial of intravenous glycerol in acute stroke. Lancet 1987, 1, 405–408. [Google Scholar] [CrossRef]
  2. Yu, Y.L.; Kumana, C.R.; Lauder, I.J.; Cheung, Y.K.; Chan, F.L.; Kou, M.; Fong, K.Y.; Cheung, R.T.; Chang, C.M. Treatment of acute cortical infarct with intravenous glycerol. A double-blind, placebo-controlled randomized trial. Stroke 1993, 24, 1119–1124. [Google Scholar] [PubMed]
  3. Biestro, A.; Alberti, R.; Galli, R.; Cancela, M.; Soca, A.; Panzardo, H.; Borovich, B. Osmotherapy for increased intracranial pressure: Comparison between mannitol and glycerol. Acta Neurochir. 1997, 139, 725–733. [Google Scholar] [CrossRef] [PubMed]
  4. Casey, T.A.; Trevor-Roper, P.D. Oral glycerol in glaucoma. Br. Med. J. 1963, 2, 851–852. [Google Scholar] [CrossRef] [PubMed]
  5. Righetti, E.; Celani, M.G.; Cantisani, T.A.; Sterzi, R.; Boysen, G.; Ricci, S. Glycerol for acute stroke: A Cochrane systematic review. J. Neurol. 2002, 249, 445–451. [Google Scholar] [CrossRef] [PubMed]
  6. Frank, M.S.; Nahata, M.C.; Hilty, M.D. Glycerol: A review of its pharmacology, pharmacokinetics, adverse reactions, and clinical use. Pharmacotherapy 1981, 1, 147–160. [Google Scholar] [CrossRef] [PubMed]
  7. Cinar, Y.; Senyol, A.M.; Kosku, N.; Duman, K. Effects of glycerol on metabolism and hemodynamics: A pilot study. Curr. Ther. Res. 1999, 60, 435–445. [Google Scholar] [CrossRef]
  8. Plank, C.; Lorbeer, E. Simultaneous determination of glycerol, and mono-, di-, and triglycerides in vegetable oil methyl esters by capillary gas chromatography. J. Chromatogr. A 1995, 697, 461–468. [Google Scholar] [CrossRef]
  9. Reddy, S.R.; Titu, D.; Chadha, A. A novel method for monitoring the transesterification reaction of oil in biodiesel production by estimation of glycerol. J. Am. Oil Chem. Soc. 2010, 87, 747–754. [Google Scholar] [CrossRef]
  10. Li, R.; Wang, W.; Ma, X.; Chen, B. Determination on content of free glycerol in biodiesel by HPLC-ELSD. J. Instrum. Anal. 2011, 30, 1167–1170. [Google Scholar]
  11. Holzgrabe, U.; Diehl, B.; Wawer, I. NMR Spectroscopy in Pharmaceutical Analysis, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2008; pp. 233–315. ISBN 978-0-444-53173-5. [Google Scholar]
  12. Simmler, C.; Napolitano, J.G.; McAlpine, J.B.; Chen, S.N.; Pauli, G.F. Universal quantitative NMR analysis of complex natural samples. Curr. Opin. Biotechnol. 2014, 25, 51–59. [Google Scholar] [CrossRef] [PubMed]
  13. Hohmann, M.; Koospal, V.; Bauer-Christoph, C.; Christoph, N.; Wachter, H.; Diehl, B.; Holzgrabe, U. Quantitative 1H NMR analysis of egg yolk, alcohol, and total sugar content in egg liqueurs. J. Agric. Food Chem. 2015, 63, 4112–4119. [Google Scholar] [CrossRef] [PubMed]
  14. Jiang, M.; Wang, C.; Zhang, Y.; Feng, Y.; Wang, Y.; Zhu, Y. Sparse partial-least-squares discriminant analysis for different geographical origins of Salvia miltiorrhiza by 1H NMR-based metabolomics. Phytochem. Anal. 2014, 25, 50–58. [Google Scholar] [CrossRef] [PubMed]
  15. Mason, S.; Reinecke, C.J.; Solomons, R.; Wevers, R.A.; Engelke, U.F.H. 1H NMR spectral identification of medication in cerebrospinal fluid of pediatric meningitis. J. Pharm. Biomed. Anal. 2017, 143, 56–61. [Google Scholar] [CrossRef] [PubMed]
  16. Bharti, S.K.; Roy, R. Quantitative 1H NMR spectroscopy. TrAC Trends Anal. Chem. 2012, 35, 5–26. [Google Scholar] [CrossRef]
  17. Shaykhutdinov, R.A.; MacInnis, G.D.; Dowlatabadi, R.; Weljie, A.M.; Vogel, H.J. Quantitative analysis of metabolite concentrations in human urine samples using 13C{1H} NMR spectroscopy. Metabolomics 2009, 5, 307–317. [Google Scholar] [CrossRef]
  18. Waugh, J.S. Sensitivity in Fourier transform NMR spectroscopy of slowly relaxing systems. J. Mol. Spectrosc. 1970, 35, 298–305. [Google Scholar] [CrossRef]
Sample Availability: Samples of glycerol and injections are available from the authors.
Figure 1. Enlarged 1H NMR spectra from δ 3.05 to δ 4.00 of (a) homemade testing sample; (b) glycerol fructose and sodium chloride injection (GFS); (c) Shenxiong glucose injection (SXG); (d) Xiaozhiling injection (XZL); and (e) etomidate injectable emulsion (EIE).
Figure 1. Enlarged 1H NMR spectra from δ 3.05 to δ 4.00 of (a) homemade testing sample; (b) glycerol fructose and sodium chloride injection (GFS); (c) Shenxiong glucose injection (SXG); (d) Xiaozhiling injection (XZL); and (e) etomidate injectable emulsion (EIE).
Molecules 23 01177 g001
Figure 2. 1H NMR spectra of (a) glycerol and 1,2-propanediol mixture and (b) glycerol and 1,3-propanediol mixture.
Figure 2. 1H NMR spectra of (a) glycerol and 1,2-propanediol mixture and (b) glycerol and 1,3-propanediol mixture.
Molecules 23 01177 g002
Figure 3. Representative 13C NMR spectra of (a) homemade testing sample; (b) glycerol fructose and sodium chloride injection (GFS); (c) Shenxiong glucose injection (SXG); (d) Xiaozhiling injection (XZL); and (e) etomidate injectable emulsion (EIE).
Figure 3. Representative 13C NMR spectra of (a) homemade testing sample; (b) glycerol fructose and sodium chloride injection (GFS); (c) Shenxiong glucose injection (SXG); (d) Xiaozhiling injection (XZL); and (e) etomidate injectable emulsion (EIE).
Molecules 23 01177 g003
Figure 4. 13C NMR spectra of (a) glycerol and 1,2-propanediol mixture and (b) glycerol and 1,3-propanediol mixture.
Figure 4. 13C NMR spectra of (a) glycerol and 1,2-propanediol mixture and (b) glycerol and 1,3-propanediol mixture.
Molecules 23 01177 g004
Figure 5. The plots of calibration curves for glycerol quantification (a) by 1H NMR method and (b) by 13C NMR method.
Figure 5. The plots of calibration curves for glycerol quantification (a) by 1H NMR method and (b) by 13C NMR method.
Molecules 23 01177 g005
Figure 6. Clycerol contents determined by sodium periodate titration (SPR), 1H NMR (qH-NMR) and 13C NMR (qC-NMR) methods (a) in the homemade testing samples with glycerol at four concentrations (the relative errors were marks on the histograms); (b) in the injection samples, including glycerol fructose and sodium chloride injection (GFS), Shenxiong glucose injection (SXG), Xiaozhiling injection (XZL), and etomidate injectable emulsion (EIE). *** p < 0.001.
Figure 6. Clycerol contents determined by sodium periodate titration (SPR), 1H NMR (qH-NMR) and 13C NMR (qC-NMR) methods (a) in the homemade testing samples with glycerol at four concentrations (the relative errors were marks on the histograms); (b) in the injection samples, including glycerol fructose and sodium chloride injection (GFS), Shenxiong glucose injection (SXG), Xiaozhiling injection (XZL), and etomidate injectable emulsion (EIE). *** p < 0.001.
Molecules 23 01177 g006
Table 1. Resonance signal assignments of glycerol, maleic acid, 1,2-propanediol, and 1,3-propanediol.
Table 1. Resonance signal assignments of glycerol, maleic acid, 1,2-propanediol, and 1,3-propanediol.
CompoundStructureNumber1H13C
glycerol Molecules 23 01177 i0011, 33.63 (2H, dd),
3.54 (2H, dd)
62.5
23.77 (1H, m)73.1
1,2-propanediol Molecules 23 01177 i00213.52 (1H, dd), 3.44 (1H, dd)67.9
23.86 (1H, m)66.6
31.13 (3H, d)17.9
1,3-propanediol Molecules 23 01177 i0031, 33.67 (4H, t)58.6
21.78 (2H, m)33.7
maleic acidHOOCCH=CHCOOH–CH=CH–6.40 (s)132.6
–COOH 170.8
Table 2. Proton T1 values of glycerol and maleic acid (n = 3)
Table 2. Proton T1 values of glycerol and maleic acid (n = 3)
Compound (Chemical Shift)T1 (s)Mean ± SD (s)RSD (%)
glycerol (δ 3.52)2.292.29 ± 0.000.22
2.28
2.29
glycerol (δ 3.63)2.242.25 ± 0.000.42
2.25
2.26
glycerol (δ 3.75)4.674.86 ± 0.173.56
4.93
5.00
maleic acid (δ 6.40)6.776.78 ± 0.020.25
6.78
6.78
Table 3. Relative errors of glycerol amounts predicted by 1H and 13C NMR methods
Table 3. Relative errors of glycerol amounts predicted by 1H and 13C NMR methods
Detected AtomIntegral WayCalculated MethodActual Mass (mg)Testing Mass (mg)Absolute ErrorRelative Error
1Hpeak areainternal standard0.550.550.000.000
1.091.110.020.018
external calibration0.550.51−0.040.073
1.091.05−0.040.037
13Cpeak areainternal standard5.024.78−0.240.048
10.049.81−0.230.023
external calibration5.024.56−0.460.092
10.048.91−1.130.113
peak heightinternal standard5.024.79−0.230.046
10.0410.240.200.020
external calibration5.025.00−0.020.004
10.049.90−0.140.014
Table 4. Validation results for NMR techniques of glycerol (n = 6)
Table 4. Validation results for NMR techniques of glycerol (n = 6)
Detected Atom1H13C
integral waypeak areapeak height
calculated methodinternal standardexternal calibration
linear regression equationy = 20.912x − 0.4351y = 0.1968x + 0.0147
correlation coefficient (r2)1.00000.9977
standard deviation of y-intercept0.0940.087
LOD (mM)0.0150.16
LOQ (mM)0.0450.48
precision (RSD %)0.360.40
stability (RSD %)0.350.96
repeatability (RSD %)0.551.48
recovery rate (RSD %)95.8 (0.68%)101.8 (0.98%)

Share and Cite

MDPI and ACS Style

Lu, J.; Wang, P.; Wang, Q.; Wang, Y.; Jiang, M. Validated 1H and 13C Nuclear Magnetic Resonance Methods for the Quantitative Determination of Glycerol in Drug Injections. Molecules 2018, 23, 1177. https://doi.org/10.3390/molecules23051177

AMA Style

Lu J, Wang P, Wang Q, Wang Y, Jiang M. Validated 1H and 13C Nuclear Magnetic Resonance Methods for the Quantitative Determination of Glycerol in Drug Injections. Molecules. 2018; 23(5):1177. https://doi.org/10.3390/molecules23051177

Chicago/Turabian Style

Lu, Jiaxi, Pengli Wang, Qiuying Wang, Yanan Wang, and Miaomiao Jiang. 2018. "Validated 1H and 13C Nuclear Magnetic Resonance Methods for the Quantitative Determination of Glycerol in Drug Injections" Molecules 23, no. 5: 1177. https://doi.org/10.3390/molecules23051177

Article Metrics

Back to TopTop