You are on page 1of 5

2018

NPC Natural Product Communications Vol. 13


No. 7
Identification of the Chemical Constituents in Ginger 869 - 873
(Zingiber officinale) Responsible for Thermogenesis
Yuto Nishidonoa, Azis Saifudinb, Mikio Nishizawac, Takashi Fujitaa, Masatoshi Nakamotoa and Ken Tanakaa,*
a
College of Pharmaceutical Science, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
b
Faculty of Pharmacy, Universitas Muhammadiyah Surakarta, Pabelan, KTS Solo, Jawa Tengah 57102,
Indonesia
c
Graduate School of Life Sciences, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan

ktanaka@fc.ritsumei.ac.jp

Received: April 20th, 2018; Accepted: May 2nd, 2018

To compare the thermogenic properties of crude drugs derived from ginger, the activities to peroxisome proliferator-activated receptor gamma coactivator 1-
alpha (PGC-1α) of methanol extracts of “Shokyo” (dried rhizome of Z. officinale var. rubens), “Kankyo” (steamed and dried rhizome of Z. officinale var.
rubens), “Red ginger” (Indonesian dried rhizome of Z. officinale var. rubrum) and “White ginger” (Indonesian dried rhizome of Z. officinale var. amarum),
were examined. The extracts of the four specimens were analyzed by liquid chromatography mass spectrometry (LC-MS). The results showed that “Shokyo”
and “White ginger” strongly stimulated PGC-1α and that the amount of [10]-shogaol (6) in these was higher than in “Kankyo” and “Red ginger”. Gingerol-
related compounds were isolated or prepared in order to identify the compounds responsible for stimulating PGC-1α. As a result, [10]-gingerol (3), [10]-
shogaol (6), [10]-gingerdiols (11, 12) and [10]-gingerdiols 3,5-diacetate (17, 18) were identified as the active constituents, while the main constituents, [6]-
gingerol (1) and [6]-shogaol (4), did not show any significant PGC-1α activity. These results suggest that gingerol-related compounds with long alkyl side
chains contribute to the thermogenic properties of ginger.

Keywords: Zingiber officinale, PGC-1α, Thermogenesis, [10]-Shogaol.

Zingiber officinale is a perennial plant of the Zingiberaceae family have been reported [7], but studies on the effects of ginger and/or its
and is widely distributed throughout the tropical and subtropical chemical constituents on PGC-1α activity have not yet been
regions of Asia, Far East Asia and Africa. Its rhizomes, commonly reported.
known as ginger, have a pungent flavor and it has been used as a
spice for at least 2000 years [1a-1c]. It has also been used as In this study, we evaluated the effects of the methanol extracts of
medicine for the treatment of catarrh, rheumatism, nervous diseases, four crude drugs, derived from “Shokyo” (dried rhizome of Z.
gingivitis, toothache, asthma, stroke, constipation and diabetes in officinale var. rubens), “Kankyo” (steamed and dried rhizome of Z.
the traditional medicines of many countries [2a]. officinale var. rubens), “Red ginger” (Indonesian dried rhizome of
Z. officinale var. rubrum) and “White ginger” (Indonesian dried
Diarylheptanoids, sesquiterpenes and phenolic compounds have rhizome of Z. officinale var. amarum), on the PGC-1α activity. The
been identified as the major constituents in ginger [2a-2d]. There four extracts were analyzed by LC-MS and the chemical
have been several studies indicating that gingerols ([6]-, [8]- and constituents identified. The gingerol-related compounds (Figure 1)
[10]-gingerol) and shogaols ([6]-, [8]- and [10]-shogaol) have a were isolated or prepared to identify the compounds responsible for
number of different beneficial effects. These include thermogenesis.
anticarcinogenic, antioxidative, antimicrobial, anti-inflammatory,
antidiabetic, antiobesity, anti-gastric ulcer and antiallergenic effects. The PGC-1α activity of the methanol extracts of “Shokyo”,
In addition, it has also been reported that the bioactivity of shogaols “Kankyo”, “Red ginger” and “White ginger” are shown in Figure 2.
is stronger than that of gingerols [2a, 2e]. “Shokyo” and “White ginger” strongly stimulated the PGC-1α,
while no activity was apparent for “Kankyo” and “Red ginger”.
In many traditional medicines, ginger is used to warm parts of the Several studies done so far indicate that the major constituents of
body and release the exterior. Several studies on the thermogenic ginger, [6]-gingerol and [6]-shogaol, are responsible for the various
properties of ginger have been reported. Iwami and Fujiwara effects. In addition, it has also been reported that [6]-shogaol shows
reported that ginger increased the oxygen consumption in rats and stronger bioactivity than [6]-gingerol [2a]. [6]-shogaol is the
raised the surface temperature of the forehead in humans [3a, 3b]. dehydrated product of [6]-gingerol, and [6]-gingerol is easily
It is believed that thermogenesis induced by the administration of transformed to [6]-shogaol during the heating process [8].
ginger is caused by activation of the transient receptor potential Therefore, it is suggested that “Kankyo” has stronger thermogenic
vanilloid subtype 1 (TRPV1) channel, and that [6]-gingerol and [6]- properties than “Shokyo”. However, the present results were not
shogaol are responsible for activating this [4a-4c]. TRPV1 consistent with previous findings.
activation improves energy metabolism by upregulating the
peroxisome proliferator-activated receptor gamma coactivator 1- To identify the active constituents, the chemical constituents of the
alpha (PGC-1α) in skeletal muscles [5]. PGC-1α is a coactivator of four methanol extracts were analyzed by LC-MS. Figure 3 shows
the peroxisome proliferator-activated receptor γ (PPARγ) and plays total ion chromatograms (TIC) with both positive and negative ion
an important role in thermogenesis through the uncoupling of scans and mass chromatograms to monitor the [M˗H]˗ ions of the
oxidative phosphorylation [6a-6c]. The effects of ginger on PPARγ gingerols (1-3), shogaols (4-6), and gingerdiols (7-12), as well the
870 Natural Product Communications Vol. 13 (7) 2018 Nishidono et al.

(3S,5S)-[6]-Gingerdiol (7) : R=H


[6]-Gingerdiol 3S,5S-diacetate (13) : R=Ac
[6]-Gingerol (1)

(3R,5S)-[6]-Gingerdiol (8) : R=H


[6]-Gingerdiol 3R,5S-diacetate (14) : R=Ac
[8]-Gingerol (2)

(3S,5S)-[8]-Gingerdiol (9) : R=H


[8]-Gingerdiol 3S,5S-diacetate (15) : R=Ac
[10]-Gingerol (3)

(3R,5S)-[8]-Gingerdiol (10) : R=H


[8]-Gingerdiol 3R,5S-diacetate (16) : R=Ac
[6]-Shogaol (4)

(3S,5S)-[10]-Gingerdiol (11) : R=H


[10]-Gingerdiol 3S,5S-diacetate (17) : R=Ac
[8]-Shogaol (5)

(3R,5S)-[10]-Gingerdiol (12) : R=H


[10]-Gingerdiol 3R,5S-diacetate (18) : R=Ac
[10]-Shogaol (6)
Figure 1: Structures of gingerol-related compounds in ginger.

3 study, a systematic analysis of gingerol-related compounds (1-18)


was performed. Regarding the PGC-1α activity, “Shokyo” and
PGC-1α Activity Fold Induction

2.5 *** “White ginger” showed strong stimulation, while “Kankyo” and
** “Red ginger” showed none. From a comparison of the amounts of
2 gingerol-related compounds, it is considered that [10]-shogaol (6)
contributes to the PGC-1α activity.
1.5 * *
Table 1: Mass spectrometric characterization and retention times (RT) of gingerol-
1 related compounds in ginger detected by LC-MS.
RT (min) Compound Formula Abundant Ions Measured Calculated
-
20.71 [6]-Gingerol (1) C17H26O4 [M-H] 293.1737 293.1758
0.5 24.04 [8]-Gingerol (2) C19H30O4 [M-H]
-
-
321.2042 321.2071
27.10 [10]-Gingerol (3) C21H34O4 [M-H] 349.2366 349.2384
-
24.90 [6]-Shogaol (4) C17H24O3 [M-H] 275.1640 275.1646
-
0 27.95 [8]-Shogaol (5) C19H28O3 [M-H] 303.1946 303.1966
-
30.60 [10]-Shogaol (6) C21H32O3 [M-H] 331.2252 331.2279
Control (A) (B) (C) (D) 19.37 (3S,5S)-[6]-Gingerdiol (7) C17H28O4 [M-H]
-
295.1918 295.1915
-
19.97 (3R,5S)-[6]-Gingerdiol (8) C17H28O4 [M-H] 295.1921 295.1915
Figure 2: PGC-1α activity of the methanol extracts of (A) “Shokyo”, (B) “White ginger”, (C) 22.83 (3S,5S)-[8]-Gingerdiol (9) C19H32O4 [M-H]
-
323.2203 323.2228
-
“Kankyo” and (D) “Red ginger”. 23.47 (3R,5S)-[8]-Gingerdiol (10) C19H32O4 [M-H] 323.2213 323.2228
-
*P < 0.05; ** P < 0.01; *** P < 0.001 vs control (DMSO alone) 25.97 (3S,5S)-[10]-Gingerdiol (11) C21H36O4 [M-H] 351.2535 351.2541
-
26.65 (3R,5S)-[10]-Gingerdiol (12) C21H36O4 [M-H] 351.2546 351.2541
+
26.10 [6]-Gingerdiol 3,5-diacetate (13, 14) C21H32O6 [M+H-2AcOH] 261.1856 261.1849
[M+Na]+ ions of [6]-gingerdiol 3,5 diacetate (13, 14). The
+
[M+H-AcOH] 321.2063 321.2060
+
[M+Na] 403.2097 403.2091
compounds observed in the chromatograms were identified based 28.85 [8]-Gingerdiol 3,5-diacetate (15, 16) C23H36O6 [M+H-2AcOH]
[M+H-AcOH]
+
+
289.2126
349.2359
289.2162
349.2373
on high resolution mass spectral data and a comparison of the [M+Na]
+
+
431.2395 431.2404
31.12 [10]-Gingerdiol 3,5-diacetate (17, 18) C25H40O6 [M+Na] 459.2735 459.2717
retention times with those of standard reference compounds, as well
as comparing the mass spectra with the reported ones shown in Table 2: Weight content of gingerol-related compounds in the methanol extracts.
Table 1 [9a-9c]. The abundant ion in the negative ion mass spectra Compound
Shokyo
Weight content (%) in the methanol extracts
White ginger Kankyo Red ginger
of the gingerols, shogaols and gingerdiols (1-12) is the [M˗H]˗ ion, [6]-Gingerol (1) 6.30 11.57 12.73 10.91
[8]-Gingerol (2) 2.56 3.55 3.87 2.70
while the abundant ions in the positive ion mass spectra of [10]-Gingerol (3) 4.96 5.81 5.86 2.10

gingerdiol 3,5-diacetates (13-18) are [M+Na]+, [M+H-AcOH]+ and


[6]-Shogaol (4) 3.58 3.58 4.11 3.90
[8]-Shogaol (5) 1.20 1.27 1.04 1.11
[M+H-2AcOH]+. Furthermore, the corresponding stereoisomers of [10]-Shogaol (6)
(3S,5S)-[6]-Gingerdiol (7)
2.21
0.58
2.18
0.39
1.08
0.26
1.10
0.30
gingerdiol 3,5-diacetates (13-18) were unable to be separated under (3R,5S)-[6]-Gingerdiol (8)
(3S,5S)-[8]-Gingerdiol (9)
0.73
0.06
0.55
0.07
0.64
0.05
1.25
0.06
the present analytical conditions as shown in Figure 3. (3R,5S)-[8]-Gingerdiol (10)
(3S,5S)-[10]-Gingerdiol (11)
0.06
0.21
0.05
0.23
0.04
0.16
0.11
0.08
(3R,5S)-[10]-Gingerdiol (12) 0.12 0.12 0.07 0.21
[6]-Gingerdiol 3,5-diacetate (13 + 14) 0.23 0.09 0.24 0.57
The amounts of gingerol-related compounds in the methanol [8]-Gingerdiol 3,5-diacetate (15 + 16) trace
a
a
trace
a
a
trace
a
a
0.05
extracts of “Shokyo”, “Kankyo”, “Red ginger” and “White ginger” a
[10]-Gingerdiol 3,5-diacetate (17 + 18) trace trace trace 0.03
Not exceed 0.01 %
are shown in Table 2. “Shokyo” and “White ginger” have higher
amounts of [10]-shogaol (6) than “Kankyo” and “Red ginger”. In Doui and Mikage have said that the ratio of shogaol to gingerol
addition, “Red ginger” contains a larger amount of [6]-gingerdiol ([S/G]) is an important factor that affects the quality of the ginger
3,5-diacetate (13, 14). So far, most of the studies on ginger [11]. In general, it is considered that “Kankyo” has a higher S/G
constituents and the quantification of them have mainly been ratio than “Shokyo” because gingerols are easily transformed to
focused on gingerols (1-3) and shogaols (4-6) [10a, 10b], and there shogaols during the heating process. Recently, Akiba et al., reported
are few reports on chemical profiling of other gingerol-related that the S/G ratio of “Kankyo” in the Japanese crude drug market
compounds (1-18) [10c]. With the analytical conditions used in this had a wide distribution and occasionally the ratio of “Kankyo” is
Thermogenic properties of ginger Natural Product Communications Vol. 13 (7) 2018 871

(x100,000,000) (x100,000,000)

(intensity)

(intensity)
(A) (C)

(1) (1)
1.0 (2) 1.0 (2)
(3) (x 10) (3) (x 
(4) (x 2) ( ) (x 2)
(5) (x 2) ( ) (x 2)
(6) (x 2) ( ) (x 2)
(7) (x 2) ( ) (x 2)
(8) (x 2) ( ) (x 2)
(9) (x 2) ( ) (x 2)
(1 ) (x 2) (10) (x 2)
(1 ) (x 2) (11) (x 2)
(1 ) (x 2) (12) (x 2)
0.0 0.0
10 20 30 (min 40 10 20 30 (min 40
(x100,000,000) (x100,000,000)
(intensity)

(intensity)
(B (D)
(1) (1)
1.0 (2) 1.0 (2)
(3) (x 10) (3) (x 10)
(4) (x 2) (4) (x 2)
(5) (x 2) (5) (x 2)
(6) (x 2) (6) (x 2)
(7) (x 2) (7) (x 2)
(8) (x 2) (8) (x 2)
(9) (x 2) (9) (x 2)
(1 ) (x 2) (10) (x 2)
(1 ) (x 2) (11) (x 2)
(1 ) (x 2) (12) (x 2)
0.0 0.0
10 20 30 (min 40 10 20 30 (min 40
Figure 3: LC-MS total ion chromatograms (TIC) and mass chromatograms of the methanol extracts of (A) “Shokyo”, (B) “White ginger”, (C) “Kankyo” and (D) “Red ginger”. (1)
TIC positive ion mode; (2) TIC negative ion mode; (3) m/z 403.2091 ([M+Na]+ of 13 and 14); (4) m/z 293.1758 ([M-H]- of 1); (5) m/z 321.2071 ([M-H]- of 2); (6) m/z 349.2384 ([M-
H]- of 3); (7) m/z 275.1646 ([M-H]- of 4); (8) m/z 303.1966 ([M-H]- of 5); (9) m/z 331.2279 ([M-H]- of 6); (10) m/z 295.1915 ([M-H]- of 7 and 8); (11) m/z 323.2228 ([M-H]- of 9 and
10); (12) m/z 351.2541 ([M-H]- of 11 and 12).

1.8 *** ***


**
PGC-1α Activity Fold Induction

1.6
** **
1.4 **

1.2

0.8

0.6

0.4

0.2

0
Control 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
Figure 4: PGC-1α activity of gingerol-related compounds. The numbers in the figure indicate the compounds in Figure 1.
*P < 0.05; ** P < 0.01; *** P < 0.001 vs control (DMSO alone).

lower than that of “Shokyo” [12]. The present results support these contribution to the activity. [10]-shogaol (6) stimulated PGC-1α in
findings. a concentration-dependent manner, as shown in Figure 5. As a
consequence (Table 2, Figure 1, 5), it was clarified that [10]-
To identify the constituents responsible for the PGC-1α activity of shogaol (6) contributes to the PGC-1α activity.
ginger, the gingerol-related compounds (1-18) shown in Figure 1
were isolated or prepared. The PGC-1α activities of compounds 1- In this study, the PGC-1α activity of methanol extracts of
18 are shown in Figure 4. [10]-gingerol (3), [10]-shogaol (6), [10]- “Shokyo”, “Kankyo”, “Red ginger” and “White ginger” was
gingerdiols (11, 12) and [10]-gingerdiols 3,5-diacetate (17, 18) investigated. By comparing the constituents in the four ginger
stimulated significant PGC-1α. These results indicate that the samples, we found that [10]-shogaol (6) contributes to the PGC-1α
length of the alkyl side chains in gingerols is a more important activity. The results for the PGC-1α activity of the gingerol-related
factor for PGC-1α activity than the substitutional groups at the 3- compounds (1-18) indicate that [10]-gingerol (3), [10]-shogaol (6),
and 5- positions. Recently, Chen et al., has reported that [10]- [10]-gingerdiols (11, 12) and [10]-gingerdiols 3, 5-diacetate (17,
gingerol (6) is metabolized in the human body into two major 18) stimulate significant PGC-1α. The various bioactivities of
metabolites, (3S,5S)-[10]-gingerdiol (11) and (3R,5S)-[10]- gingerol- related compounds with long alkyl side chains have been
gingerdiol (12) [13]. Our results indicate that both [10]-gingerol (3) reported. van Breemen et al., has reported that [10]-shogaol (6)
and the [10]-gingerdiols (11, 12) showed PGC-1α activity. specifically inhibited COX-2 [14a], and Ho et al., has reported that
Therefore, it is considered that the PGC-1α activity due to [10]- 10-gingerol (3) showed anti-neuroinflammatory properties through
gingerol (3) is retained even if [10]-gingerol (3) metabolizes in the blocking NF-κB activation [14b]. In addition, several studies on
human body. Additionally, the PGC-1α activity of [10]-shogaol (6) 5-HT3 receptor inhibition, NO inhibition and antioxidation by
in different concentrations (1-8 M) was examined to confirm its gingerol-related compounds have been reported [14c-14e]. The
872 Natural Product Communications Vol. 13 (7) 2018 Nishidono et al.

B-811 LSV (BUCHI, Flawil, Switzerland) yielding 2.06 g of “Red


ginger” extract, 1.78 g of “Kankyo” extract and 2.32 g of “White
ginger” extract. “Shokyo” (500 g) was extracted with methanol
three times under reflux for 1 hr, then concentrated to yield the
methanol extract (50.92 g). To analyze the chemical constituents by
LC-MS, the methanol extracts were dissolved in methanol (10
mg/ml).

Isolation of Gingerols and Shogaols: The methanol extract of


“Shokyo” was suspended in water and extracted with ethyl acetate
(AcOEt), n-BuOH and water to give the AcOEt fraction (26.58 g),
the n-BuOH fraction (5.73 g) and the water fraction (11.94 g). The
Figure 5: PGC-1α activity of [10]-shogaol.
ethyl acetate fraction was subjected to Wakogel C-200 column
*P < 0.05; ** P < 0.01; *** P < 0.001 vs control (DMSO alone) chromatography (Wako Pure Chemical Industries) using a solvent
of hexane-ethyl acetate (7:3 → 0:10), from which 8 fractions were
present results and these studies indicated that gingerol-related obtained. A part of fraction 6 (2.0 g) was subjected to an ODS
compounds also contribute to the bioactivities of ginger. column (2L UNIVERSAL COLUMN Packed with High
Performance Silica Gel, Yamazen Corporation, Osaka, Japan) with
Experimental a solvent of methanol-water (7:3 → 10:0) to yield compounds 1
(903 mg), 2 (392 mg) and 3 (283 mg). A part of fraction 3 (2.5 g)
Plant materials and Reagents: Japanese Pharmacopoeia grade was subjected to an ODS column with a solvent of methanol-water
“Shokyo” (dried rhizome of Z. officinale var. rubens) and “Kankyo” (7:3 → 10:0) to yield compounds 4 (725 mg), 5 (287 mg) and 6
(steamed and dried rhizome of Z. officinale var. rubens) were (367 mg). The following compounds were identified by comparing
purchased from Tochimoto Tenkaido (Osaka, Japan). “Red ginger” the EI-MS and 13C-NMR spectral data with those reported [15, 16].
(Indonesian dried rhizome of Z. officinale var. rubrum) and “White
ginger” (Indonesian dried rhizome of Z. officinale var. amarum) Synthesis of Gingerdiols: [6]-, [8]-, [10]-gingerdiols were obtained
were purchased from M&K Laboratories (Nagano, Japan). All the by reduction of the respective [6]-, [8]-, [10]-gingerols using
specimens were deposited in the Museum of Materia Medica, sodium borohydride. As a result of the reduction, two diastereomers
College of Pharmaceutical Science, Ritsumeikan University (RIN). were obtained. Further subjection to the silica gel column (2L
Analytical grade chemicals and LC-MS grades of chromatographic UNIVERSAL COLUMN Packed with High Performance Silica
solvents and reagents were purchased from Wako Chemical Co. Gel, Yamazen Corporation, Japan), separated out the (3S,5S) and
Ltd. (Tokyo, Japan). (3R,5S) isomers. The isomers were identified by comparison with
the reported 13C-NMR spectral data [17].
Analytical instruments: The LC-MS analyses were performed
using a Shimadzu LC-IT-TOF mass spectrometer (Shimadzu, Synthesis of Gingerdiol 3,5-diacetates: The respective gingerdiol
Kyoto, Japan) equipped with an ESI interface. The ESI parameters 3,5-diacetates were obtained by the acetylation of the corresponding
were as follows: source voltage, +4.5 kV (positive ion mode) or gingerdiols using acetic anhydride and pyridine followed by
˗3.5kV (negative ion mode); capillary temperature, 200 °C; hydrolysis using ammonium hydroxide. These compounds were
nebulizer gas flow rate, 1.5 l/min. The mass spectrometer was identified by the EI-MS and 13C-NMR spectral data.
operated in the positive and negative ion modes, scanning from m/z
150 to 1500. A Waters Atlantis T3 column (2.1 mm × 150 mm, 5 Quantification: The compounds 1-18 were dissolved in methanol to
m) was used and the column temperature was maintained at 40 °C. make standard solutions (0.1 mg/mL). The [M˗H]˗ ion shown in
The mobile phase was a binary eluent of (A) 5 mM (NH4)OAc Table 1 was selected for quantification of compounds 1-12 and the
solution and (B) CH3CN under the following gradient conditions: individual [M+Na]+ ion was selected for compounds 13-18.
0–30 min, linear gradient from 10% to 100% B, 30–40 min,
isocratic at 100% B. The flow rate was 0.2 ml/min. GC-MS Measurement of the PGC-1α activity using a one-hybrid system:
analyses were performed with a Shimadzu QP 2010 mass The PGC-1α activity assay system was established, and the activity
spectrometer (Shimadzu, Kyoto, Japan) equipped with a Shimadzu was measured using the method presented in our previous paper
GC 2010 gas chromatography system and AOC-20i autosampler. A [18]. The cells were allowed to adhere for 3 hr and then treated
DB-5MS capillary column (0.25 mm × 30 m, film thickness 0.25 with the samples dissolved in DMSO. Sample solutions were
μm, Agilent Technologies, Santa Clara, CA) was used. The mass prepared at a final concentration of 0.001 % (w/v) for the methanol
spectrometry conditions were as follows: ionization mode, EI; extracts and 5 μM for the gingerol-related compounds. After 20 hr
ionization current, 300 μA; ionization voltage, 70 eV. The GC incubation, the cells were lysed, and the luciferase activity was
parameters were as follows: the injector and transfer line were measured by a ONE-Glo™ Luciferase Assay System (Promega).
maintained at 270 °C. The oven temperature was programmed as The fold induction was calculated.
follows: initial temperature, 50 °C; initial hold time, 3 min;
temperature ramp-up rate, 10 °C/min; final temperature, 300 °C; Statistical analysis: Data were expressed as mean ± SD (n = 3).
final hold time, 5 min. The flow rate of the carrier gas (helium) was Differences were analyzed using Student’s t test. Statistical
1 ml/min. 1H- and 13C-NMR spectra were measured in CDCl3 using significance was set at *P < 0.05, ** P < 0.01 and *** P < 0.001 vs
a JNM-ECS400 NMR spectrometer (JEOL Ltd., Tokyo, Japan) with control (DMSO alone).
tetramethylsilane as an internal standard.
Acknowledgments - This work was supported by a research grant
Sample preparation: Pulverized “Red ginger” (15.49 g), “Kankyo” from the Program for Asia-Japan Research Development,
(21.73 g) and “White ginger” (15.42 g) were extracted with Ritsumeikan University.
methanol under reflux for 40 min using an Buchi Extraction System
Thermogenic properties of ginger Natural Product Communications Vol. 13 (7) 2018 873

References
[1] (a) Shukla Y, Singh M. (2007) Cancer preventive properties of ginger: A brief review. Food and Chemical Toxicology, 45, 683−690; (b) Altman
RD, Marcussen KC. (2001) Effects of a ginger extract on knee pain in patients with osteoarthritis. Arthritis & Rheumatology, 44, 2531–2538; (c)
Wang S, Zhang C, Yang G, Yang Y. (2014) Biological properties of 6-gingerol: a brief review. Natural Product Communications, 9, 1027-1030.
[2] (a) Ali BH, Blunden G, Tanira MO, Nemmar A. (2008) Some phytochemical, pharmacological and toxicological properties of ginger (Zingiber
officinale Roscoe): a review of recent research. Food and Chemical Toxicology, 46, 409−420; (b) Hoferl M, Stoilova I, Wanner J, Schmidt E,
Jirovetz L, Trifonova D, Stanchev V, Krastanov A. (2015) Composition and comprehensive antioxidant activity of ginger (Zingiber officinale)
essential oil from Ecuador. Natural Product Communications, 10, 1085-1090; (c) Gupta S, Pandotra P, Ram G, Anand R, Gupta AP, Husain MK,
Bedi YS, Mallavarapu GR. (2011) Composition of a monoterpenoid-rich essential oil from the rhizome of Zingiber officinale from North Western
Himalayas. Natural Product Communications, 6, 93-96; (d) Stappen I, Hoelzl A, Randjelovic O, Wanner J. (2016) Influence of essential ginger oil
on human psychophysiology after inhalation and dermal application. Natural Product Communications, 11, 1565-1568; (e) Semwal RB, Semwal
DK, Combrinck S, Viljoen AM. (2015) Gingerols and shogaols: important nutraceutical principles from ginger. Phytochemistry, 117, 554−568.
[3] (a) Iwami M, Terada S, Sunahara M, Shimooka R, Shimazu T. (2003) Effect of ginger and its effective components on energy expenditure in rats.
Journal of Japan Society of Nutrition and Food Science, 56, 159–165; (b) Fujisawa F, Nadamoto T, Fushiki T. (2005) Effect of intake of ginger on
peripheral body temperature. Journal of Japan Society of Nutrition and Food Science, 58, 3–9.
[4] (a) Morera E, De PL, Morera L, Moriello AS, Nalli M, Di MV, Ortar G. (2012) Synthesis and biological evaluation of [6]-gingerol analogues as
transient receptor potential channel TRPV1 and TRPA1 modulators. Bioorganic & Medicinal Chemistry Letters, 22, 1674−1677; (b) Yoshitomi T,
Oshima N, Goto Y, Nakamori S, Wakana D, Anjiki N, Sugimura K, Kawano N, Fuchino H, Iida O, Kagawa T, Jinno H, Kawahara N, Kobayashi Y,
Maruyama T. (2017) Construction of Prediction Models for the Transient Receptor Potential Vanilloid Subtype 1 (TRPV1)-Stimulating Activity of
Ginger and Processed Ginger Based on LC-HRMS Data and PLS Regression Analyses. Journal of Agricultural and Food Chemistry, 65, 3581-
3588; (c) Iwasaki Y, Morita A, Iwasawa T, Kobata K, Sekiwa Y, Morimitsu Y, Kubota K, Watanabe T. (2006) A nonpungent component of
steamed ginger-[10]-shogaol―increases adrenaline secretion via the activation of TRPV1. Nutritional Neuroscience, 9, 169-178.
[5] Luo Z, Ma L, Zhao Z, He H, Yang D, Feng X, Ma S, Chen X, Zhu T, Cao T, Liu D, Nilius B, Huang Y, Yan Z, Zhu Z. (2012) TRPV1 activation
improves exercise endurance and energy metabolism through PGC-1α upregulation in mice. Cell Research, 22, 551−564.
[6] (a) Puigserver P, Wu Z, Park CW, Graves R, Wright M, Spiegelman BM. (1998) A cold-inducible coactivator of nuclear receptors linked to
adaptive thermogenesis. Cell, 92, 829-839; (b) Wu Z, Puigserver P, Andersson U, Zhang C, Adelmant G, Mootha V, Troy A, Cinti S, Lowell B,
Scarpulla RC, Spiegelman BM. (1999) Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator
PGC-1. Cell, 98, 115-124; (c) Liang H, Ward WF. (2006) PGC-1alpha: a key regulator of energy metabolism. Advances in Physiology Education,
30, 145-151.
[7] Sekiya K, Yoshida R, Saito T, Abe D. (2010) Enhancement of adipose differentiation of 3T3-L1 cells and activation of PPARγ by ginger
constituents. Kampo & the Newest Therapy, 19, 241-246.
[8] Doui M, Mikage M. (2012) Herbological Study on the Processing of Ginger. Kampo Medicine, 63, 266-274.
[9] (a) Park JS, Jung MY. (2012) Development of high-performance liquid chromatography-time-of-flight mass spectrometry for the simultaneous
characterization and quantitative analysis of gingerol-related compounds in ginger products. Journal of Agricultural and Food Chemistry, 60,
10015−10026; (b) Tao Y, Li W, Liang W, Van BRB. (2009) Identification and Quantification of Gingerols and Related Compounds in Ginger
Dietary Supplements Using High-Performance Liquid Chromatography-Tandem Mass Spectrometry. Journal of Agricultural and Food Chemistry,
57, 10014−10021; (c) Cheng XL, Liu Q, Peng YB, Qi LW, Li P. (2011) Steamed ginger (Zingiber officinale): Changed chemical profile and
increased anticancer potential. Food Chemistry, 129, 1785–1792.
[10] (a) He X, Bernart MW, Lian L, Lin L. (1998) High-performance liquid chromatography-electrospray mass spectrometric analysis of pungent
constituents of ginger. Journal of Chromatography A, 796, 327-334; (b) Schwertner HA, Rios DC. (2007) High-performance liquid
chromatographic analysis of 6-gingerol, 8-gingerol, 10-gingerol, and 6-shogaol in ginger-containing dietary supplements, spices, teas, and
beverages. Journal of Chromatography B, 856, 41-47; (c) Jiang H, Solyom AM, Timmermann BN, Gang DR. (2005) Characterization of gingerol-
related compounds in ginger rhizome (Zingiber officinale Rosc.) by high-performance liquid chromatography/electrospray ionization mass
spectrometry. Rapid Communications in Mass Spectrometry, 19, 2957–2964.
[11] Doui M, Mikage M. (2012) The relationship between the color and pungent compound contents of ginger subjected to heating, soaking in hot water,
or steaming. Journal of Traditional Medicines, 29, 115-123.
[12] Akiba S, Sahasi Y, Nihe K, Komiya H, Saito R, Suzuki T, Mitsuma T. (2017) Effects on pungent ingredients of “Kankyo” by processing. Abstract
of the 34th congress of the medicinal and pharmaceutical society for Wakan-Yaku, 124.
[13] Chen H, Soroka DN, Haider J, Ferri-Lagneau KF, Leung T, Sang S. (2013) [10]-Gingerdiols as the Major Metabolites of [10]-Gingerol in Zebrafish
Embryos and in Humans and Their Hematopoietic Effects in Zebrafish Embryos. Journal of Agricultural and Food Chemistry, 61, 5353-5360.
[14] (a) van BRB, Tao Y, Li W. (2011) Cyclooxygenase-2 inhibitors in ginger (Zingiber officinale). Fitoterapia, 82, 38−43; (b) Ho SC, Chang KS, Lin
CC. (2013) Anti-neuroinflammatory capacity of fresh ginger is attributed mainly to 10-gingerol. Food Chemistry, 141, 3183-3191; (c) Abdel-Aziz
H, Nahrstedt A, Petereit F, Windeck T, Ploch M, Verspohl EJ. (2005) 5-HT3 receptor blocking activity of arylalkanes isolated from the rhizome of
Zingiber officinale. Planta Medica, 71, 609-616; (d) Hong SS, Oh JS. (2012) Phenylpropanoid ester from Zingiber officinale and their inhibitory
effects on the production of nitric oxide. Archives of Pharmacal Research, 35, 315-320; (e) Masuda Y, Kikuzaki H, Hisamoto M, Nakatani N.
(2004) Antioxidant properties of gingerol related compounds from ginger. BioFactors, 21, 293-296.
[15] Lee SW, Lim J, Kim MS, Jeong J, Song G, Lee WS, Rho M. (2011) Phenolic compounds isolated from Zingiber officinale roots inhibit cell
adhesion. Food Chemistry, 128, 778-782.
[16] Jolad SD, Lantz RC, Chen GJ, Bates RB, Timmermann BN. (2017) Commercially processed dry ginger (Zingiber officinale): Composition and
effects on LPS-stimulated PGE2 production. Phytochemistry, 66, 1614-1635.
[17] Kikuzaki H, Tsai SM, Nakatani N. (1992) Gingerdiol related compounds from the rhizomes of Zingiber officinale. Phytochemistry, 31, 1783-1786.
[18] Nishidono Y, Fujita T, Kawanami A, Nishizawa M, Tanaka K. (2017) Identification of PGC-1α acivating constituents in Zingiberaceous crude
drugs. Fitoterapia, 122, 40-44.

You might also like