Sie sind auf Seite 1von 11

Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332

Contents lists available at ScienceDirect

Journal of Pharmaceutical and Biomedical Analysis


journal homepage: www.elsevier.com/locate/jpba

Method development and application of offline two-dimensional


liquid chromatography/quadrupole time-of-flight mass
spectrometry-fast data directed analysis for comprehensive
characterization of the saponins from Xueshuantong Injection
Wenzhi Yang a,1 , Jingxian Zhang a,1 , Changliang Yao a , Shi Qiu a , Ming Chen b , Huiqin Pan a ,
Xiaojian Shi a , Wanying Wu a,∗ , Dean Guo a
a
Shanghai Research Center for Modernization of Traditional Chinese Medicine, National Engineering Laboratory for TCM Standardization Technology,
Shanghai Institute of Materia Medica, Chinese Academy of Sciences, 501 Haike Road, Shanghai 201203, People’s Republic of China
b
Guangxi Wuzhou Pharmaceutical (Group) Co., Ltd., Wuzhou 543000, People’s Republic of China

a r t i c l e i n f o a b s t r a c t

Article history: Xueshuantong Injection (XSTI), derived from Notoginseng total saponins, is a popular traditional Chi-
Received 12 March 2016 nese medicine injection for the treatment of thrombus-resultant diseases. Current knowledge on its
Received in revised form 19 May 2016 therapeutic basis is limited to five major saponins, whereas those minor ones are rarely investigated.
Accepted 20 May 2016
We herein develop an offline two-dimensional liquid chromatography/quadrupole time-of-flight mass
Available online 3 June 2016
spectrometry-fast data directed analysis (offline 2D LC/QTOF-Fast DDA) approach to systematically char-
acterize the saponins contained in XSTI. Key parameters affecting chromatographic separation in 2D
Keywords:
LC (including stationary phase, mobile phase, column temperature, and gradient elution program) and
Xueshuantong Injection
Minor saponins
the detection by QTOF MS (involving spray voltage, cone voltage, and ramp collision energy) were
Offline two-dimensional liquid optimized in sequence. The configured offline 2D LC system showed an orthogonality of 0.84 and a
chromatography theoretical peak capacity of 8976. Total saponins in XSTI were fractionated into eleven samples by the
Quadrupole time-of-flight mass first-dimensional hydrophilic interaction chromatography, which were further analyzed by reversed-
spectrometry phase UHPLC/QTOF-Fast DDA in negative ion mode. The fragmentation features evidenced from 36
Fast data directed analysis saponin reference standards, high-accuracy MS and Fast-DDA-MS2 data, elemental composition (C < 80,
Structural elucidation H < 120, O < 50), double-bond equivalent (DBE 5–15), and searching an in-house library of Panax notogin-
seng, were simultaneously utilized for structural elucidation. Ultimately, 148 saponins were separated
and characterized, and 80 have not been isolated from P. notoginseng. An in-depth depiction of the chem-
ical composition of XSTI was achieved. The results obtained would benefit better understanding of the
therapeutic basis and significant promotion on the quality standard of XSTI as well as other homologous
products.
© 2016 Published by Elsevier B.V.

1. Introduction components difficult to be detected, analyzed, and isolated. Aside


from the conventional phytochemical approaches, liquid chro-
Clarification of the chemical composition of traditional Chinese matography coupled to mass spectrometry (LC–MS) has been
medicine (TCM) is a premise for elucidation of the mechanism of proven as powerful in rapid characterization of herbal compo-
action and also a key segment for TCM modernization and inter- nents, in particular for discovery of minor and potentially new ones
nationalization. However, this work is often hindered because of [1]. Unfortunately, due to the limited peak capacity and the sin-
the ubiquitous interference from the chemical matrix. Additionally, gle separation mechanism used, only tens of compounds (typically
the significantly altered content difference renders those minor <50) can be characterized by once chromatographic analysis [2–4].
Despite the ongoing development of ultra-high performance liq-
uid chromatography (UHPLC) coupled with high-resolution mass
spectrometry (such as TOF, FT-ICR, Orbitrap, QTOF, IT-TOF, and LTQ-
∗ Corresponding author.
Orbitrap, etc) can greatly improve the efficiency in separation and
E-mail address: wanyingwu@simm.ac.cn (W. Wu).
1
These authors contributed equally to this work.
the reliability in structural elucidation of herbal components, the

http://dx.doi.org/10.1016/j.jpba.2016.05.035
0731-7085/© 2016 Published by Elsevier B.V.
W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332 323

characterized compounds by one-dimensional chromatographic which MSE and Fast DDA (fast data directed analysis) are two rou-
separation are, in general, still less than 100 [5–8], in which tine ones [31]. Comparatively, MSE is data-independent allowing
many minor, structurally novel, possibly pharmacologically active MS/MS fragmentation of all precursor ions without ion selec-
molecules fail to be exposed and identified. tion or separation [32], whilst Fast DDA as a data-dependent
This situation has been greatly improved in recent years due mode enables automatic intensity-ranking of precursors-triggered
to the development of enhanced chromatographic separation MS/MS fragmentation [6,31], and can greatly ease data process-
techniques and hybrid mass spectrometric scan strategies: (1) two- ing for structural elucidation. First, the parameters essential for
dimensional liquid chromatography coupled with multistage mass establishment of an offline 2D LC–MS system were optimized.
spectrometry (2D LC–MSn ) [9–11]; (2) fractionation and enrich- Orthogonality evaluation was performed using asterisk equa-
ment effects by column chromatography or knockout of major tions proposed by Camenzuli and Schoenmakers [33]. Second,
components prior to LC–MS analysis [12–15]; and (3) enhanced, the total saponins contained in XSTI were fractionated by the
integral targeted or nontargeted scan methods performed on a first-dimensional (1 D) separation. Third, each fraction sample was
triple quadrupole (QqQ) or a hybrid triple quadrupole/linear ion- analyzed by the second-dimensional (2 D) separation coupled with
trap (QTrap) mass spectrometer [16–18]. Amongst them, 2D LC, QTOF MS detection by data-dependent Fast DDA in negative mode.
with significantly extended peak capacity and versatile separation Fourth, the fragmentation pathways evidenced from 36 reference
mechanisms hybridation, has arisen much interest of the chemists standards, high-accuracy MS and MS/MS data, elemental composi-
in resolution of a complex mixture, such as an herbal extract. 2D LC tion (EC), and double bond equivalent (DBE), were simultaneously
can operate in either online or offline mode. Complicated pump- employed to characterize the substructures (including sapogenin,
ing and switching valves are always necessary for configuration of sugar residue, and acyl substituent). Fifth, an in-house library of P.
an online 2D LC system to achieve heart-cutting or comprehensive notoginseng was constructed and used to discern potentially new
analysis [9]. Alternatively, flexible integration of different separa- saponin structures. Ultimately, 148 saponins were identified or
tion mechanisms is easily accomplished in the offline mode for 2D tentatively characterized, including 80 not isolated from P. notogin-
LC without instrumental limitation [19]. A multiple heart-cutting seng, which demonstrates the remarkable superiority of the offline
2D LC approach was recently developed in our lab that enabled the 2D LC/QTOF-Fast DDA approach over the conventional methods
simultaneous quality evaluation of Panax notoginseng in eight dif- applied for the qualitative analysis of XSTI. This study focuses on
ferent Chinese patent medicines [20]. Offline comprehensive 2D LC the method development, by systematic optimization of the key
coupled with LTQ-Orbitrap MS was developed to comprehensively parameters in both offline 2D LC and QTOF MS.
characterize ginsenosides from the stems/leaves of Panax ginseng
and quinochalcone C-glycosides from Carthamus tinctorius [10,21].
Offline comprehensive 2D LC is deemed as a desirable solution to
the exposure and characterization of minor herbal components. 2. Experimental
Panax notoginseng (Burk.) F.H. Chen is used as a traditional herbal
medicine or as functional food worldwide [22,23], which has been 2.1. Chemicals and reagents
officially recorded in USP Herbal Medicines Compendium and China
Pharmacopeia [24,25]. Diverse categories of primary and secondary A total of 36 saponins (purity >95%) isolated from the roots
metabolites, such as saponins, nonprotein amino acids, polysaccha- of P. notoginseng were used as the reference standards, involving
rides, polyacetylenes, phytosterols, and flavonoids, etc, have been notoginsenoside-R1 , −R2 , −R3 , −R4 , −G, −K, −M, −T, −T5 , 20(R)-
reported or isolated from P. notoginseng. Saponins are considered noto-R2 , ginsenoside-Ra3 , −Rb1 , −Rb2 , −Rc, −Rd, −Re, −Re3 , −Rf,
closely related to the pharmacology and clinical effects of P. notogin- −Rg1 , −Rg2 , −Ro, −F2 , −Rk3 , 20(S)-Rg3 , 20(R)-Rg3 , 20-O-glc-Rf,
seng on the cardiovascular and digestive systems [26], which refer 20(S)-Rh1 , 20(R)-Rh1 , 20(S)-Rh2 , 20(R)-Rh2 , 20(S)-sanchirhinoside-
to the oligo-glycosides of dammarane type protopanaxadiol (PPD), A3 , −A4 , −A5 , vinaginsenoside-R4 , 5,6-didehydroginsenoside-Rb1 ,
protopanaxatriol (PPT), and versatile derivatives [27]. Given the and compound K. Detailed information with regard to the isolation
significant bioactivities of Notoginseng saponins on blood circula- and structural elucidation is offered as Electronic Supplemen-
tion system, diverse therapeutic agents derived from Notoginseng tary Data. Fig. 1 displays structures of these reference standards.
total saponins are clinically available, and Xueshuantong Injection Notably, Ro, Re3 , F2 , 20(S)-Rh2 , 20(R)-Rh2 , and compound K, are
(XSTI; “Zhusheyongxueshuantong” or “Xueshuantong Lyophilized first isolated from the roots of P. notoginseng, while 20(R)-noto-R2
Powder”) is ranked among the most famous. is first obtained from plants. The XSTI sample (Batch No. 14010709,
XSTI is a very popular TCM injection agent in China. Accord- 150 mg/bottle) was kindly provided by Guangxi Wuzhou Pharma-
ing to the statistic information provided by the vendor, the ceutical (Group) Co., Ltd., with the voucher specimen deposited
average annual sales between 2012 and 2014 reached 360 mil- at the author’s laboratory (ID: XSTI-14010709). HPLC-grade ace-
lion US dollars. XSTI could improve cerebral blood perfusion tonitrile, methanol (Merck, Darmstadt, Germany), formic acid (ROE
and enhance hematoma adsorption and neurological function in Scientific Inc., USA; FA), and ammonium acetate (Sigma-Aldrich,
patients [28,29]. However, systematic chemical analysis of XSTI has MO, Switzerland; AA), and ultra-pure water (18.2 M cm at 25 ◦ C)
been rarely reported. Our previous research characterized thirty prepared by a Millipore Alpha-Q water purification system (Milli-
saponins from XSTI by use of a well-developed HPLC/IT-MS method pore, Bedford, USA), were applied in this study. Leucine-enkephalin
[30]. Very recently, Wang et al. used UHPLC/QTOF-MS combined was purchased from Sigma-Aldrich (St. Louis, MO, USA).
with peaks-knockout to characterize 94 saponins [15]. We believe
that, by integration of different separation mechanisms in 2D LC,
more minor, unknown saponin molecules can be discovered and
identified from XSTI. 2.2. Sample preparation
In this study, we develop an offline 2D LC/QTOF MS method
to achieve the comprehensive characterization of minor saponins A stock solution was prepared by dissolving a quantity of the
from XSTI following a protocol we recently established [10]. Dif- lyophilized power in 50% aqueous methanol (v/v) to reach a con-
ferent from the former studies [10,21], the MS detection was centration of 10 mg/mL. The solution was then filtered through a
performed on a Waters Xevo® G2-S QTOF mass spectrometer in 0.2-␮m PTFE filter (Agilent Technologies, USA) and stored at 4 ◦ C
the current study. It facilitates multiple scan methods, among prior to use.
324 W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332

Fig. 1. Structure of 36 saponin reference standards used in this study. Glc: glucose; GlurA: glucuronic acid; Rha: rhamnose; Ara(p)/Ara(f): pyran-/furan-arabinose; Xyl: xylose.

2.3. Offline 2D LC separation of the saponins in XSTI Each 40 ␮L stock solution of XSTI (10 mg/mL) was injected onto the
column, with five replications. Pooled eluate for each fraction was
Combination of HILIC × RPLC was used to separate the saponin dried under a steady flow of nitrogen (N2 ) at ambient temperature.
components contained in XSTI. In detail, the total saponins were The dried residues of eleven fractions were separately reconstituted
initially separated in HILIC mode on an Agilent 1100 HPLC system in 400 ␮L 50% aqueous methanol and filtered through a 0.2-␮m
(Agilent Waldbronn, Germany) using a Waters Xbridge Amide col- PTFE filter ahead of RP-UHPLC/QTOF-Fast DDA analysis.
umn (4.6 × 150 mm, 3.5 ␮m) maintained at 25 ◦ C. A binary mobile Afterwards, eleven fractionated samples were subjected to RPLC
phase consisting of acetonitrile (A) and water (B) at a flow rate separation on a Waters ACQUITY I-Class UPLC® system (Waters,
of 1.0 mL/min was used following a gradient elution program: Milford, MA, USA) equipped with a binary solvent manager, a sam-
0–12 min: 90%–83% (A); 12–25 min: 83%–70% (A); 25–28 min: ple manager, a column manager, and a PDA detector. An ACQUITY
70%–50% (A); 28–30 min: 50% (A); 30–31 min: 50%–90% (A); and UPLC HSS T3 column (2.1 × 100 mm, 1.8 ␮m) at 25 ◦ C was used for
31–41 min: 90% (A). The VWD detector was set at 203 nm for detec- chromatographic separation with a binary mobile phase consisting
tion of saponins. Eluate collection was based on time-dependent of acetonitrile (A) and water containing 0.1% formic acid (B) at a flow
fractionation, giving eleven fractions: 0–5 min as Fr. 1, 5–10 min rate of 0.3 mL/min. An optimized gradient elution program was as
as Fr. 2, 10–11 min as Fr. 3, 11–13 min as Fr. 4, 13–16 min as Fr. 5, follows: 0–3 min: 20%–23% (A); 3–6 min: 23%–24% (A); 6–6.5 min:
16–18.5 min as Fr. 6, 18.5–20.5 min as Fr. 7, 20.5–21.5 min as Fr. 8, 24%–30% (A); 6.5–11.5 min: 30%–32% (A); 11.5–16 min: 32%–65%
21.5–24 min as Fr. 9, 24–27 min as Fr. 10, and 27–30 min as Fr. 11. (A); 16–17.5 min: 65%–95% (A); 17.5–21 min: 95% (A); 21–21.5 min:
W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332 325

95%–20% (A); and 21.5–25 min: 20% (A). The PDA detector was set Eqs. (6)–(9) for indication of peaks spreading degrees. Ultimately,
between 190 and 400 nm and at 203 nm for detection of saponins. A0 was employed to assess the orthogonality according to Eq. (10).
An injection volume of 5 ␮L was set for each sample. tI − tD
tR,norm(i) = (1)
tG − tD

Sz- = {1 t R,norm(i) − 2 t R,norm(i) } (2)


2.4. QTOF-Fast DDA
Sz+ = {2 t R,norm(i) − (1 − 1 t R,norm(i) )} (3)
High-accuracy MS data of saponins were acquired on a Waters
1
Xevo® G2-S QTOF mass spectrometer (Waters, Manchester, UK) Sz1 = { t R,norm(i) − 0.5} (4)
connected to UPLC system via a ZprayTM ESI source. In order to
obtain more reliable characterization results, data-dependent Fast Sz2 = {2 t R,norm(i) − 0.5} (5)
DDA in negative, Resolution mode was utilized. The ESI source Z- = |1 − 2.5|Sz - − 0.4|| (6)
parameters were set as follows: spray voltage, 2.5 kV; cone volt-
age, 140 V; source offset voltage, 80 V; source temperature, 100 ◦ C; Z+ = |1 − 2.5|Sz + − 0.4|| (7)
desolvation temperature, 400 ◦ C; cone gas flow rate, 40 L/h; and Z1 = 1 − |2.5 × Sz 1 × 2 − 1| (8)
desolvation gas flow rate, 800 L/h. For Fast DDA settings, the mass
analyzer scanned over a mass range of 400–1500 Da in full scan with Z2 = 1 − |2.5 × Sz 2 × − 1| (9)
a scan time of 0.15 s under the collision energy of 6 V, and over m/z 
A0 = Z− × Z+ × Z1 × Z2 (10)
100–1400 for MS/MS by the same scan time. Dual-dynamic colli-
sion energies, 20–40 V for LM CE (low-mass collision energy) and
60–80 V for HM CE (high-mass collision energy), were employed 3. Results and discussion
to acquire more balanced fragmentation independent of molecular
mass difference. Automatic switching to MS/MS mode was enabled 3.1. Parameters optimization for the offline 2D LC system
when TIC intensity rose above 5000 intensity/s, and switching off
when 0.45 s had elapsed. Tolerance window of ±3.0 Da and peak 3.1.1. Selection of chromatographic columns
extract window of 7.0 Da were set in deisotope peak detection Given that integration of different separation mechanisms is
mode. Data calibration was performed using an external reference able to enhance the selectivity and thus benefits the exposure of
(LockSprayTM ) by constant infusion of the leucine-enkephalin solu- minor components in a larger number, different stationary phases
tion (1 ng/␮L) at a flow rate of 5 ␮L/min. MS data were referenced representing versatile separation mechanisms were examined in
to m/z 554.2615. MassLynx V4.1 software (Waters, Milford, USA) the first step. Meanwhile, taking into account the compatibil-
was used in data acquisition and processing. For prediction of EC of ity with mass spectrometer and the retention information as
an observed precursor ion, the ranges of EC and DBE were strictly important evidence for unknown saponins characterization and
predefined: (1) C < 80, H < 120, O < 50; and (2) DBE 5–15. An in- isomers discrimination by having recourse to literature, we initially
house library that records the structure information of 95 saponins employed RPLC for 2 D separation due to its wide application in
ever isolated from P. notoginseng was established and employed in Ginseng analysis [2,3,7,12–15,18,30,34]. Six columns suitable for
structural elucidation. UHPLC separation from three different vendors (Waters, Agilent,
and Phenomenex) were tested in respect of selectivity and reso-
lution (Supplementary Fig. A.1). Apparently, HSS T3 and CSH C18
columns exhibited stronger retention capacity to saponins than the
2.5. Method validation for offline 2D LC/QTOF-Fast DDA other four, in particular, the most polar ginsenosides (eluted before
noto-R1 , tR 5.20 min) were resolved on the HSS T3 column. Addi-
The established offline 2D LC/QTOF-Fast DDA method was val- tionally, the largest number of medium-polarity saponins (eluted
idated in respect of intra-day/inter-day precision (both 1 D and 2 D between Rg1 and Rb1 ) were separated by HSS T3. It suggested HSS
separation), repeatability, and limit of detection (LOD). Precision T3 had the potential to resolve the most minor saponins in 2 D
was evaluated by six consecutive injections on the first day and separation.
another three injections separately on the second and third days. Subsequently, another six columns having different separation
Repeatability was assessed by analysis of six reduplicative samples mechanisms or bonded with different functional groups were com-
of Fr. 5 (containing noto-R1 , Re, and Rd). LOD, defined as the low- pared in terms of their separation difference to HSS T3, determined
est amount at which a saponin compound can be characterized by by linearity regression correlation coefficient (R2 ) using fourteen
the Fast DDA method, was determined using five major saponins saponins as the index components. As shown in Supplementary
(noto-R1 , Rg1 , Re, Rb1 , and Rd). Fig. A.2, BEH Phenyl, CSH Phenyl-Hexyl, HSS Cyano, and HSS PFP
columns, all showed R2 > 0.9, indicating their similar selectivity to
HSS T3 on separation of saponins. In contrast, two HILIC columns
2.6. Orthogonality evaluation exhibited high orthogonality to HSS T3, with R2 < 0.1. We finally
selected the Xbridge Amide column in 1 D separation because of
Orthogonality of the offline 2D LC system was assessed by cal- its greater separation difference from HSS T3, consistent with our
culating the distribution of 50 index components using an in-house previous report in comprehensive characterization of ginsenosides
executable program we have reported [21]. Peak distribution was from the stems and leaves of P. ginseng [10]. The results demon-
depicted according to a series of asterisk equations reported in lit- strated the combined use of HILIC and RPLC was potent in resolution
erature [33]. In brief, normalized retention time (tR ,norm(i) ) of each of polar or medium-polarity herbal components [10,19,21,35].
index component in each dimension was calculated according to
Eq. (1) (tD : dead time; : tG total gradient elution time). Spreading 3.1.2. The 1 HILIC separation
of all the index components (Sz- , Sz+ , Sz1 , and Sz2 ) around four One of the advantages of offline mode of 2D LC over the online
lines were calculated based on Eqs. (2)–(5) (␴: standard deviation mode is embodied in the convenient, individual optimization of
of the values of the bracketed equations for all 50 index compo- chromatographic conditions in each dimension, without consid-
nents), which were then used to determine four Z parameters by ering solvent effect and limitations in sampling frequency and
326 W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332

injection volume [10,21]. The 1 D HILIC condition was optimized in


terms of different compositions of mobile phase, column tempera-
ture, and gradient elution program, on an Xbridge Amide column. It
was found that the resolution of XSTI saponins was seldom affected
by use of different mobile phases (CH3 CN-H2 O, CH3 CN-0.1%FA,
and CH3 CN-2 mM AA) (Supplementary Fig. A.3). Acyl ginsenosides
(such as malonyl-ginsenosides) in the roots of P. notoginseng have
been removed in XSTI, leaving neutral saponins [15,30]. Retention
of neutral saponins was rarely affected by different pH values of
the mobile phase due to the addition of different additives. On the
other hand, when the column temperature increased from 25 ◦ C to
40 ◦ C, the retention times of major saponins only slightly decreased
(Supplementary Fig. A.4). Low column temperature favored the
separation of a small peak eluted between Re and noto-R1 . Con-
sequently, the 1 D HILIC separation was performed on an Xbridge
Amide column maintained at 25 ◦ C with a binary mobile phase
composed of CH3 CN-H2 O at a flow rate of 1.0 mL/min.

3.1.3. The 2 D chromatographic separation


The 2 D chromatographic separation condition, including the
mobile phase, column temperature, and gradient elution program,
was optimized. As shown in Supplementary Fig. A.5, different
compositions of mobile phase (CH3 CN-H2 O, CH3 CN-0.1% FA, and
CH3 CN-2 mM AA) remarkably influenced both the chromato-
graphic separation and ion response of saponins by using QTOF MS
as the detector. The largest number of peaks were resolved by use
of CH3 CN-0.1% FA as the mobile phase, despite the most abundant Fig. 2. Orthogonality evaluation of the offline 2D LC system by a series of asterisk
equations.
peak Rg1 was weaker than that obtained using CH3 CN-2 mM AA as
the mobile phase. In addition, much less complicated ion species
corresponding to the same compound (such as Ra3 , Supplementary 3.2. Parameters optimization of QTOF-Fast DDA
Fig. A.6) was generated by addition of 0.1% FA in the mobile phase,
which is crucial for intensity-ranked automatic selection of precur- Even if a few of reports are available that employ Fast DDA
sors in Fast DDA. On the other hand, little influence was observed for characterization of herbal components, a systematic parame-
when different column temperature was set (Supplementary Fig. ters optimization study has not been reported hitherto. Herein, the
A.7). Taking into account the resolution of minor peaks around Rb1 , ESI source parameters and ramp collision energy were optimized,
we performed the 2 D separation with the HSS T3 UPLC® column aiming to achieve sensitive detection and characterization of XSTI
maintained at 25 ◦ C and a binary mobile phase of CH3 CN-0.1% FA saponins based on abundant precursor-product ions information
at a flow rate of 0.3 mL/min. recorded by Fast DDA.

3.2.1. ESI source parameters


The ESI source parameters, including spray voltage (SV) and
3.1.4. Peak capacity and orthogonality evaluation cone voltage (CV), affect ions generation in terms of relative inten-
The separation potency of the established offline 2D LC system sity of the molecular ions, adduct ions, and product ions, due to
was assessed by peak capacity and orthogonality. Theoretical peak in-source fragmentation. The objective in this step was to obtain
capacity of a comprehensive 2D LC system (n2D ) is the product of the the best ion response and, meanwhile, to suppress the adduct
peak capacity in 1 D and 2 D separation (n1 and n2 ) [35]. The aver- ions and avoid the occurrence of in-source fragmentation. Tak-
age base-line peak width of three peaks eluted at the beginning, ing into account the molecular mass coverage (600–1400 Da) of
the middle, and the end of the chromatograms in 1 D and 2 D sep- saponins ever isolated from P. notoginseng [27,36], four representa-
aration were 0.29 min and 0.24 min, respectively, based on which tive saponins, comprising noto-R2 (C41 H70 O13 , 770; a biglycoside),
n1 and n2 were calculated to be 102 and 88. Thereby, n2D should noto-R1 (C47 H80 O18 , 932; a triglycoside), Rb1 (C54 H92 O23 , 1108;
be 8976. Clearly, compared to the conventional one-dimensional a tetraglycoside), and Ra3 (C59 H100 O27 , 1240; a pentaglycoside),
chromatography, even in the UHPLC mode, the peak capacity of were taken as index components.
the 2D LC system was enhanced by approximate 100-fold. Improve- Influence of different SV settings ranging from 1.0 kV to 3.5 kV
ment on peak capacity is beneficial to the exposure of more minor was assessed by comparing the ion response of four index com-
components, laying a solid foundation for the completion of a com- ponents. In general, the peak areas positively correlated with the
prehensive elucidation of the therapeutic basis of XSTI. increasing of SV from 1.0 kV to 2.5 kV, however, slightly decreased
Orthogonality evaluation was based on a series of asterisk to a low level when 3.0 kV and 3.5 kV were set (Supplementary Fig.
equations proposed by Camenzuli and Schoenmakers [33] using A.8). No remarkable in-source fragmentation occurred under these
normalized retention time of 50 saponins (Table A.1). As a conse- SV settings (1.0–3.5 kV). As a result, SV was set at 2.5 kV.
quence, the spreading degrees (SZ ) of 50 saponins around lines Z+ , Subsequently, different CV settings (from 130 V to 180 V) were
Z- , Z1 , and Z2 , were 0.97, 0.81, 0.91, and 0.98, respectively (Fig. 2). compared by observing the transition tendency of adduct ions,
The orthogonality was calculated to be 0.84, ultimately, indicating molecular ions, and in-source fragmentation product ions. As evi-
a powerful 2D LC system for exposure of minor saponins contained denced in Supplementary Fig. A.9, with the increasement of CV
in XSTI. from 130 V to 160 V, FA-adduct ions of noto-R1 , Rb1 , and Ra3 ,
W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332 327

exhibited decreasing tendency. Significant in-source fragmenta- respectively. In addition, miscellaneous sapogenins, 5-ene-PPD in
tion (m/z 637.43, due to the neutral elimination of an external 5,6-didehydro-Rb1 and 7-OH-5-ene-PPD in noto-G exhibited rich
Xyl residue) was observed for noto-R2 when CV was set higher sapogenin ions at m/z 457.36 and m/z 473.37, respectively.
than 140 V, which was not detected for the tri- (noto-R1 ), tetra-
(Rb1 ), and penta- (Ra3 ) glycosides. Therefore, CV was set at 140 V, 3.4.2. Characterization of sugar residues
under which no in-source decay occurred and the adduct ions were The negative mode CID of ginsenosides offers indirect evidence
suppressed in lower abundance than the genuine molecular ions. (masses of neutral elimination) for the sugar moieties, and, mean-
while, the low-mass region (m/z 100–400) in the Fast DDA-MS2
3.2.2. Ramp collision energies spectra informs direct information (DPIs) for the nature, composi-
Ramp (or dynamic) collision energy (RCE) setting enables tion, and even attachment patterns of the oligosaccharide chains.
mass-dependent collision-induced dissociation (CID) of molecu- The neutrally lost masses of 162.05 Da, 146.05 Da, 132.04 Da, and
lar ions, conducing to the generation of more balanced MS/MS 176.03 Da, were separately consistent with the sugars of glucose
spectra independent of molecular mass difference. Constant col- (Glc), rhamnose (Rha), arabinose (furan- or pyran-, Ara)/xylose
lision energy (CCE) was initially set from 20 V to 70 V to acquire (Xyl), and glucuronic acid (GlurA) [10,15]. By comparing the MS2
the fragmentation behaviors of noto-R2 , noto-R1 , Rb1 , and Ra3 . spectra of 20(S)-Rh1 , Rf, 20-O-glc-Rf, and Rb1 that contain 1–4 Glc
From Supplementary Fig. A.10, rather different CID features were residue(s), the DPIs at m/z 161.05, 113.02, and 101.03, were indica-
observed. Considering the diversity of product ions and generation tive of a single Glc residue, while m/z 221.07 (not detected for
of sapogenin ions, we could infer the appropriate collision energies Re3 that contains Glc(4,1)Glc), was informative of Glc(2,1)Glc,
were 30–50 V for noto-R2 , 40–50 V for noto-R1 , and 60–70 V for Rb1 which might result from the 0,2 X0 fragmentation (Supplementary
and Ra3 . Based on these findings and considering the mass range for Fig. A.11). Likewise, the product ions at m/z 163.06 and 205.07,
full scan (400–1500 Da), as a consequence, RCE was set at 20–80 V observed for 20(S)-Rg2 and Re, were correlated to the Rha residue
in this study, by which rich CID information concerning the pre- and the oligosaccharide chain of Glc(2,1)Rha, respectively, and m/z
cursor ions, product ions, sapogenin ions, and even the secondary 337.08 in Ro and 353.11 in noto-T were ascribed to GlurA(2,1)Glc
product ions of sapogenins, were obtained sufficient for structural and Glc(2,1)Glc(2,1)Xyl, respectively. Differently, the DPIs associ-
elucidation of the saponin structures. ated with Glc(2,1)Ara in 20(S)-sanchi-A5 , Glc(2,1)Xyl in noto-R2 ,
Glc(6,1)Ara(p) in Rb2 , and Glc(3,1)Xyl in noto-T5 , were all detected
3.3. Validation of the offline 2D LC/QTOF-Fast DDA method at m/z 191.06, nonspecific for a disaccharide chain of a Glc and a
pentose. These DPIs and the eliminated masses are complementary
The established offline 2D LC/QTOF-Fast DDA method was sim- for characterization of the sugar moieties.
ply validated in precision, repeatability, and LOD. The intra-day
and inter-day precision evaluated by five major saponins for 1 D 3.5. Systematic characterization of the saponins contained in XSTI
HILIC-UV (RSD, in%) varied from 0.19%–1.48% and 0.62%–3.13%,
respectively, while 9.39%–11.14% of intra-day precision and The total saponins in XSTI were initially separated by 1 D HILIC
12.94%–16.84% of inter-day precision were determined for the 2 D separation yielding eleven fractions, which were then individually
UPLC/QTOF-Fast DDA. RSD of repeatability was less than 7%. LOD analyzed by an optimized UHPLC/QTOF-Fast DDA method (Fig. 5).
was determined at 0.1 ng for noto-R1 and Rd, and 0.5 ng for Rg1 , Re, Because of the orthogonal separation facilitated by the offline 2D LC,
and Rb1 . These data indicated an acceptable qualitative analysis numerous minor saponins that were easily covered up by the dom-
approach by means of offline 2D LC/QTOF-Fast DDA. inant ones were well exposed, and thus got characterized. Under
the current 2 D chromatographic condition, the saponins were ion-
3.4. CID features of 36 saponin reference standards ized into predominant FA-adduct ions ([M+HCOO]− ), concomitant
with [M+H2 PO4 ]− ions (except 50, 53, 90, 113, 130, 134, 138, 142,
Despite the CID features of ginsenosides have been extensively and 143 that have the richest [M−H]− ). Thereby, the structural
studied [3.4,7,10,12–15,17,18,30], different fragmentations may characterization work was carried out based on a workflow consist-
occur on different mass spectrometers. Herein, the negative mode ing of four steps: (1) simultaneous detection of dominant adduct
CID of 36 saponin reference standards (Fig. 1), representing five ions of [M+CH3 COO]− and [M+H2 PO4 ]− informs a saponin com-
different subclasses, that is, protopanaxatriol (PPT, 17 in total), pound; (2) EC of the precursor ion is deduced by strict atoms setting
protopanaxadiol (PPD, 14), oleanolic acid (OA, 1), C17-side chain (C < 80, H < 120, and O < 50), DBE screening (5–15), and isotope dis-
varied (C17-SCV, 2), and miscellaneous (2), were comparatively tribution matching (i-FIT); 3) the fragmentation pathways and DPIs
investigated to outline the characteristic fragmentation pathways detected from the Fast DDA-MS2 spectra are interpreted to charac-
and potential diagnostic product ions (DPIs) for characterization of terize the sapogenins and sugar moieties; 4) possible structures
the sapogenins and sugar residues. Fig. 3 displays the base peak are retrieved in the in-house library of P. notoginseng (record-
chromatograms (BPCs) of mixed reference standards and an XSTI ing 95 saponins, Table A.2), to discern potentially new saponins.
sample. Full-scan MS and Fast DDA-MS2 spectra of four represen- Glc, Rha, GlurA, and Xyl (used to represent all possible pentose
tative saponins (comprising 20-O-glc-Rf, Rd, Ro, and noto-T5 ) are residues), are employed to interpret the observed neutral loss of
interpreted as shown in Fig. 4. In general, the CID behaviors of 162.05 Da, 146.06 Da, 176.03 Da, and 132.04 Da, in characterization
saponin reference standards obtained on the QTOF mass spectrom- of unknown saponins from XSTI [10].
eter were similar to those on an LTQ-Orbitrap mass spectrometer The characteristic fragmentations, m/z 475.38 → 391.29 and
as we previously reported [10]. 459.38 → 375.29, were employed as major diagnostic informa-
tion to identify PPT and PPD type saponins, respectively, whilst
3.4.1. Characterization of sapogenins the sapogenin ion at m/z 455.35 and neutral loss of 176.03 Da
Sequential elimination of the attached sugar residues could gen- were informative of the OA-type characterization. Compound 76
erate DPIs associated with the sapogenins, involving m/z 475.38 (tR 7.84 min in Fr. 4) is a typical PPT-type saponin, giving rich
for PPT, m/z 459.38 for PPD, m/z 455.35 for OA, and m/z 457.36 for [M+HCOO]− and [M+H2 PO4 ]− ions at m/z 1169.59 and 1221.56,
dehydrated PPT (Fig. 4). Furthermore, relatively weak secondary respectively (Fig. 6). The molecular formula was thus determined as
product ions of deprotonated PPD and PPT, due to the neutral elimi- C54 H92 O24 . High intensity of [M−H]− at m/z 1123.59, [M−H−2Glc]−
nation of C6 H12 (84.09 Da), were observed at m/z 391.29 and 375.29, at m/z 799.49, [M−H−3Glc]− at m/z 637.43, the sapogenin ion at m/z
328 W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332

Fig. 3. Negative mode base peak chromatograms (BPC) of 36 reference standards and the XSTI sample. 30: noto-R3 ; 33: Re3 ; 34: noto-M; 39: 20-O-glc-Rf; 49: noto-R1 ; 54:
20(S)-sanchi-A5 ; 60: Rg1 ; 62: Re; 79: 20(S)-sanchi-A4 ; 80: noto-G; 89: 20(S)-sanchi-A3 ; 97: vina-R4 ; 104: noto-T; 110: Rf; 113 noto-R4 ; 118: 20(S)-noto-R2 ; 123: Ra3 ; 125:
5,6-didehydro-Rb1 ; 127: 20(R)-noto-R2 ; 1*: 20(S)-Rg2 ; 128: 20(S)-Rh1 ; 130: Rb1 ; 131: 20(R)-Rh1 ; 132: Rc; 134: Ro; 135: Rb2 ; 141: Rd; 144: noto-K; 2*: noto-T5 ; 3*: F2 ; 145:
Rk3 ; 147: 20(S)-Rg3 ; 148: 20(R)-Rg3 ; 4*: compound K; 5*: 20(S)-Rh2 ; 6*: 20(R)-Rh2 . Peaks 1*–6* represent six saponins isolated from the roots of P. notoginseng, but are rare
or not detected from XSTI in this study.

475.39, together with the oligosaccharide chain fragments at m/z the PPT framework. Thereby compound 76 was characterized as
323.10 and 221.07, were readily obtained by CID-MS/MS. All these PPT-Glc-Glc-(Glc-Glc). Compound 142 (tR 14.29 min in Fr. 6) is a
fragments informed the presence of four Glc residues attached to typical OA-type saponin since the characteristic sapogenin ion at

Fig. 4. The negative ion mode full-scan MS and Fast DDA-MS2 spectra of four saponin reference standards, representative of the PPT-type, PPD-type, OA-type, and C17-side
chain varied type. The dashed box region shows diagnostic product ions of sugar moieties.
W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332 329

Fig. 5. Fractionation of XSTI saponins by the first-dimensional HILIC. A-the UV spectrum (at 203 nm) of XSTI; B-the base peak chromatograms (BPC) of mixed reference
standards and eleven fraction samples determined by UHPLC/QTOF MS.

m/z 455.36 and neutral loss of 176 Da were acquired by CID of the amount. However, owing to the unexpected structural variation
deprotonated molecules at m/z 793.43 (Fig. 6), as those occurring to and lack of relevant reference standards, their characterization
the reference standard Ro. We thus tentatively characterized 142 as is putative in this study. Characterization of compound 63 (tR
chikusetsusaponin IVa (OA-GlurA-Glc), an OA ginsenoside reported 6.87 min in Fr. 7) was taken as an example (Fig. 6). The adduct ions
from P. japonicum [37]. Consequently, 67 PPT-type, 24 PPD-type, at m/z 1025.65 and 1077.52 could inform the molecular formula
and 2 OA-type saponins, were characterized. Notably, this is the C48 H84 O20 . The MS/MS product ions at m/z 979.54, 799.48, 637.42,
first report of OA saponins (134 and 142) from XSTI, which are 496.39, and 475.39, were ascribed to [M−H]− , [M−H−Glc−H2 O]− ,
generally considered absent in P. notoginseng [36,38]. [M−H−2Glc−H2 O]− , [sapogenin−H]− , and [sapogenin−H−H2 O]− ,
Thirty-seven saponins encompassing varied C17-side chains respectively, based on which it could be inferred that H2 O-addition
were characterized from XSTI, accounting for 31.1% of the total PPT and three Glc residues were involved in compound 63. By
330 W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332

Fig. 6. Interpretation of the MS and Fast-DDA-MS/MS spectra of compounds 76, 142, 63, and 42, representative of the PPT, OA, C17-SCV, and 7-OH-5-ene-PPD type, respectively.

comparative analysis of the fragmentation behaviors, the modifi- determined as C59 H98 O28 . CID of the adduct precursor ion at m/z
cations on C17-side chains were primarily characterized to involve 1299.62 generated MS/MS product ions at m/z 797.47, 635.42,
dehydrogenation of PPD (−H2 ; 71), hydrogenation of PPT (+H2 ; 473.35, and 353.11, which were assigned as [M−H−2Glc−Xyl]− ,
68), H2 O-addition of PPT (+H2 O; 16, 17, 35, 50, 63, and 70) or PPD [M−H−3Glc−Xyl]− , [sapogenin−H]− ([M−H−4Glc−Xyl]− ), and
(100), hydroxylation of PPT (+O; 12–15, 19, 20, 22, 38, and 44), [GlcGlcXyl + H2 O−H−120 Da]− , respectively. Given its weaker
and combination of above reactions or unknown variations (such retention than noto-G (Fig. 1) on the 2 D RPLC column, we speculat
as hydroxylation and H2 O-addition of PPT in 1, 3, 5, 6, 8, and 10). they both have the 7-OH-5-ene-PPD sapogenin. The sugar fragment
Saponins with variations on Ring B, including four contain- at m/z 353.11, as that obtained by CID of noto-T (Supplementary Fig.
ing 5-ene-PPD (90, 99, 125, and 138) and fourteen involving A.11), inferred a possible trisaccharide chain of Glc-Glc-Xyl. Based
7-OH-5-ene-PPT (42, 57, 58, 64, 66, 73, 77, 80, 87, 91, 94, 96, on these evidences, compound 42 was tentatively characterized as
112, and 115), were characterized from XSTI as well. Compound 7-OH-5-ene-PPD-Glc-Glc-(Glc-Glc-Xyl).
42 (tR 4.67 min in Fr. 10) gave abundant adduct ions at m/z As a result, a total of 148 saponins were identified or tentatively
1299.62 and 1351.59, and accordingly its molecular formula was characterized (Table A.3), comprising 30 unambiguously identi-
fied by comparing with the reference standards (including tR , MS,
and MS/MS fragments) (Fig. 3). By further searching the in-house
library, 80 of them have not been isolated from P. notoginseng. The
versatile variations on C17-side chains, different compositions of
the sugar chains or linkage patterns may result in the novelty of
these potentially new saponins. Moreover, a chemical depiction by
plotting the tR -m/z information for all characterized saponins is
implemented as shown in Fig. 7. The structural features of saponins
in XSTI can be outlined as follows: (1) PPT and C17-SCV are two
most common subclasses; (2) polarization on the C17-side chains
weakens the retention of some saponins; (3) both PPT and PPD
types exhibit a general relationship between the molecular weight
and the retention behavior, particularly the PPD-type saponins
(scattered in the blue ellipse of Fig. 7), which is useful for their
auxiliary characterization.

Fig. 7. A 2D scatter plot of all the characterized saponins from XSTI. The blue ellipse 4. Conclusion
region shows certain linearity between the molecular weight and the retention time
of PPD-type saponins on the HSS T3 UPLC® column. (For interpretation of the ref-
erences to colour in this figure legend, the reader is referred to the web version of Comprehensive characterization of the components contained
this article.) in XSTI by conventional LC–MS approaches is confronted with
W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332 331

increasing challenges arising from the cover-up of five major [8] Y.J. Chen, Z.T. Liang, Y. Zhou, G.Y. Xie, M. Tian, Z.Z. Zhao, M.J. Qin,
saponins, and the limited peak capacity and selectivity, which are Tissue-specific metabolites profiling and quantitative analyses of flavonoids
in the rhizome of Belamcanda chinensis by combing laser-microdissection
also a common issue that impedes elucidation of the therapeutic with UHPLC-Q/TOF-MS and UHPLC-QqQ-MS, Talanta 130 (2014) 585–597.
basis for most herbal medicines as well as their products. Here, [9] K.Y. Li, Q. Fu, H.X. Xin, Y.X. Ke, Y. Jin, X.M. Liang, Alkaloids analysis using
we developed an offline 2D LC/QTOF-Fast DDA approach with the off-line two-dimensional supercritical fluid chromatography × ultra-high
performance liquid chromatography, RSC Adv. 139 (2014) 3577–3587.
view of systematic exposure and characterization of the minor [10] S. Qiu, W.Z. Yang, X.J. Shi, C.L. Yao, M. Yang, X. Liu, B.H. Jiang, W.Y. Wu, D.A.
saponins contained in XSTI. Hybridation of HILIC and RPLC in offline Guo, A green protocol for efficient discovery of novel natural compounds:
2D LC showed high orthogonality for separating XSTI saponins and characterization of new ginsenosides from the stems and leaves of Panax
ginseng as a case study, Anal. Chim. Acta 893 (2015) 65–76.
approximate 100-fold improvement in peak capacity, in contrast to
[11] W.Y. Sun, L. Tong, J.Z. Miao, J.Y. Huang, D.X. Li, Y.F. Li, H.T. Xiao, H. Sun, K.S. Bi,
conventional one-dimensional RPLC separation. After parameters Separation and analysis of phenolic acids from Salvia miltiorriza and its
optimization, the negative mode data-dependent Fast DDA facili- related preparations by off-line two-dimensional hydrophilic interaction
chromatography × reversed-phase liquid chromatography coupled with ion
tated a more reliable structural characterization of minor, unknown
trap time-of-flight mass spectrometry, J. Chromatogr. A 1431 (2016) 79–88.
saponins from XSTI, particularly by searching an in-house library [12] W.Z. Yang, M. Ye, X. Qiao, C.F. Liu, W.J. Miao, T. Bo, H.Y. Tao, D.A. Guo, A
of P. notoginseng. As many as 143 minor saponins were separated strategy for efficient discovery of new natural compounds by integrating
and characterized, comprising 80 not ever isolated from P. noto- orthogonal column chromatography and liquid chromatography/mass
spectrometry analysis: its application in Panax ginseng, Panax quinquefolium
ginseng. Remarkable superiority, due to orthogonal separation and and Panax notoginseng to characterize 437 potential new ginsenosides, Anal.
high structural characterization capacity facilitated by QTOF MS, Chim. Acta 739 (2012) 56–66.
was embodied over the reported LC–MS methods and knock-out [13] Y.Y. Liu, J.B. Li, J.M. He, Z. Abliz, J. Qu, S.S. Yu, S.G. Ma, J. Liu, D. Du,
Identification of new trace triterpenoid saponins from the roots of Panax
strategies in separation and characterization of minor components notoginseng by high-performance liquid chromatography coupled with
from XSTI. Integration of offline 2D LC and high-resolution mass electrospray ionization tandem mass spectrometry, Rapid Commun. Mass
spectrometry is proven as a promising vehicle for elucidation of the Spectrom. 23 (2009) 667–679.
[14] W.Z. Yang, T. Bo, S. Ji, X. Qiao, D.A. Guo, M. Ye, Rapid chemical profiling of
therapeutic basis for herbal medicines together with their prod- saponins in the flower buds of Panax notoginseng by integrating MCI gel
ucts and composite preparations. As the continuous studies, on column chromatography and liquid chromatography/mass spectrometry
one hand, the molecular descriptors mainly responsible for the analysis, Food Chem. 139 (2013) 762–769.
[15] L.L. Wang, L.F. Han, H.S. Yu, M.M. Sang, E.W. Liu, Y. Zhang, S.M. Fang, T. Wang,
observed relationship between structures and retention features
X.M. Gao, Analysis of the constituents in Zhu She Yong Xue Shuan Tong by
will be disclosed, and, on the other hand, the chemical difference ultra high performance liquid chromatography with quadrupole
between XSTI and its homogenous products, such as Xuesaitong, time-of-flight mass spectrometry combined with preparative high
performance liquid chromatography, Molecules 20 (2015) 20518–20537.
will be systematically compared.
[16] X. Qiao, X.H. Lin, S. Ji, Z.X. Zhang, T. Bo, D.A. Guo, M. Ye, Global profiling and
novel structure discovery using multiple neutral loss/precursor ion scanning
combined with substructure recognition and statistical analysis (MNPSS):
Acknowledgements characterization of terpene-conjugated curcuminoids in Curcuma longa as a
case, Anal. Chem. 88 (2016) 703–710.
We are grateful to acknowledge the financial support from [17] Z.X. Yan, G. Lin, Y. Ye, Y.T. Wang, R. Yan, Triterpenoid saponins profiling by
adducts-targeted neutral loss triggered enhanced resolution and product ion
the 56th China Post-doctoral Science Foundation (2014M560364), scanning using triple quadrupole linear ion trap mass spectrometry, Anal.
National Natural Science Foundation of China (81473344 and Chim. Acta 819 (2014) 56–64.
81503240), and National Science and Technology Major Project for [18] Y.L. Song, N. Zhang, S.P. Shi, J. Li, Y.F. Zhao, Q. Zhang, Y. Jiang, P.F. Tu,
Homolog-focused profiling of ginsenosides based on the integration of
Major Drug Development (2014ZX09304-307-001-007). step-wise formate anion-to-deprotonated ion transition screening and
scheduled multiple reaction monitoring, J. Chromatogr. A 1406 (2015)
136–144.
Appendix A. Supplementary data [19] Q.Q. Xing, T. Liang, G.B. Shen, X.L. Wang, Y. Jin, X.M. Liang, Comprehensive
HILIC × RPLC with mass spectrometry detection for the analysis of saponins in
Supplementary data associated with this article can be found, in Panax notoginseng, Analyst 137 (2012) 2239–2249.
[20] C.L. Yao, W.Z. Yang, W.Y. Wu, J. Da, J.J. Hou, J.X. Zhang, Y.H. Zhang, Y. Jin, M.
the online version, at http://dx.doi.org/10.1016/j.jpba.2016.05.035. Yang, B.H. Jiang, X. Liu, D.A. Guo, Simultaneous quantitation of five
notoginseng saponins by multi heart-cutting two-dimensional liquid
chromatography: method development and application to the quality control
References of eight notoginseng containing Chinese patent medicines, J. Chromatogr. A
1402 (2015) 71–81.
[1] H.F. Wu, J. Guo, S.L. Chen, X. Liu, Y. Zhou, X.P. Zhang, X.D. Xu, Recent [21] W.Z. Yang, W. Si, J.X. Zhang, M. Yang, H.Q. Pan, J. Wu, S. Qiu, C.L. Yao, J.J. Hou,
developments in qualitative and quantitative analysis of phytochemical W.Y. Wu, D.A. Guo, Selective and comprehensive characterization of the
constituents and their metabolites using liquid chromatography–mass quinochalcone C-glycoside homologs in Carthamus tinctorius L. by offline
spectrometry, J. Pharm. Biomed. Anal. 72 (2013) 267–291. comprehensive two-dimensional liquid chromatography/LTQ-Orbitrap MS
[2] W. Yang, M. Ye, M. Liu, D.Z. Kong, R. Shi, X.W. Shi, K.R. Zhang, Q. Wang, L.D. coupled with versatile data mining strategies, RSC Adv. 6 (2016) 495–506.
Zhang, A practical strategy for the characterization of coumarins in Radix [22] C.Z. Wang, E. McEntee, S. Wicks, J.A. Wu, C.S. Yuan, Phytochemical and
Glehniae by liquid chromatography coupled with triple quadrupole-linear ion analytical studies of Panax notoginseng (Burk.) F.H. Chen, J. Nat. Med. 60
trap mass spectrometry, J. Chromatogr. A 1217 (2010) 4587–4600. (2006) 97–106.
[3] X.Q. Ma, H.B. Xiao, X.M. Liang, Identification of ginsenosides in Panax [23] H.B. Guo, X.M. Cui, N. An, G.P. Cai, Sanchi ginseng (Panax notoginseng (Burkill)
quinquefolium by LC–MS, Chromatographia 64 (2006) 31–36. F. H. Chen) in China: distribution, cultivation and variations, Genet. Resour.
[4] T.H. Kao, S.C. Huang, B.S. Inbaraj, B.H. Chen, Determination of flavonoids and Crop. Evol. 57 (2010) 453–460.
saponins in Gynostemma pentaphyllum (Thunb.) Makino by liquid [24] https.//hmc.usp.org/monographs/for-comment.
chromatography–mass spectrometry, Anal. Chim. Acta 626 (2008) 200–211. [25] Chinese Pharmacopoeia Commission, Pharmacopoeia of the People’s Republic
[5] Y. Zhao, L.P. Kang, H.S. Yu, J. Zhang, C.Q. Xiong, X. Pang, Y. Gao, C. Liu, B.P. Ma, of China, vol. 1, Chinese Medical Science and Technology Press, Beijing, 2015,
Structure characterization and identification of steroidal saponins from the pp. 11.
rhizomes of Anemarrhena asphodeloides by ultra performance liquid [26] T.B. Ng, Pharmacological activity of sanchi ginseng (Panax notoginseng),
chromatography and hybrid quadrupole time-of-flight mass spectrometry, Pharm. Pharmacol. 58 (2006) 1007–1019.
Int. J. Mass Spectrom. 341–342 (2013) 7–17. [27] W.Z. Yang, Y. Hu, W.Y. Wu, M. Ye, D.A. Guo, Saponins in the genus Panax L.
[6] J. Tchoumtchoua, D. Njamen, J.C. Mbanya, A.L. Skaltsounis, M. Halabalaki, (Araliaceae): A systematic review of their chemical diversity, Phytochemistry
Structure-oriented UHPLC-LTQ Orbitrap-based approach as a dereplicatioin 106 (2014) 7–24.
strategy for the identification of isoflavonoids from Amphimas pterocarpoides [28] Q.F. Gui, Y.M. Yang, S.H. Ying, M.M. Zhang, Xueshuantong improves cerebral
crude extract, J. Mass Spectrom. 48 (2013) 561–575. blood perfusion in elderly patients with lacunar infarction, Neural Regen. Res.
[7] Q. Mao, M. Bai, J.D. Xu, M. Kong, L.Y. Zhu, H. Zhu, Q. Wang, S.L. Li, 8 (2013) 792–801.
Discrimination of leaves of Panax ginseng and P. quinquefolius by ultra high [29] L. Gao, H.P. Zhao, Q. Liu, J.X. Song, C.M. Xu, P. Liu, W. Gong, R.L. Wang, K.J. Liu,
performance liquid chromatography quadrupole/time-of-flight mass Y.M. Luo, Improvement of hematoma absorption and neurological function in
spectrometry based metabolomics approach, J. Pharm. Biomed. Anal. 97 patients with acute intracerebral hemorrhage treated with Xueshuantong, J.
(2014) 129–140. Neurol. Sci. 323 (2012) 236–240.
332 W. Yang et al. / Journal of Pharmaceutical and Biomedical Analysis 128 (2016) 322–332

[30] M. Chen, W.Z. Yang, W.Y. Wu, D.A. Guo, Chemical analysis of xue-shuan-tong [35] Y.M. Liu, X.Y. Xie, Z.M. Guo, Q. Xu, F.F. Zhang, X.M. Liang, Novel
lyophilized powder by LC–MS profiling, Chin. Herbal Med. 7 (2015) 54–61. two-dimensional reversed-phase liquid chromatography/hydrophilic
[31] W. Si, W.Z. Yang, D.A. Guo, J. Da, J.X. Zhang, S. Qiu, C.L. Yao, Y.J. Cui, W.Y. Wu, interaction chromatography, an excellent orthogonal system for practical
Selective ion monitoring of quinochalcone C-glycoside markers for the analysis, J. Chromatogr. A 1208 (2008) 133–140.
simultaneous identification of Carthamus tinctorius L. in eleven Chinese patent [36] J.J. Liu, Y.T. Wang, L. Qiu, Y.Y. Yu, C.M. Wang, Saponins of Panax notoginseng:
medicines by UHPLC/QTOF MS, J. Pharm. Biomed. Anal. 117 (2016) 510–521. chemistry, cellular targets and therapeutic opportunities in cardiovascular
[32] R.S. Plumb, K.A. Johnson, P. Rainville, B.W. Smith, L.D. Wilson, J.M. diseases, Expert Opin. Invest. Drugs 23 (2014) 523–539.
Castro-Perez, J.K. Nicholson, UPLC/MSE , a new approach for generating [37] T.D. Lin, N. Kondo, J. Shoji, Studies on the constituents of Panacis Japonici
molecular fragment information for biomarkers structure elucidation, Rapid Rhizoma. V. The structures of chikusetsusaponin I, Ia Ib, IVa and glycoside P1 ,
Commun. Mass Spectrom. 20 (2006) 1989–1994. Chem. Pharm. Bull. 24 (1976) 253–261.
[33] M. Camenzuli, P.J. Schoenmakers, A new measure of orthogonality for [38] C.Z. Wang, E. McEntee, S. Wicks, J.A. Wu, C.S. Yuan, Phytochemical and
multi-dimensional chromatography, Anal. Chim. Acta 838 (2014) 93–101. analytical studies of Panax notoginseng (Burk.) F.H. Chen, J. Nat. Med. 60
[34] L.W. Qi, C.Z. Wang, C.S. Yuan, Isolation and analysis of ginseng: advances and (2006) 97–106.
challenges, Nat. Prod. Rep. 28 (2011) 467–495.

Das könnte Ihnen auch gefallen