ArticlePDF Available

Hypericin: Source, Determination, Separation, and Properties

Authors:

Abstract and Figures

Hypericin is a naturally occurring compound synthesized by certain species of the genus Hypericum, with various pharmacological effects. It is used as a natural photosensitizing agent with great potential in photodynamic therapy. This review discusses the latest results about the biosynthetic pathways and chemical synthetic routes to obtain hypericin. Although many analysis methods can be used for the determination of hypericin purity, HPLC has become the method of choice due to its fast and sensitive analyses. The extraction and purification of hypericin are also described. Hypericin can be used as a photosensitizer due to a large and active π-electron conjugated system in its structure. Medical applications of hypericin are not easy due to several unsolved practical problems, which include hypericin phototoxicity, poor solubility in water, and extreme sensitivity to light, heat, and pH.
Content may be subject to copyright.
Hypericin: Source, Determination, Separation, and Properties
Jie Zhang
a
, Ling Gao
a
, Jie Hu
a
, Chongjun Wang
a
, Peter-Leon Hagedoorn
b
, Ning Li
a
, and Xing Zhou
c
a
Chongqing Engineering Research Center for Processing, Storage and Transportation of Characterized Agro-Products, College of Environment and
Resources, Chongqing Technology and Business University, Chongqing, China;
b
Department of Biotechnology, Delft University of Technology, Delft,
The Netherlands;
c
Chongqing Academy of Chinese Materia Medica, Chongqing, China
ABSTRACT
Hypericin is a naturally occurring compound synthesized by certain species of the genus Hypericum, with
various pharmacological eects. It is used as a natural photosensitizing agent with great potential in
photodynamic therapy. This review discusses the latest results about the biosynthetic pathways and
chemical synthetic routes to obtain hypericin. Although many analysis methods can be used for the
determination of hypericin purity, HPLC has become the method of choice due to its fast and sensitive
analyses. The extraction and purication of hypericin are also described. Hypericin can be used as
a photosensitizer due to a large and active π-electron conjugated system in its structure. Medical
applications of hypericin are not easy due to several unsolved practical problems, which include hypericin
phototoxicity, poor solubility in water, and extreme sensitivity to light, heat, and pH.
ARTICLE HISTORY
Received 8 February 2020
Revised 4 July 2020
Accepted 7 July 2020
KEYWORDS
Hypericin; synthesis;
extraction; photosensitivity;
solubility; stability
INTRODUCTION
Hypericum or Saint John’s wort, is one of the nine genera
belonging to the Clusiaceae Lindl family widely spread
throughout the world. A large number of Hypericum species,
including Hypericum perforatum L., Hypericum perfoliatum L.,
Hypericum ascyron L., Hypericum androsaemum L., and
Hypericum chinense L., have been identified in Europe, Asia,
North Africa, and North America.
[1]
In China, there are 55
species of Hypericum, 18 of which have been used as local
resources for medicinal purposes in traditional Chinese
medicine.
[2,3]
The plants of the genus Hypericum contain
numerous bioactive substances, such as naphthodianthrones,
flavonoids, phloroglucinols, and polyphenols.
[1,4]
Naphthodianthrones are considered as characteristic constitu-
ents for the identification of Hypericum species
[5]
and one of
the most important kinds of compounds, which includes
hypericin and its biosynthetic precursors: protohypericin,
pseudohypericin, and protopseudohypericin (Figure 1).
Hypericin (4,5,7,4,5,7-hexahydroxy-2,2- dimethylnaphto-
dianthrone, C
30
H
16
O
8
, m.w. 504) is a brownish-black powder
with a unique bitter taste that is mainly found in Hypericum
plants.
[6]
Bucher first discovered that hypericin was an active
ingredient of Hypericum perforatum, and it was renamed
hypericin by Cerny in 1911.
[7]
Hypericin is one out of the most biologically active sub-
stances in the genus Hypericum,
[1,4,8]
and has drawn much
interest in recent years. Evidence of antidepressant properties
has been reported.
[9]
Hypericin was active against chronic
unpredictable mild stress-induced depression and metabolic
dysfunction by affecting excitatory amino acids and monoa-
mine neurotransmitters.
[10]
It also exhibits antitumor activity
as an antineoplastic and photocytotoxic agent, a property
attributed to its photosensitivity.
[1114]
Studies demonstrated
that hypericin possesses immunomodulatory properties and
can induce the production of interferon.
[15,16]
It was found to
be particularly effective as an antiviral agent against the herpes
virus,
[17]
infectious bronchitis virus,
[18]
hepatitis C virus,
[19]
human immunodeficiency virus,
[20]
and novel duck
reovirus.
[21]
Finally, hypericin was considered as an antimicro-
bial agent, antioxidant, and as a promising candidate for
photodynamic diagnosis.
[22,23]
Recently, investigations on the pharmaceutical and clinical
purposes of hypericin surged, and Hypericum perforatum as
a source of hypericin has gradually become one of the three
most popular Chinese herbal medicines.
[24]
The aim of this review
is to describe the recent advances on hypericin research, focusing
on biosynthesis, chemical synthesis, analysis, extraction, purifica-
tion, photosensitivity, solubility in water and stability.
HYPERICIN SOURCES
Natural Sources
Hypericin as a natural bioactive compound can be obtained
from plants, insects, and protozoa.
[3]
It is found in the integu-
ment of Australian Lac insects of the Coccoidea family,
[25,26]
and
the blue-green ciliate, Stentor coerulus, which is a form of
protozoa.
[27]
However, the Hypericum genus has spread
throughout the temperate and tropical regions worldwide, and
is therefore the leading natural source of hypericin. The genus
contains 484 species divided into 36 subgroups.
[28]
The
Hypericum genus has 30 species in Italy
[29]
and 89 in
Turkey.
[5,30]
In China, the 55 various Hypericum species are
widely spread across the country, but the main Hypericum con-
taining areas are concentrated in southwest China. An early
CONTACT Jie Zhang lang19880107@hotmail.com Chongqing Technology and Business University, College of Environment and Resources, Chongqing 400067,
China; Xing Zhou sheya2624@163.com Chongqing Academy of Chinese Materia Medica, Chongqing, China
SEPARATION & PURIFICATION REVIEWS
https://doi.org/10.1080/15422119.2020.1797792
Copyright © Taylor & Francis Group, LLC
survey of circa 200 species of Hypericum indicated that almost all
hypericin-containing species belong to the sections Euhypericum
and Campylosporus of Keller’s classification.
[31]
The most impor-
tant and well-known species is Hypericum perforatum which is
commonly known as St. John’s wort.
[32]
Hypericum perforatum
is a perennial herbaceous plant widely distributed in the world
and it has been included in numerous pharmacopeia. Hypericin
is produced in specialized minute glands on all aerial parts of the
plant, predominantly in flowers and leaves. The hypericin con-
centration varies depending on the species, the geographical
locations of Hypericum,
[1,8]
and the part of the plant,
[30]
as
shown in Table 1. Additionally, the developmental stage of the
plant and seasonal variations also influences the hypericin
concentration.
[30,33]
Although there are numerous other
Hypericum species known to contain approximately similar
amounts of hypericin as Hypericum perforatum,
[34]
information
from the literature on these species is scarce.
Biosynthesis of Hypericin
The biosynthesis of hypericin in Hypericum is more compli-
cated than known chemical synthetic routes and involves the
OOH OH
HO
HO R
OH O OH
OOH OH
HO
HO R
OH O OH
R=CH3 Hypericin
R=CH2OH Pseudohypericin
R=CH3 Protohypericin
R=CH2OH Protopseud ohypericin
Figure 1. Chemical structures of hypericin, protohypericin, pseudohypericin, and protopseudohypericin.
Table 1. The hypericin concentration in some species of the genus Hypericum.
Hypericum species Provenance [ref] Plant part
Hypericin
(mg·g
−1
)
Hypericum perforatum Italy (Sicily)
[1]
Flowering tops 15–20 cm 3.69
Italy (Latium)
[9]
Flowering tops 0.27
Italy (Trentino)
[9]
Flowering tops 0.22
Italy (Tuscany)
[9]
Flowering tops 0.16
Italy (Molise)
[9]
Flowering tops 0.13
Turkey (Samsun)
[5]
Top 1/3 of the crown 2.82
Japan (Tokyo)
[4]
Flowering tops 1.20
China (Hubei)
[89]
Above the ground 1.50
China (Guizhou)
[90]
Above the ground 0.25
Hypericum aviculariifolium Turkey (Gumus)
[5]
Top 1/3 of the crown 2.14
Hypericum aegypticum Italy (Sicily)
[1]
Flowering tops 15–20 cm 0.03
Hypericum enshiense China (Hubei)
[89]
Above the ground 3.00
Hypericum empetrifolium Greece (Crete)
[6]
Above the ground 0.09
Hypericum faberi China (Guizhou)
[90]
Above the ground 0.05
Hypericum hirsutum Italy (Bulgaria)
[6]
Flowers 0.43
Italy (Sicily)
[1]
Flowering tops 15–20 cm 0.15
Italy (Siena)
[9]
Flowering tops 0.02
China (Xinjiang)
[91]
Above the ground 0.06
Serbia (Rudina Planina)
[92]
Above the ground 0.024
Hypericum linarioides Serbia (Rudina Planina)
[92]
Above the ground 0.02
Hypericum lydium Turkey (Havza)
[5]
Top 1/3 of the crown 0.18
Hypericum maculatum Serbia (Rudina Planina)
[92]
Above the ground 0.03
Hypericum montanum Italy (Sicily)
[1]
Flowering tops 15–20 cm 1.42
Hypericum montbretii Turkey (Samsun)
[5]
Top 1/3 of the crown 1.39
Hypericum origanifolium Turkey (Samsun)
[5]
Top 1/3 of the crown 1.43
Hypericum patulum Italy (Sicily)
[1]
Flowering tops 15–20 cm 0.02
Hypericum perfoliatum Turkey (Samsun)
[30]
Top 1/3 of the crown, floral budding 1.06
Turkey (Samsun)
[30]
Top 1/3 of the crown, full flowering 0.96
Turkey (Samsun)
[30]
Top 1/3 of the crown, fresh fruiting 0.41
Italy (Sicily)
[1]
Flowering tops 15–20 cm 0.93
Hypericum pruinatum Turkey (Cumus)
[5]
The top 1/3 of the crown 0.79
Hypericum rumeliacum Serbia (Rudina Planina)
[92]
Above the ground 0.18
Hypericum sampsonii China (Jiangxi)
[90]
Flowering tops 20 cm 0.04
Hypericum scabrum China (Xinjiang)
[91]
Above the ground 0.06
Hypericum tetrapterum Italy (Stia)
[9]
Flowering tops 0.84
Italy (Sicily)
[8]
Flowering tops 15–20 cm 0.40
Serbia (Rudina Planina)
[92]
Above the ground 0.09
Hypericum wightianum China (Guizhou)
[90]
Above the ground 0.023
2J. ZHANG ET AL.
expression of multiple genes. The precise regulation of hyper-
icin biosynthesis remains uncertain until today. The generally
accepted biosynthetic hypericin pathway can be divided into
two main parts: the formation of emodin anthrone and the
conversion of emodin anthrone to hypericin (Figure 2).
[35]
Emodin anthrone is most likely the immediate precursor of
hypericin. Early studies presumed that emodin anthrone
synthesis followed the polyketide pathway.
[36,37]
The cycliza-
tion of a linear polyketide starting with the condensation of
acetyl-CoA and malonyl-CoA results in the formation of emo-
din anthrone catalyzed by polyketide synthase (PKS). Two
cDNAs encoding for PKS designated as HpPKS1 and
HpPKS2 were cloned and identified from Hypericum
perforatum.
[38]
Although the expression of HpPKS2 was corre-
lated with the concentrations of hypericin, the recombinant
HpPKS2 protein failed to convert the acetylated substrates to
emodin or hypericin under in vitro conditions.
[39]
Pillai and
Nair
[40]
provided direct biochemical and molecular evidence in
support of the PKS hypothesis of hypericin biosynthesis in
2014. Auxin inducible culture systems of Hypericum hookeria-
num were applied as a model system to study the metabolic
pathway of hypericin synthesis. The results demonstrated the
presence of additional protein components besides PKS
activity.
In later steps of hypericin biosynthesis, emodin anthrone is
oxidized to emodin by emodin anthrone oxygenase.
[35]
Emodin dianthrone can be produced by a condensation reac-
tion with emodin and emodin anthrone. This subsequently
undergoes oxidation to form protohypericin, then protohyper-
icin produces hypericin on irradiation. Bais, et al.
[41]
discov-
ered that hypericin biosynthesis is related to a gene termed
hyp-1. Based on an in vitro study, the phenolic oxidative
coupling protein (Hyp-1) was shown to catalyze the dimeriza-
tion of emodin and emodin anthrone, the dehydration of the
intermediate to emodin dianthrone, and further phenolic oxi-
dation to protohypericin and hypericin. A high-resolution
crystal structure of the Hyp-1 protein indicated that it has
a pathogenesis-related class 10 protein structure.
[42]
However,
it is unable to dimerize emodin to hypericin using Hyp-1
protein as biocatalyst. Kosuth, et al.
[43]
also found that the
hyp-1 gene is not a limiting factor for hypericin biosynthesis.
The low expression of the genes in the early stages of the
hypericin biosynthetic route may be the potential key factors
in the accumulation and biosynthesis of hypericin.
[44]
The role
of the hyp-1 gene should be further verified by functional
validation experimental approaches. In addition, Kimakova
et al.
[45]
identified new compounds present in the genus
Hypericum and proposed that the anthraquinone skyrin is the
key intermediate in hypericin biosynthesis. Further research on
the role of anthraquinone derivatives in plant metabolism
should be performed.
Chemical Synthesis of Hypericin
In 1957, Brockmann, et al.
[46]
published the first multistep
chemical synthesis method for the production of hypericin.
This synthesis route starts with the reaction of 3,5-dimethox-
ybenzoic acid methyl ester and chloral hydrate. The procedure
with 12 steps is complicated, and the overall yield is 6% - 9%.
Such a low yield and complex synthetic route are no longer
acceptable for the industrial production of hypericin, and new
synthetic pathways have been studied to increase the synthesis
yield and simplify the route. Hypericin can be obtained from
2-methyl anthraquinone in eight steps .
[47]
The key step in this
process is to obtain emodin by steps such as nitration, reduc-
tion, bromination, deamination, and substitution. Emodin is
condensed in the presence of hydroquinone under alkaline
conditions to give protohypericin, which is subsequently
photochemically converted to hypericin by irradiation with
a halogen lamp. Although the synthetic route was optimized
to perform under mild conditions, it involves multiple synth-
esis steps resulting in a low yield.
In 2007, Motoyoshiya, et al.
[48]
proposed a six-step method
for synthesizing hypericin (Figure 3). The regioselective two-
fold Diels-Alder reaction of 1,4-benzoquinone with (1-meth-
oxy-3- methylbuta-1,3-dienyloxy)trimethylsilane results in
7-methyljuglone in the first step. Emodin and its
O-methylated derivative are subsequently produced from
7-methyljuglone and (1,3-dimethoxybuta- 1,3-dienyloxy)tri-
methylsilane. The reduction of both compounds with SnCl
2
in acidic media was accompanied by an acid hydrolysis that
produced emodin anthrone, which after oxidative dimerization
with FeCl
3
hydrate gave bianthrone in high yield. Bianthrone is
oxidized by N-ethyldiisopropylamine (i-Pr
2
EtN) to produce
protohypericin, which is then converted to hypericin by irra-
diation. This shorter route realizes the straightforward synth-
esis of hypericin from a simple compound. However, the yield
in the final step is low and should be improved.
In general, the synthesis of emodin is a necessary step in the
synthesis of hypericin. The synthesis methods for emodin are
Figure 2. The proposed biosynthesis of hypericin. Figure 3. Synthesis of hypericin from 1,4-benzoquinone.
SEPARATION & PURIFICATION REVIEWS 3
numerous and well optimized. Therefore, it has also been
reported that emodin can be used as the starting
material.
[49,50]
After interaction with SnCl
2
, emodin is first con-
verted into emodin anthrone. Then, emodin anthrone is used as
an intermediate reactant. It is reacted with pyridine, piperidine,
pyridine N-oxide and FeSO
4
· 7H
2
O to form hypericin.
[51]
Some
scholars have developed an even more direct method to synthe-
size hypericin. Emodin was converted to hypericin
[52,53]
using
hydroquinone as catalyst under nitrogen and light illumination
after 2 weeks (Figure 4). A large number of scientists have
focused their attention on the synthesis of hypericin, and the
synthesis technology of hypericin has gradually matured.
However, the methods can still be improved further in terms
of atom economy and environmental impact (E-factor).
DETERMINATION AND SEPARATION OF HYPERICIN
Analysis of Hypericin
Various hypericin determination methods have been devel-
oped including: ultraviolet-visible spectroscopy (UV-VIS),
chemiluminescence-flow injection analysis (CL-FIA), thin-
layer chromatography (TLC), and high-performance liquid
chromatography (HPLC).
UV-VIS spectroscopy was applied to determine hypericin
and pseudohypericin in the extracts of Hypericum perforatum,
and the solution of the compounds in methanol was measured
at 588 nm.
[54]
The molar extinction coefficient was difficult to
establish because none of the routine purity criteria can be
applied to hypericin and pseudohypericin. However, specific
reference spectra can successfully be used to analyze hypericin
and pseudohypericin.
Shi
[55]
combined chemiluminescence and flow injection
analysis to establish a CL-FIA method to detect the hypericin
content in Hypericum perforatum. Hypericin has a sensitizing
effect on the chemiluminescence intensity of the Luminol-
KMnO
4
system in an alkaline medium. Under optimized
experimental conditions, the hypericin mass concentration is
linearly related to the luminescence intensity in the range of
1.9 × 10
−5
- 3.8 × 10
−4
g·L
−1
with a limit of detection (LOD) of
3.8 µg·L
−1
. The content of hypericin in Hypericum perforatum
was detected as 0.492 mg·g
−1
. The experimental results also
indicated that CL-FIA and UV methods are equally sensitive.
Moreover, CL-FIA shows some advantages in terms of larger
linear range, higher sensitivity, and higher speed of analysis.
Mulinacci, et al.
[56]
established a TLC-densitometry method
with fluorescence detection to detect the hypericin content in
Hypericum perforatum, and this method was compared with
reversed-phase HPLC-DAD (diode array detection). The
mobile phase was optimized by adjusting the ratio of toluene,
ethyl acetate, and formic acid. The TLC densitometry was
performed without the use of spray or dipping reagents
which improved the speed of the analytical procedure. The
method is cost-effective because of the short analysis time
and the low solvent consumption. The accuracy and reprodu-
cibility of TLC densitometry were comparable with those
obtained by HPLC-DAD. However, HPLC-DAD does provide
much more information than TLC densitometry.
Currently, various HPLC methods for hypericin analysis are
gradually emerging. Wang, et al.
[57]
established an HPLC-
visible spectroscopy method for hypericin determination in
the extracts of Hypericum perforatum. The flow rate of mobile
phase was 1.0 mL·min
−1
(Table 2, A). The linear relationship of
hypericin was good in the range of 6–36 mg·L
−1
(r = 0.9996),
the average recovery rate was 98.86% (n = 6), the relative
standard deviation of the peak area was 3.15%. This HPLC
detection method has a good reproducibility and accuracy and
is suitable to determine hypericin in complex samples. In 2006,
Ruckert, et al.
[58]
presented an HPLC method for the quantita-
tion of hypericin using a new and sensitive amperometric
detection. Using Ag/AgCl as a reference electrode in the detec-
tor, hypericin was eluted isocratically using a mobile phase
consisting of ammonium acetate, methanol, and acetonitrile
(Table 2, B). Hypericin eluted at a retention time of 12 min.
Linearity was obtained over the range 0.035–1.30 mg·L
−1
(r = 0.9994). The LOD of hypericin was 0.010 ng injected on-
column. These parameters show that the method is selective,
simple, rapid, and accurate. Zhang, et al.
[59]
established a high-
resolution HPLC method for determining hypericin by com-
paring different chromatographic conditions. The different
methods showed different detection efficiencies depending on
the mobile phase and detection wavelength. A purchased crude
extract of Hypericum perforatum was applied to optimize the
chromatographic conditions for the determination of hyperi-
cin in a complex sample. The results demonstrated that the best
resolution was obtained at 590 nm and with the mobile phase
composition: methanol/acetonitrile/0.1 mol·L
−1
sodium dihy-
drogen phosphate 200/300/100 v/v/v (Table 2, C). The hyper-
icin calibration curve showed a good linearity in the range of
4–14 mg·L
−1
(r = 0.9986), and the established method is fast
and easy to use.
[59]
In addition, the stationary phase in the column has
a significant effect on the resolution of HPLC. Dolezal, et al.
[60]
selected four stationary phases modified by the pentafluoro-
phenyl group to investigate the contribution of π-π interactions
to the improvement of hypericin separation in comparison to
the separation obtained with a conventional C18 reversed-
phase column. The best analytical method employing the pen-
tafluorophenyl stationary phase (Table 2, E) showed sufficient
linearity, accuracy, and precision and was used for the deter-
mination of hypericin in Hypericum perforatum. Besides the
HPLC methods discussed above, other reported HPLC meth-
ods with different mobile and stationary phases are presented
in Table 2 (F-I). For routine hypericin analyzes, UV detection
at 278 and 284 nm, and VIS detection at 579–590 nm, have
been proposed. However, the VIS detection has a better sensi-
tivity and selectivity than the UV detection.
[58,59,61]
HPLC
Figure 4. The direct synthesis of hypericin with emodin.
4J. ZHANG ET AL.
analyzes with various detection modes will replace the other
method for hypericin analysis because it is fast, accurate, and
sensitive.
[58,60]
Separation of Hypericin
The hypericin content in Hypericum is extremely low, in most
cases below 3 mg·per gram of dry weight of plant material.
Therefore, effective hypericin extraction and purification
methods are needed. The most common method is solvent
extraction using methanol, ethanol, and polar alcohols. In
addition, microwave-assisted extraction has been used to
reduce the extraction time and enzyme-assisted extraction to
increase the extraction yield. The lye (sodium hydroxide)
extraction method has been rarely used because of its higher
energy consumption and longer extraction time. Separation
and purification of hypericin has been performed by using
macroporous resin column chromatography (MRCC) and
molecular imprinting techniques. Additionally, counter-
current chromatography (CCC) has been used widely due to
its high purification rate.
Extraction of Hypericin
Cossuta et al.
[62]
used Soxhlet extraction to extract hypericin
from Hypericum perforatum with four different solvents
(n-hexane, ethyl acetate, 2-propanol, and ethanol), and the
extracts were analyzed by UV-HPLC. Ethanol was the best
solvent to extract hypericin producing a maximum of
0.060 mg·g
−1
. However, the 16 h Soxhlet extraction time was
discouraging needing to be optimized.
Xing
[63]
took a two-step impurity removal method to
extract hypericin. Ether was first used to remove apolar impu-
rities including chlorophyll, and water-soluble impurities
including sugars were removed by suspension in warm water.
After that, ethyl acetate was utilized to extract hypericin. Using
40% and 80% ethanol as eluents, respectively, the extracts were
separated by MRCC. The red MRCC effluent was concentrated
under reduced pressure to obtain a hypericin paste. Finally, the
content of hypericin in the paste, determined by UV-HPLC,
was 1.2%, which far exceeds the 0.3% hypericin content
required for the international market of medicinal products.
This hypericin optimized extraction method has a high
extraction rate and is simple and effective. Moreover, the
MRCC product has the characteristics of low hygroscopicity
and low tackiness.
Punegov, et al.
[64]
studied the extraction of hypericin and
pseudohypericin from raw Hypericum perforatum using
microwave activation. The maximal extraction efficiency was
achieved when 55% ethanol or isopropanol was used as extrac-
tant at 0.0205 W·cm
−3
microwave irradiation power density
and a microwave frequency of 2450 MHz for 60 s. Microwave
activation was found to improve 10-fold the extraction effi-
ciency, reducing the time necessary to fully extract hypericin
and pseudohypericin compared to classical extraction meth-
ods. Zhang, Feng, Xu, Tan, Hagedoorn, and Ding
[59]
used
a xylanase-assisted associated with a microwave-assisted
extraction to improve the hypericin extraction. This method
was found to improve the extraction yield of hypericin signifi-
cantly compared to unassisted extraction. Microwave-assisted
extraction after xylanase-assisted extraction was found to be
the most efficient strategy for extracting hypericin. The yield
was 0.32 ± 0.006 mg·g
−1
, which was a 210% increase over
unassisted extraction. Compared to conventional solvent
extraction, enzyme-assisted extraction can be accomplished
using a low enzyme concentration. It reduces energy consump-
tion due to less wastewater generation and lower temperatures.
In addition, the unique microwave heating method decreases
the extraction time and increases the yield even further. In
conclusion, enzyme and microwave assistance are effective
strategies to improve the mass transfer rate of hypericin during
the extraction process. Scalability of the equipment has to be
addressed in order to use microwave assistance on an industrial
scale.
Purification of Hypericin
The purity of hypericin in the extracts is not sufficient to meet
the requirements for pharmaceutical products. Further
research on how to purify hypericin is necessary. Xue, et al.
[65]
selected six macroporous adsorption resins (D101, AB-8, HZ-
801, HZ-818, HZ-806, X-5) with different specific surface areas
and pore sizes to separate hypericin. The crude extracts of
Hypericum perforatum were used to investigate the hypericin
separation and purification performance of six resins. Optimal
conditions were determined using a medium operating
Table 2. Different chromatographic conditions for determining hypericin.
Methods
[ref]
Mobile phase
v:v:v Stationary phase
Temperature
(ºC)
Detection
wavelength
(nm)
A
[57]
Acetonitrile, 0.02 M sodium dihydrogen phosphate, volume ratio of 85: 15 v:v ODS-A 25 588
B
[58]
Ammonium acetate, methanol and acetonitrile, volume ratio of 10: 40: 50 LiChroCart
Purospher RP18e
22 254
C
[59]
Methanol, acetonitrile, sodium dihydrogen phosphate solution, volume ratio of 200: 300: 100 v:v:v Kromasil C18 40 590
D
[59]
Methanol, acetonitrile, sodium dihydrogen phosphate solution, volume ratio of 200: 300: 100 v:v:v Kromasil C18 40 284
E
[60]
Solvent A: H
2
O, 0.1 M acetic acid, 0.1 M trimethylamine, solvent B: acetonitrile, 0.1 M acetic acid,
0.1 M trimethylamine
0–1 min (A: B = 15: 85), 1–8 min (A: B = 0: 100), 8.0–10.0 min (A: B = 15: 85).
Pentafluorophenyl 25 278
F
[93]
Methanol, ethyl acetate, 0.1 M Na
2
PO
4
, volume ratio of 72: 23: 5 v:v:v Ultracarb 7 ODS 23 579
G
[94]
Ethylacetate, 15.6 g L
−1
sodium dihydrogen phosphate (pH = 2 with phosphoric acid) and
methanol, volume ratio of 39: 41: 160 v:v:v
ACE C18 40 590
H
[95]
Methanol, acetonitrile, water and 3% aqueous phosphoric acid, volume ratio of 45: 50: 4.5: 0.5 with
triethylamine adjusted to pH = 6
Supelcosil ODS 25 590
I
[96]
Methanol, ethyl acetate and phosphate buffer (pH = 2, 0.1 M), volume ratio of 60: 20: 20 v:v:v VP-ODS C18 25 590
SEPARATION & PURIFICATION REVIEWS 5
pressure separation method. The results showed that HZ-801
resin is the preferred separation material with a high adsorp-
tion rate and a high elution rate, and the purity of hypericin
after elution with ethanol solution was 79%. Hypericin separa-
tion using MRCC is feasible on industrial scale due to its easy
operation and recyclability. However, contaminating sub-
stances which are similar in structure to hypericin affect the
results of hypericin separation due to the physical adsorption
principle of the resins.
Molecular imprinting is a newly established method for
chemical separation and purification in recent years. The syn-
thetic molecularly imprinted polymer used for separation
shows high selectivity and specificity for the template (target)
molecule. A core-shell structure molecularly imprinted mag-
netic nanospheres of hypericin (Fe
3
O
4
@MIPs) were prepared
by mercapto-alkyne click polymerization.
[66]
The Fe
3
O
4
@MIPs
showed a good adsorption capacity of 3.43 mg·g
−1
, high fast
mass transfer rates, and good reusability. Hypericin, acryla-
mide, and pentaerythritol triacrylate were used as a template
molecule, functional monomer, and molecular imprint pre-
assembly cross-linker, respectively.
[67]
A cooperative hydro-
gen-bonding complex between hypericin and acrylamide was
formed at the ratio of 1:6 in the prepolymerized system. A high
recovery of 82.3% was achieved by molecular-imprinted poly-
mers to extract hypericin from Hypericum perforatum extracts.
Molecular imprinting is simple, rapid, accurate, and reliable.
The main disadvantage is the amount of pure hypericin which
has to be used to prepare the molecular-imprinted polymers.
Cao, et al.
[68]
found that CCC combined with pre-separation
by ultrasonic solvent extraction was successful for the separation
of series of bioactive compounds from the crude extracts of
Hypericum perforatum. The ethyl acetate extract was separated
by using the solvent system hexane-ethylacetate- methanol-
water (1:1: 1:1 and 1:3: 1:3) in gradient through both reverse
phase and normal phase elution mode. The hypericin purity was
determined to be 95% by HPLC-DAD. CCC does not require
a solid carrier, and the hypericin obtained in less than 5 hours by
this method has high purity. However, CCC is difficult to realize
on an industrial scale and it is not environmentally friendly due
to the amount of organic solvents used.
The extraction and purification technologies that have been
developed provide an indispensable foundation to prepare
hypericin products with different purities. Further studies are
required to achieve a method that combines low cost and
a green process. Furthermore, preparative liquid chromatogra-
phy is a feasible technology to prepare a high purity product,
can be utilized to separate hypericin.
[6971]
This technology is
also amenable to industrial scale-up.
PROPERTIES OF HYPERICIN
Photosensitivity
Hypericin is one of the most effective natural photosensitizers,
and shows a good photosensitivity due to the extensive elec-
tron-conjugated system in its structure (Figure 1). Hypericin is
extremely sensitive to light and it is photoactivated to produce
peroxides. The photosensitive mechanism of hypericin has
been described as follows: in the photodynamic action, light
quanta are absorbed by the sensitizer, generating the excited
singlet state. This excited singlet state may undergo intersystem
crossing to the triplet excited state. The triplet sensitizer will
excite singlet oxygen that is produced when the energy is
transferred to ground state triplet oxygen, and the singlet oxy-
gen forms a peroxide subsequently.
[72]
Moreover, it has been
confirmed that the photosensitivity of hypericin induces cell
apoptosis and inhibits the growth of cancer cells. This optical
activity of hypericin has been widely used in optical
diagnostics.
[73]
Hypericin can produce superoxide-free radicals under the
irradiation of visible light and in the presence of oxygen.
[74]
Hypericin shows an electron paramagnetic resonance (EPR)
signal caused by a semiquinone-like radical formed by inter-
molecular electron transfer in the absence of light and electron
donors. The amplitude of the radical EPR signal for the water-
dispersed lysine salt of hypericin is significantly increased
under visible light irradiation. This finding indicates that the
free hypericin radicals and the superoxide radicals are formed
during light irradiation, and may also be implicated in the
biological activities.
The photosensitivity of hypericin is commonly used in
photodynamic therapy, providing an effective treatment of can-
cer. Rabbits and mice xenografted with P3 human squamous cell
carcinoma were used to assess the usefulness of hypericin for
laser photoinactivation of solid tumors.
[75]
The tissue uptake and
distribution of hypericin in rabbits and mice were measured.
The degree of absorption of hypericin by intravenous injection at
4 and 24 hours in both animal tissues was determined by ethanol
extraction and quantitative fluorescence spectrophotometry.
Experimental results show that elimination of hypericin was
rapid in most animal organs with residual hypericin under
10% of the maximum after 7 days. The retention rate of squa-
mous cell tumors is only 25% to 30%. It indicated that photo-
dynamic therapy using hypericin can, to a certain extent,
eliminate some cancer cells.
[76]
In addition, hypericin and laser
irradiation induced cell death mediated by the intracellular
reactive oxygen species and mitochondrial damage. These data
demonstrate that hypericin is an effective photosensitizer with
potential for human cancer therapy.
Hypericin can absorb light in the ultraviolet and visible
range. Although hypericin has potential for cancer treatment,
some studies have shown that hypericin is phototoxic to the
skin and human eye. Ingestion of hypericin containing drugs is
potentially phototoxic to the retina, which may lead to retinal
or early macular degeneration.
[77]
In addition, hypericin is not
cytotoxic in the dark.
[78]
Further research into the practical
application of hypericin in photodynamic therapy should be
performed.
Water Solubility of Hypericin
Hypericin is readily dissolved in dimethylsulfoxide, methanol,
ethanol, and alkaline aqueous solution, and it is red-colored at
pH <11.5 and green-colored above pH 11.5.
[79]
However, it is
a serious drawback that hypericin exhibits low level of solubi-
lity in neutral water because of its hydrophobicity.
[80,81]
Hypericin forms nonsoluble aggregates in an aqueous
environment.
6J. ZHANG ET AL.
Fluorescence spectroscopy and diffusion coefficient mea-
surements were used to investigate the self-association of
hypericin molecules in DMSO/water mixtures.
[82]
Fluorescence measurements revealed that hypericin remained
in its monomeric form in DMSO/water mixtures containing up
to 20% - 30% water. As the proportion of water passes 30%,
hypericin gradually formed non-fluorescent aggregates, and
the size of the aggregates increased with increasing water con-
centrations. Hypericin presumably produces large molecular
weight stacked aggregates in a neutral aqueous environment.
In addition, molecules of hypericin remain in the monomeric
state in an aqueous environment at alkaline pH.
The insoluble hypericin aggregates in aqueous solutions do
not possess biological activity. This characteristic restricts
hypericin applications in medicine. Polymeric micelles made
with polyethylene glycol (PEG) have been utilized to improve
the solubility of hypericin.
[83]
PEGs with low molecular weight
(<1000 g·mol
−1
) did not significantly contribute to the hyper-
icin solubilization. However, PEGs with molecular weight
>2000 g·mol
−1
efficiently transformed hypericin aggregates to
the monomeric state. The solubility of hypericin in water
increased significantly by adding cromolyn disodium salt
(DSCG).
[81]
The monomerization of hypericin under these
conditions can be explained as a result of the hydrotropic effect
of DSCG. This hydrotropic effect is most likely a result of
interactions between the two relative rigid aromatic rings of
DSCG and a delocalized charge on the surface of the hypericin
molecule. Kubin, et al.
[84]
prepared a non-covalently bound
hypericin-polyvinylpyrrolidone (PVP) complex to enhance the
water solubility of hypericin. The hypericin-PVP complex
bound more than 1000 mg of hypericin in presence of 100 g
PVP and the resulting complex was soluble in 1 L of pure
water. The proposed methods provide strategies to improve
the solubility of hypericin in water, which will facilitate medical
applications. The effects of these additives on the biological
activities of the resulting hypericin preparations have to be
investigated.
Hypericin Stability
Due to its photosensitivity, light inevitably affects the stability
of hypericin. Therefore, exposure to light has a significant
impact on the biological activities of hypericin. The stability
of hypericin in extract solutions of Hypericum perforatum and
standard solutions has been evaluated under different light
conditions monitored by HPLC-VIS.
[85]
Hypericin was extre-
mely unstable after exposure to light, and light was the main
factor reducing the effective hypericin concentration.
Additionally, temperature was a factor affecting hypericin sta-
bility. Wang, et al.
[86]
investigated the effects of light and
temperature on the long-term stability of hypericin extracts.
After 8 weeks, the content of hypericin in the extracts
decreased by 49% under constant light at room temperature,
and it declined by only 8.5% in the dark. The hypericin content
of the extracts was unchanged under dark and low temperature
(−24°C) conditions. Long-term storage is possible for hyper-
icin dissolved in a polar solvent, under a nitrogen atmosphere
at freezing temperature (<-30°C).
[87]
Wang and Zhang
[88]
stu-
died the effects of visible light, temperature, pH, Na
2
SO
3
and
ascorbic acid on the stability of hypericin by UV-VIS spectro-
scopy. The results indicated that light is the major factor
influencing the stability of hypericin. Light and temperature
were found to have a greater effect on stability under alkaline
conditions than acidic conditions. Therefore, alkaline solvents
should be avoided during the extraction of hypericin. However,
the stability of hypericin improved when ascorbic acid or Na
2
SO
3
was added. The instability of hypericin has always been
a major challenge in the separation and purification process.
CONCLUSIONS
Hypericin is one of the effective bioactive substances primarily
extracted from the Hypericum plants, and has various pharma-
cological activities such as anti-depressive, anti-tumor, and
anti-viral. The biosynthetic pathways in the plants are known,
but the precise regulation of these pathways remains uncertain.
Chemical synthesis routes for hypericin from different starting
compounds have been developed. However, a novel synthesis
route combining a high overall yield, low cost, and less envir-
onmental pollution is still desired. HPLC is widely used for
hypericin analysis, and it may replace all other analytical
methods due to its fast analysis times and high sensitivity.
Extraction of hypericin from Hypericum can provide low pur-
ity products or extracts. Microwave-assisted extraction and
enzyme-assisted extraction contribute to a higher hypericin
yield. Macroporous adsorption resin, molecular imprinting
techniques, CCC, and preparative liquid chromatography sys-
tems were used to prepare high purity hypericin. Especially,
preparative liquid chromatography system is a feasible strategy
to realize the industrial production of high purity hypericin.
Hypericin is photosensitive due to its extensive system of
conjugated C = C double bonds. It can be utilized for photo-
dynamic therapy. However, the phototoxicity of hypericin to
the skin and human lens should be considered. Hydrophobic
groups in hypericin account for its low solubility in water.
PEG, DSCG, and PVP can significantly improve the solubility.
In addition, the storage and operation conditions, such as light,
temperature, and pH will affect the hypericin stability. Without
a doubt, the disadvantages of hypericin can be overcome with
technical solutions, which are worthwhile of investigation
because its great medicinal value.
Funding
The study has been carried out with financial support from the Natural
Science Foundation Project of CSTC [No. cstc2017shms-xdny100003];
Project of China Scholarship Council [No. 201808500035].
ORCID
Jie Zhang http://orcid.org/0000-0002-3938-0878
REFERENCES
[1] Napoli, E.; Siracusa, L.; Ruberto, G.; Carrubba, A.; Lazzara, S.;
Speciale, A.; Cimino, F.; Saija, A.; Cristani, M. Phytochemical
Profiles, Phototoxic and Antioxidant Properties of Eleven
Hypericum species-A Comparative Study. Phytochemistry. 2018,
152, 162–173. DOI: 10.1016/j.phytochem.2018.05.003.
SEPARATION & PURIFICATION REVIEWS 7
[2] Pan, W.; Zhang, S.; He, X. Y.; Wang, H. F.; Bian, S. Research Status
of Wild Hypericum Plant. Hortic. Seed. 2014, 11, 40–41. DOI:
10.3969/j..2095-0896.2014.11.014.
[3] Huang, L. F.; Cheng, S. L. Hypericin in Hypericum: Chemistry,
Botanical Sources and Biological Activities. J. Pharm. Sci. 2012, 21
(5), 388–400. DOI: 10.5246/jcps.2012.05.052.
[4] Onoue, S.; Seto, Y.; Ochi, R. I.; Hideyuki, I.; Tsutomu, H. B.;
Shizuo, Y. In Vitro Photochemical and Phototoxicological
Characterization of Major Constituents in St. John’s Wort
(Hypericum Perforatum) Extracts. Phytochemistry. 2011, 50(14–-
15), 1814–1820. DOI: 10.1016/j.phytochem.2011.06.011.
[5] Cirak, C.; Radusiene, J.; Janullis, V.; Lvanauskas, L.; Arslan, B.
Chemical Constituents of Some Hypericum Species Growing in
Turkey. J. Plant. Biol. 2007, 50(6), 632–635. DOI: 10.1007/
bf03030606.
[6] Kitanov, G. M.;. Hypericin and Pseudohypericin in Some
Hypericum Species. Biochem. Syst. Ecol. 2001, 29(2), 171–178.
DOI: 10.1016/s0305-1978(00)00032-6.
[7] Brockmann, H.; Sanne, W. Pseudo-hypericin, ein neuer, roter
Hypericum farbsto. Sci. Nat. 1953, 40(17), 461. DOI: 10.1007/
bf00628841.
[8] Lazzara, S.; Carrubba, A.; Napoli, E. Variability of Hypericins and
Hyperforin in Hypericum Species from the Sicilian Flora. Chem.
Biodivers. 2020, 17(1), 1–33. DOI: 10.1002/cbdv.201900596.
[9] Lazzara, S.; Carrubba, A.; Napoli, E. Hypericum Spp.: A Resource
from Wild Mediterranean Flora for the Treatment of Mild
Depression. In Bioactive Phytochemicals: Perspectives for Modern
Medicine; Gupta, V. K., Ed.; New Delhi: Daya Publishing House,
2015; Vol. 3, pp 337–352.
[10] Zhai, X. J.; Chen, F.; Chen, C.; Zhu, C. R.; Lu, Y. N. LC-MS/MS
Based Studies on the Anti-depressant Effect of Hypericin in the
Chronic Unpredictable Mild Stress Rat Model. J. Ethnopharmacol.
2015, 169, 363–369. DOI: 10.1016/j.jep.2015.04.053.
[11] Stanicova, J.; Verebova, V.; Benes, J. Interaction of a Potential
Anticancer Agent Hypericin and Its Model Compound Emodin
with DNA and Bovine Serum Albumin. In Vivo. 2018, 32(5),
1063–1070. DOI: 10.21873/invivo.11347.
[12] Han, C.; Zhang, C.; Ma, T.; Zhang, C.; Luo, J. G.; Xu, X.; Zhao, H. J.;
Chen, Y.; Kong, L. Y. Hypericin-functionalized Graphene Oxide
for Enhanced Mitochondria-targeting and Synergistic Anticancer
Effect. Acta. Biomater. 2018, 77, 268–281. DOI: 10.1016/j.
actbio.2018.07.018.
[13] Misuth, M.; Horvath, D.; Miskovsky, P.; Huntosova, V. Synergism
between PKC Delta Regulators Hypericin and Rottlerin Enhances
Apoptosis in U87 MG Glioma Cells after Light Stimulation.
Photodiagnosis. Photodyn. Ther. 2017, 18, 267–274. DOI: 10.1016/
j.pdpdt.2017.03.018.
[14] Xu, Y. X.; Wang, D. X.; Zhuang, Z. Z.; Jin, K. K.; Zheng, L. Z.;
Yang, Q.; Guo, K. Y. Hypericin-mediated Photodynamic Therapy
Induces Apoptosis in K562 Human Leukemia Cells through JNK
Pathway Modulation. Mol. Med. Report. 2015, 12(5), 6475–6482.
DOI: 10.3892/mmr.2015.4258.
[15] Barathan, M.; Mariappan, V.; Shankar, E. M.; Abdullah, B. J. J.;
Goh, K. L.; Vadivelu, J. Hypericin-photodynamic Therapy Leads to
Interleukin-6 Secretion by HepG2 Cells and Their Apoptosis via
Recruitment of BH3 Interacting-domain Death Agonist and
Caspases. Cell. Death. Dis. 2013, 4, 1–10. DOI: 10.1038/cddis.2013.219.
[16] Koon, H. K.; Lo, K. W.; Leung, K. N.; Lung, M. L.; Chang, C. C.;
Wong, R. N.; Leung, W. N.; Mak, N. K. Photodynamic
Therapy-mediated Modulation of Inflammatory Cytokine
Production by Epstein-Barr Virus-infected Nasopharyngeal
Carcinoma Cells. Cell. Mol. Immunol. 2010, 7(4), 323–326. DOI:
10.1038/cmi.2010.4.
[17] Cohen, P. A.; Hudson, J. B.; Towers, G. H. N. Antiviral Activities of
Anthraquinones, Bianthrones and Hypericin Derivatives from
Lichens. Experientia. 1996, 52(2), 180–183. DOI: 10.1007/
bf01923366.
[18] Chen, H. J.; Feng, R.; Muhammad, I.; Abbas, G.; Zhang, Y.;
Ren, Y. D.; Huang, X. D.; Zhang, R. L.; Diao, L.; Wang, X. R.
Protective Effects of Hypericin against Infectious Bronchitis
Virus Induced Apoptosis and Reactive Oxygen Species in
Chicken Embryo Kidney Cells. Poult. Sci. 2019, 98(12),
6367–6377. DOI: 10.3382/ps/pez465.
[19] Shih, C. M.; Wu, C. H.; Wu, W. J.; Hsiao, Y. M.; Ko, J. L. Hypericin
Inhibits Hepatitis C Virus Replication via Deacetylation and
Down-regulation of Heme Oxygenase-1. Phytomedicine. 2018, 46,
193–198. DOI: 10.1016/j.phymed.2017.08.009.
[20] Xu, Y. M.; Lu, C. Z. Raman Spectroscopic Study on Structure of
Human Immunodeficiency Virus (HIV) and Hypericin-induced
Photosensitive Damage of HIV. Sci. China Ser. C-Life Sci. 2005,
48(2), 117–132. DOI: 10.1360/04yc0015.
[21] Du, X. X.; Xiao, R.; Fu, H. L.; Yuan, Z. X.; Zhang, W.; Yin, L. Z.;
He, C. L.; Li, C. F.; Zhou, J. W.; Liu, G. Q.; et al. Hypericin-loaded
Graphene Oxide Protects Ducks against a Novel Duck Reovirus.
Mater. Sci. Eng. C-Mater. Biol. Appl. 2019, 105, 1–8. DOI: 10.1016/
j.msec.2019.110052.
[22] Penjweini, R.; Loew, H. G.; Eisenbauer, M.; Kratky, K. W.
Modifying Excitation Light Dose of Novel Photosensitizer
PVP-hypericin for Photodynamic Diagnosis and Therapy.
J. Photochem. Photobiol. B-Biol. 2013, 120, 120–129. DOI:
10.1016/j.jphotobiol.2012.12.013.
[23] Noell, S.; Feigl, G. C.; Serifi, D.; Mayer, D.; Naumann, U.;
Gobel, W.; Ehrhardt, A.; Ritz, R. Microendoscopy for Hypericin
Fluorescence Tumor Diagnosis in a Subcutaneous Glioma Mouse
Model. Photodiagnosis. Photodyn. Ther. 2013, 10(4), 552–560. DOI:
10.1016/j.pdpdt.2013.06.001.
[24] Wang, Y. C.; Sun, Y.; Zhou, Y. F.; Li, Q.; Li, Y. Research Status and
Existing Problems of Medicinal Plant Hypericum Perforatum. GX.
Forest. Sci. 2017, 46(4), 403–405. DOI: 10.3969/j..1006-
1126.2017.04.013.
[25] Cameron, D. W.; Raverty, W. D. Pseudohypericin and Other
Phenanthroperylene Quinones. Aust. J. Chem. 1976, 29(7),
1523–1533. DOI: 10.1071/ch9761523.
[26] Banks, H. J.; Cameron, D. W.; Raverty, W. D. Chemistry of the
Coccoidea. II. Condensed Polycyclic Pigments from Two
Australian Pseudococcids (Hemiptera). Aust. J. Chem. 1976, 29
(7), 1509–1521. DOI: 10.1071/CH9761509.
[27] Walker, E. B.; Lee, T. Y.; Song, P. S. Spectroscopic Characterization
of the Stentor Photoreceptor. Biochim. Biophys. Acta. 1979, 587(1),
129–144. DOI: 10.1016/0304-4165(79)90227-7.
[28] Crockett, S. L.; Robson, N. K. B. Taxonomy and Chemotaxonomy
of the Genus Hypericum. Med. Aromat. Plant. Sc. Biotechnol. 2011,
5(1), 1–13.
[29] Conti, F.; Abbate, G.; Alessandrini, A.; Blasi, C. An Annotated
Checklist of the Italian Vascular Flora; Palombi Editori: Roma,
2005.
[30] Cirak, C.; Radusiene, J.; Janulis, V.; Ivanauskas, L. Secondary
Metabolites in Hypericum Perfoliatum: Variation among Plant
Parts and Phenological Stages. Bot. Helv. 2007, 117(1), 29–36.
DOI: 10.1007/s00035-007-0777-z.
[31] Mathis, C.; Ourisson, G. Etude chimio-taxonomique du genre
Hypericum I. Repartition de lhypericine. Phytochemistry. 1963, 2
(2), 157–171. DOI: 10.1016/s0031-9422(00)82976-3.
[32] Ozturk, N.; Tuncel, M.; Erkara, I. P. Phenolic Compounds and
Antioxidant Activities of Some Hypericum Species: A Comparative
Study with H. Perforatum. Pharm. Biol. 2009, 47(2), 120–127. DOI:
10.1080/13880200802437073.
[33] Wang, D.; Bai, J.; Sun, F.; Yang, D. Chemical Constituents and
Antidepressant Activity of the New Species Hypericum Enshiense
Occurring in China. Phytomedicine. 2010, 17(6), 410–413. DOI:
10.1016/j.phymed.2009.07.015.
[34] Zheng, Q. M.; Qin, L. P.; Zheng, H. C.; Chen, Y.; Zhang, C.;
Zhang, Q. Y.; Han, T.; Guo, C. Quantitative Phytochemical
Analysis of 11 Hyperium Species Growing in China. Acad. J. Sec.
Mil. Med. Univ. 2003, 24(4), 457–459. DOI: 10.16781/j.0258-
879x.2003.04.039.
[35] Xiong, Y. J.; Gurina, S.; Jia, X. G.; Murati, K.; Shen, S. K.
Determination of Hypericin in Three Species of Hypericum from
Xinjiang. Chin. Tradit. Herb. Drugs. 2003, 9, 99–103. DOI: 10.3321/
j.0253-2670.2003.09.043.
8J. ZHANG ET AL.
[36] Smelcerovic, A.; Verma, V.; Spiteller, M.; Ahmad, S. M.; Puri, S. C.;
Qazi, G. N. Phytochemical Analysis and Genetic Characterization
of Six Hypericum Species from Serbia. Phytochemistry. 2006, 67(2),
171–177. DOI: 10.1016/j.phytochem.2005.10.021.
[37] Cellarova, E.; Daxnerova, Z.; Kimakova, K.; Haluskova, J. The
Variability of the Hypericin Content in the Regenerants of
Hypericum Perforatum. Acta. Biotechnol. 1994, 14(3), 267–274.
DOI: 10.1002/abio.370140309.
[38] Stojanovic, G.; Dordevic, A.; Smelcerovic, A. Do Other Hypericum
Species Have Medical Potential as St. John’s Wort (Hypericum
Perforatum)? Curr. Med. Chem. 2013, 20(18), 2273–2295. DOI:
10.2174/0929867311320180001.
[39] Huang, L. F.; Wang, Z. H.; Chen, S. L. Hypericin: Chemical
Synthesis and Biosynthesis. Chin. J. Nat. Med. 2014, 12(2), 81–88.
DOI: 10.1016/s1875-5364(14)60014-5.
[40] Kirakosyan, A.; Sirvent, T. M.; Gibson, D. M.; Kaufman, P. B. The
Production of Hypericins and Hyperforin by in Vitro Cultures of
St. John’s Wort (Hypericum Perforatum). Biotechnol. Appl.
Biochem. 2004, 39, 71–81. DOI: 10.1042/ba20030144.
[41] Zobayed, S. M. A.; Afreen, F.; Kozai, T. Temperature Stress Can
Alter the Photosynthetic Efficiency and Secondary Metabolite
Concentrations in St. John’s Wort. Plant. Physiol. Biochem. 2005,
43(10–11), 977–984. DOI: 10.1016/j.plaphy.2005.07.013.
[42] Karppinen, K.; Hohtola, A. Molecular Cloning and Tissue-specific
Expression of Two cDNAs Encoding Polyketide Synthases from
Hypericum Perforatum. J. Plant. Physiol. 2008, 165(10), 1079–1086.
DOI: 10.1016/j.jplph.2007.04.008.
[43] Karppinen, K. Biosynthesis of Hypericins and Hyperforins in
Hypericum Perforatum L. (St. John’s Wort) - Precursors and
Genes Involved. Ph. D. Dissertation, University of Oulu, Oulu,
Finland, 2010.
[44] Pillai, P. P.; Nair, A. R. Hypericin Biosynthesis in Hypericum
Hookerianum Wight and Arn: Investigation on Biochemical
Pathways Using Metabolite Inhibitors and Suppression
Subtractive Hybridization. C. R. Biol. 2014, 337(10), 571–580.
DOI: 10.1016/j.crvi.2014.08.002.
[45] Bais, H. P.; Vepachedu, R.; Lawrence, C. B.; Stermitz, F. R.;
Vivanco, J. M. Molecular and Biochemical Characterization of an
Enzyme Responsible for the Formation of Hypericin in St. John’s
Wort (Hypericum Perforatum L.). J. Biol. Chem. 2003, 278(34),
32413–32422. DOI: 10.1074/jbc.M301681200.
[46] Michalska, K.; Fernandes, H.; Sikorski, M.; Jaskolski, M. Crystal
Structure of Hyp-1, a St. John’s Wort Protein Implicated in the
Biosynthesis of Hypericin. J. Struct. Biol. 2010, 169(2), 161–171.
DOI: 10.1016/j.jsb.2009.10.008.
[47] Kosuth, J.; Smelcerovic, A.; Borsch, T.; Zuehlke, S.; Karppinen, K.;
Spiteller, M.; Hohtola, A.; Cellarova, E. The Hyp-1 Gene Is Not
a Limiting Factor for Hypericin Biosynthesis in the Genus
Hypericum. Funct. Plant. Biol. 2011, 38(1), 35–43. DOI: 10.1071/
fp10144.
[48] Kosuth, J.; Hrehorova, D.; Jaskolski, M.; Cellarova, E. Stress-
induced Expression and Structure of the Putative Gene Hyp-1 for
Hypericin Biosynthesis. Plant. Cell. Tiss. Organ. Cult. 2013, 114(2),
207–216. DOI: 10.1007/s11240-013-0316-0.
[49] Kimakova, K.; Kimakova, A.; Idkowiak, J.; Stobiecki, M.;
Rodziewicz, P.; Marczak, L.; Cellarova, E. Phenotyping the Genus
Hypericum by Secondary Metabolite Profiling: Emodin Vs. Skyrin,
Two Possible Key Intermediates in Hypericin Biosynthesis. Anal.
Bioanal. Chem. 2018, 410(29), 7689–7699. DOI: 10.1007/s00216-
018-1384-0.
[50] Brockmann, H.; Kluge, F.; Muxfeldt, H. Total Synthese Des
Hypericins. Chem. Ber. 1957, 90, 2302–2311. DOI: 10.1002/
cber.19570901027.
[51] Kim, S. W.; Park, J. H.; Yang, S. D.; Hur, M. G.; Kim, Y. S.;
Chai, J. S.; Yu, K. H. Facile Synthesis and Radioiodine Labeling of
Hypericin. J. Cheminform. 2005, 36(3), 1147–1150. DOI: 10.1002/
chin.200503214.
[52] Motoyoshiya, J.; Masue, Y.; Nishi, Y.; Aoyama, H. Synthesis of
Hypericin via Emodin Anthrone Derived from a Two-fold
Diels-alder Reaction of 1,4-benzoquinone. Nat. Prod. Commun.
2007, 2(1), 67–70. DOI: 10.1177/1934578x0700200113.
[53] Falk, H.; Meyer, J.; Oberreiter, M. A Convenient Semisynthetic
Route to Hypericin. Monatsh. Chem. 1993, 124(3), 339–341. DOI:
10.1007/bf00810594.
[54] Falk, H.; Sarhan, A.-E.-W. A. O.; Tran, H. T. N.; Altmann, R.
Synthesis and Properties of Hypericins Substituted with Acidic
and Basic Residues: Hypericin Tetrasulfonic acid–A Water
Soluble Hypericin Derivative. Monatsh. Chem. 1998, 129(3),
309–318. DOI: 10.1007/pl00000089.
[55] Falk, H.; Schoppel, G. A Synthesis of Emodin Anthrone. Monatsh.
Chem. 1991, 23(5), 739. DOI: 10.1002/chin.199205246.
[56] Rodewald, G.; Arnold, R.; Griesler, J.; Steglich, W. Synthesis of
Hypericin and Related Meso-naphthodianthrones by Alkaline
Dimerization of Hydroxyanthraquinones. Angew. Chem-Int. Edit.
Engl. 1977, 16(1), 46–47. DOI: 10.1002/anie.197700462.
[57] Spitzner, D. Synthesis of Proto-hypericin from Emodin. Angew.
Chem-Int. Edit. Engl. 1977, 16(1), 46. DOI: 10.1002/
anie.197700461.
[58] Nait-Si, Y.; Fourneron, J. D. Hypericin and Pseudohypericin.
Purity Criteria and Quantitative Determination in Extracts of
St. John’s Wort (Hypericum Perforatum). Monatsh. Chem. 2004,
135(10), 1319–1326. DOI: 10.1007/s00706-004-0207-5.
[59] Shi, H. Y. Study on the Application of Flow
Injection-chemiluminescence Method in the Analysis of Active
Components of Hypericum Perforatum L. M. D. Dissertation,
Shaanxi University of Technology, Shaanxi, China, 2015.
[60] Mulinacci, N.; Bardazzi, C.; Romani, A.; Pinelli, P.; Vincieri, F. F.
Costantini, A. HPLC-DAD and TLC-densitometry for
Quantification of Hypericin in Hypericum Perforatum L. Extracts.
J. Chromatogr. 1999, 49(3–4), 197–201. DOI: 10.1007/bf02575285.
[61] Wang, J. F.; Cui, D. F.; Mu, X. Determination of Hypericin in the
Extract of St. John’s Wort. HB. J. Anim. Vet. Sci. 2014, 35(5), 13–14.
DOI: 10.16733/j.cnki.1007-273x.2014.05.053.
[62] Ruckert, U.; Eggenreich, K.; Likussar, W.; Wintersteiger, R.;
Michelitsch, A. A High-performance Liquid Chromatography
with Electrochemical Detection for the Determination of Total
Hypericin in Extracts of St. John’s Wort. Phytochem. Anal. 2006,
17(3), 162–167. DOI: 10.1002/pca.908.
[63] Zhang, J.; Feng, C. R.; Xu, H.; Tan, X. M.; Hagedoorn, P. L.;
Ding, S. G. Enhanced Hypericin Extraction from Hypericum
Perforatum L. By Coupling Microwave with Enzyme-assisted
Strategy. Ind. Crop. Prod. 2019, 137, 231–238. DOI: 10.1016/j.
indcrop.2019.05.036.
[64] Dolezal, R.; Houdkova, I.; Kalasz, H.; Andrys, R.; Novak, M.;
Maltsevskaya, N. V.; Karaskova, N.; Kolar, K.; Novotna, E.;
Kuca, K.; et al. Determination of Hypericin in Hypericum
Perforatum (St. John’s Wort) Using Classical C18 and
Pentafluorophenyl Stationary Phases: Contribution of Pi-Pi
Interactions to High-performance Liquid Chromatography
(HPLC). Anal. Lett. 2019, 52(11), 1788–1812. DOI: 10.1080/
00032719.2019.1571076.
[65] Jensen, K. I. N.; Gaul, S. O.; Specht, E. G.; Doohan, D. J. Hypericin
Content of Nova Scotia Biotypes of Hypericum Perforatum L. Can.
J. Plant. Sci. 1995, 75(4), 923–926. DOI: 10.4141/cjps95-155.
[66] Ozkan, E. E.; Ozden, T. Y.; Ozsoy, N.; Mat, A. Evaluation of
Chemical Composition, Antioxidant and
Anti-acetylcholinesterase Activities of Hypericum
Neurocalycinum and Hypericum Malatyanum. S. Afr. J. Bot. 2018,
114, 104–110. DOI: 10.1016/j.sajb.2017.10.022.
[67] Li, Y. H.; Wang, F. S. Comparison of Hypericin Contents in
Hypericum Hubeiense and Hypericum Perforatum. Med. Plant.
2010, 1(8), 29–30. DOI: 10.13989/j.cnki.0517-6611.2010.28.165.
[68] Wang, X. Determination of Hypericin Content in Hypericum
Perforatum L. By HPLC. Med. Plant. 2010, 1(9), 79–80.
[69] Liu, F. F.; Ang, C. Y. W.; Heinze, T. M.; Rankin, J. D.; Beger, R. D.;
Freeman, J. P.; Lay, J. O. Evaluation of Major Active Components
in St. John’s Wort Dietary Supplements by High-performance
Liquid Chromatography with Photodiode Array Detection and
SEPARATION & PURIFICATION REVIEWS 9
Electrospray Mass Spectrometric Confirmation. J. Chromatogr. A.
2000, 888, 85–92. DOI: 10.1016/s0021-9673(00)00555-0.
[70] Cossuta, D.; Vatai, T.; Bathori, M.; Hohmann, J.; Keve, T.;
Simandi, B. Extraction of Hyperforin and Hypericin from
St. John’s Wort (Hypericum Perforatum L.) With Different
Solvents. J. Food. Process. Eng. 2012, 35(2), 222–235. DOI:
10.1111/j.1745-4530.2010.00583.x.
[71] Xing, G. Z. Study on Extraction and Chemical Synthesis of
Hypericin. M. D. Dissertation, Chinese Academy of Agricultural
Sciences, Beijing, China, 2007.
[72] Punegov, V. V.; Kostromin, V. I.; Fomina, M. G.; Zaynullin, V. G.;
Yushkova, E. A.; Belyh, D. V.; Chukicheva, I. U.; Zaynullin, G. G.
Microwave-assisted Extraction of Hypericin and Pseudohypericin
from Hypericum Perforatum. Russ. J. Bioorg. Chem. 2015, 41(7),
757–761. DOI: 10.1134/s1068162015070122.
[73] Xue, X. L.; Zhang, X. H.; Song, H. Y.; Zhang, K. Q. Separation and
Purification of Hypericin from Hypericum Perforatum by
Macroporous Resin. Chin. J. Vet. Drug. 2016, 50(7), 20–25.
[74] Wang, X. X.; Pei, Y. X.; Hou, Y.; Pei, Z. C. Fabrication of Core-shell
Magnetic Molecularly Imprinted Nanospheres Towards Hypericin
via Click Polymerization. Polymers. 2019, 11(2), 313–326. DOI:
10.3390/polym11020313.
[75] Li, Z. Z.; Qin, C. L.; Li, D. M.; Hou, Y. Z.; Li, S. B.; Sun, J. J.
Molecularly Imprinted Polymer for Specific Extraction of
Hypericin from Hypericum Perforatum L. Herbal Extract.
J. Pharm. Biomed. Anal. 2014, 98, 210–220. DOI: 10.1016/j.
jpba.2014.05.031.
[76] Cao, X. L.; Wang, Q.; Li, Y.; Bai, G.; Ren, H.; Yiochiro, I. Isolation
and Purification of Series Bioactive Components from Hypericum
Perforatum L. By Counter-current Chromatography.
J. Chromatogr. B. Analyt. Technol. Biomed. Life. Sci. 2011, 879
(7–8), 480–488. DOI: 10.1016/j.jchromb.2011.01.007.
[77] Wei, Y.; Xie, Q. Q.; Dong, W. T.; Ito, Y. Separation of
Epigallocatechin and Flavonoids from Hypericum Perforatum
L. By High-speed Counter-current Chromatography and
Preparative High-performance Liquid Chromatography.
J. Chromatogr. A. 2008, 1216(19), 4313–4318. DOI: 10.1016/j.
chroma.2008.12.056.
[78] Kuang, P.; Song, D.; Yuan, Q.; Yi, R.; Lv, X.; Liang, H. Separation
and Purification of Sulforaphene from Radish Seeds Using
Macroporous Resin and Preparative High-performance Liquid
Chromatography. Food. Chem. 2013, 136(2), 342–347. DOI:
10.1016/j.foodchem.2012.08.082.
[79] Zhang, J.; Zhou, X.; Fu, M. Integrated Utilization of Red Radish
Seeds for the Efficient Production of Seed Oil and Sulforaphene.
Food. Chem. 2016, 192, 541–547. DOI: 10.1016/j.
foodchem.2015.07.051.
[80] Duran, N.; Song, P. S. Hypericin and Its Photodynamic Action.
Photochem. Photobiol. 1986, 43(6), 677–680. DOI: 10.1111/j.1751-
1097.1986.tb05646.x.
[81] Zhang, C. L.; Fan, J. Application of Hypericin in Tumor Treatment
and Diagnosis. J. Int. Pharm. Res. Int. 2012, 39(5), 402–408. DOI:
0.13220/j.cnki.jipr.2012.05.008.
[82] Weiner, L.; Mazur, Y. EPR Study of Hypericin-photogeneration of
Free-radicals and Superoxide. J. Chem. Soc-Perkin. Trans 2. 1992,
100(9), 1439–1442. DOI: 10.1039/p29920001439.
[83] Chung, P. S.; Saxton, R. E.; Paiva, M. B.; Rhee, C. K.; Soudant, J.;
Mathey, A.; Foote, C.; Castro, D. J. Hypericin Uptake in Rabbits
and Nude-mice Transplanted with Human Squamous-cell
Carcinoma-study of a New Sensitizer for Laser Phototherapy.
J. Laryngosc. 1994, 104(12), 1471–1476. DOI: 10.1288/00005537-
199412000-00008.
[84] Hyejin, K.; Sung, W. K.; Wang, H. S.; Chi, W. H.; Ahn, J. C.;
Jin, J. O.; Kang, H. W. Hypericin-assisted Photodynamic Therapy
against Anaplastic Thyroid Cancer. Photodiagnosis. Photodyn.
Ther. 2018, 24, 15–21. DOI: 10.1016/j.pdpdt.2018.08.008.
[85] Wielgus, A. R.; Chignell, C. F.; Miller, D. S.; Houten, B. V.;
Meyer, J.; Hu, D. N.; Roberts, J. E. Phototoxicity in Human
Retinal Pigment Epithelial Cells Promoted by Hypericin,
a Component of St. John’s Wort. Photochem. Photobiol. 2007, 83
(3), 706–713. DOI: 10.1562/2006-08-09-ra-1001.
[86] Jendzelovska, Z.; Jendzelovsky, R.; Kucharova, B.; Fedorocko, P.
Hypericin in the Light and in the Dark: Two Sides of the Same
Coin. Front. Plant. Sci. 2016, 7, 560–580. DOI: 10.3389/
fpls.2016.00560.
[87] Alali, F.; Tawaha, K.; Al-Eleimat, T. Determination of Hypericin
Content in Hypericum Triquetrifolium Turra (Hypericaceae)
Growing Wild in Jordan. Nat. Prod. Res. 2004, 18(2), 147–151.
DOI: 10.1080/14786410310001608046.
[88] Uzdensky, A. B.; Bragin, D. E.; Kolosov, M. S.; Kubin, A.;
Loew, H. G.; Moan, J. Photodynamic Effect of Hypericin and a
Water-soluble Derivative on Isolated Crayfish Neuron and
Surrounding Glial Cells. J. Photochem. Photobiol. B-Biol. 2003, 72
(1–3), 27–33. DOI: 10.1016/j.jphotobiol.2003.08.008.
[89] Kesa, P.; Jancura, D.; Kudlacova, J.; Valusova, E.; Antalik, M.
Excitation of Triplet States of Hypericin in Water Mediated by
Hydrotropic Cromolyn Sodium Salt. Spectroc. Acta Pt. A-Molec.
Biomol. Spectr. 2018, 193, 185–191. DOI: 10.1016/j.saa.2017.12.004.
[90] Bano, G.; Stanicova, J.; Jancura, D.; Marek, J.; Bano, M.; Ulicny, J.;
Strejckova, A.; Miskovsky, P. On the Diffusion of Hypericin in
Dimethylsulfoxide/water Mixtures-the Effect of Aggregation. J. Phys.
Chem. B. 2011, 115(10), 2417–2423. DOI: 10.1021/jp109661c.
[91] Buzova, D.; Kasak, P.; Miskovsky, P.; Jancura, D. Solubilization of
Poorly Soluble Photosensitizer Hypericin by Polymeric Micelles
and Polyethylene Glycol. Gen. Physiol. Biophys. 2013, 32(2),
201–208. DOI: 10.4149/gpb_2013023.
[92] Kubin, A.; Loew, H. G.; Burner, U.; Jessner, G.; Kolbabek, H.;
Wierrani, F. How to Make Hypericin Water-soluble. Pharmazie.
2008, 63(4), 263–269. DOI: 10.1691/ph.2008.7292.
[93] Wirz, A.; Meier, B.; Sticher, O. Stability of Hypericin and
Pseudohypericin in Extract Solutions of Hypericum Perforatum
and in Standard Solutions. Pharm. Ind. 2001, 63(4), 410–415.
DOI: 10.3109/10837450.2010.529148.
[94] Wang, J. L.; Li, Q. Q.; Wang, X. L.; He, L. S.; Zhang, J. S. Stability of
Hypericin Extract. Chin. Pharm. Ind. 2014, 23(13), 9–11.
[95] Bai, W. B.; Liang, J. P.; Wei, Y. M.; Cui, Y.; Wang, X. H.; Hua, L. Y.;
Sang, R. F.; Niu, J. R.; Lv, J. W.; Guo, D. S. Study on the Extraction
and Purification of Hypericin from Hypericum Perforatum L. With
Ethanol. J. Tradit. Chin. Vet. Med. 2005, 3, 17–18. DOI: 10.13823/j.
cnki.jtcvm.2005.03.005.
[96] Wang, X. J.; Zhang, L. W. Stability of Hypericin. Chin. J. Spec. Lab.
2005, 22(4), 797–800. DOI: 10.3969/j..1004-8138.2005.04.039.
10 J. ZHANG ET AL.
... Hypericin can be extracted from flowers and fruits, while pseudohypericin is predominantly found in the aerial parts of H. sampsonii (Sun et al., 2023). The antiviral properties of both compounds have been established in numerous studies, encompassing herpes simplex types 1 and 2, and HIV-1 ( Barnes et al., 2001;Zhang et al., 2022). Hypericin is actively employed in the development of drugs or intermediate compounds for photodynamic therapy (PDT) and photodynamic diagnosis (PDD) (Jendželovská et al., 2016;Yuan et al., 2023). ...
Article
Monkeypox (Mpox) is a viral zoonotic and human-to-human disease with no specific drug or treatment protocol targeting the monkeypox virus (MPXV). In the MPXV life cycle, viral kinase phosphorylation plays a crucial role in early morphogenesis in the cytoplasm, making inhibition of MPXV kinase a potential therapeutic approach for controlling Mpox. Hypericum sampsonii contains several bioactive compounds, such as hypericin and pseudohypericin, which are known for their antiviral properties. In this study, a computational investigation of the physicochemical properties of hypericin and pseudohypericin revealed drug-like characteristics. Pharmacokinetic predictions indicated that hypericin and pseudohypericin are non-toxic to the central nervous system, hepatic system, and cardiac system. Molecular docking results indicated a strong binding affinity of hypericin/pseudohypericin with MPXV thymidylate kinase. As a result, these compounds are being considered as potential Mpox control candidates.
... John's wort), basidiomycetes (Dermocybe spp.) and endophytic fungi (Thielavia subthermophila). As a natural substance with unexpectedly diverse and beneficial medical effects, it has been explored in recent decades for its wide pharmacological spectrum [1][2][3][4][5][6]. HY has antibacterial and antiviral, anticancer, antioxidant and neuroprotective properties, and has therefore been used in the treatment of many diseases, either as an independent substance or, using its photosensitizing properties, in photodynamic therapy [7,8]. ...
Article
Full-text available
Determination of the hypericin–photodynamic (HY–PDT) effect on the secretion of cytokines secreted by the skin cells, may be the basis for using the immunomodulatory effect of photodynamic action in the treatment of inflammatory skin diseases. The study aimed to evaluate the cytotoxic and immunomodulatory effects of hypericin (HY) in photodynamic therapy (PDT) performed in vitro on cultures of selected skin cell lines. The study used two human cell lines, primary dermal fibroblast (HDFa) and primary epidermal keratinocytes (HEKa). The MTT test was used to define the metabolic activity of treated cells. Cell supernatants subjected to sublethal PDT were assessed to determine the interleukins: IL-2, IL-8, IL-10, IL-11, IL-19, IL-22, and metalloproteinase 1 (MMP-1). The results confirm the destructive effect of HY–PDT and the immunomodulatory effects of sublethal doses on the selected skin cells, depending on the concentration of HY and the light doses. No statistically significant differences were noted in IL-2 and IL-10 concentration after HY–PDT for HEKa and HDFa lines. After using HY–PDT, the concentration of IL-8, MMP-1, IL-22, and IL-11 significantly decreased in the HEKa line. Moreover, the concentration of IL-19 and MMP-1 significantly decreased in the HDFa line. The concentration of IL-11 in the HDFa line after using only the HY, without the light, increased but decreased after HY–PDT. Our experiment confirmed that HY–PDT has not only a cytotoxic effect but, used in sublethal doses, also presents immunomodulatory properties. These may be an advantage of HY–PDT when used in the treatment of persistent skin inflammation, connected with the release of pro-inflammatory cytokines resistant to conventional treatment methods.
... Natural anthraquinones which contains the anthraquinone rings and some groups such as hydroxyl and carboxyl, are mainly extract from rhubarb or Saint John's Wort, have attracted everincreasing attention for their multiple pharmacological activities. [49][50][51][52] Moreover, these molecules also possess good photosensitivity and can be considered as potential PSs (Figure 3d,e). [23,[53][54] For instance, emodin and aloe-emodin are two types of well-studied natural PSs and show outstanding PDT performance. ...
Article
Full-text available
Natural photosensitizers (PSs) and their derivatives have drawn ever‐increasing attention in photodynamic therapy (PDT) for their wild range of sources, desirable biocompatibility, and good photosensitivity. Nevertheless, many factors such as poor solubility, high body clearance rate, limited tumor targeting ability, and short excitation wavelengths severely hinder their applications in efficient PDT. In recent years, fabricating nanostructures by utilizing molecular assembly technique is proposed to solve these problems. This technique is easy to put into effect, and the assembled nanostructures could improve the physical properties of the PSs so as to meet the requirement of PDT. In this concept, we focus on the construction of natural PSs and their derivatives nanostructures through molecular assembly technique to enhance PDT efficacy (Figure 1). Furthermore, current challenges and future perspectives of natural PSs and their derivatives for efficient PDT are discussed.
... Hypericin is a natural compound extracted from Hypericum perforatum [1,2]. It exhibits various pharmacological effects, such as anti-depressive, anti-tumor, anti-inflammatory, and antiviral activity [3,4]. ...
Article
Full-text available
The aim of this study was to explore the potential of hypericin, a naturally occurring photosensi-tizer, for photodynamic therapy (PDT) in skin cancer, investigating its phototoxic effects and mechanisms of action in cancer cells compared to normal skin keratinocytes, squamous cell cancer (SCC-25) cells and melanoma (MUG-Mel2) cells. Hypericin was applied at concentrations ranging from 0.1–40 μM to HaCaT, SCC-25, and MUG-Mel2 cells. After 24 h of incubation, the cells were exposed to orange light at 3.6 J/cm2 or 7.2 J/cm2. Phototoxicity was assessed using MTT and SRB tests. Cellular uptake was measured by flow cytometry. Apoptosis-positive cells were estimated through TUNEL for apoptotic bodies’ visualization. Hypericin exhibited a higher phototoxic reaction in cancer cells compared to normal keratinocytes after irradiation. Cancer cells demonstrated increased and selective uptake of hypericin. Apoptosis was observed in SCC-25 and MUG-Mel2 cells following PDT. Our findings suggest that hypericin-based PDT is a promising and less invasive approach for treating skin cancer. The higher phototoxic reaction, selective uptake by cancer cells, and observed proapoptotic properties support the promising role of hypericin-based PDT in skin cancer treatment.
... In addition, the antiviral activity has been reported for hypericin (199) and pseudohypericin (200) against herpes simplex virus types 1 and 2 and HIV-1 in vitro. Hypericin (199) has also exhibited activity against HCV, murine cytomegalovirus (MCMV), Sindbis virus, infectious bronchitis virus, and novel duck reovirus (Barnes et al., 2001;Zhang et al., 2022). ...
Article
Full-text available
Ethnopharmacological relevance: Hypericum sampsonii Hance, also known as Yuanbao Cao in Chinese, is a traditional medicinal herb from the Guttiferae family and has been widely used in China to treat various conditions, including dysentery, enteritis, mastitis, scrofula, and contusion. Aim of the review: This review aims to provide a comprehensive overview of the botany, traditional uses, phytochemistry, biological activity and safety of H. sampsonii and to highlight its potential for medical application and drug development. Materials and methods: We searched several databases, i.e., Web of Science, SciFinder, PubMed, CBM, CNKI, Google Scholar, etc., for relevant information on H. sampsonii. Additionally, we also consulted some books on Chinese medicine. Results: To date, 227 secondary metabolites have been isolated from H. sampsonii, including polycyclic polyprenylated acylphloroglucinols (PPAPs), benzophenones, xanthones, flavonoids, naphthodianthrones, anthraquinones and aromatic compounds. These metabolites exhibit various biological activities such as anti-inflammatory, anti-tumor, anti-depressant, anti-oxidant, anti-viral and anti-bacterial effects. PPAPs are considered the main active metabolites with rich biological activities. Despite being known as rich source of PPAPs, the full extent of H. sampsonii biological activities, including their potential as PDE4 inhibitors, remained unclear. Since, previous studies have mainly been based on structural identification of metabolites in H. sampsonii, and efficacy evaluations of these metabolites based on clinical applications of H. sampsonii lack sufficient data. However, current evidence suggest that PPAPs are the most likely material basis for efficacy. From the limited information available so far, there is no evidence of potential safety issues and the safety data are limited. Conclusion: Collectively, this review provides a comprehensive overview of the botany, traditional uses, phytochemistry, pharmacology, and safety of H. sampsonii, a valuable medicinal plant in China with various pharmacological activities. Based on pharmacological studies, H. sampsonii shows potential for treating gastrointestinal and gynecological disorders as well as traumatic injuries, which aligns with traditional medicinal use due to the presence of PPAPs, benzophenones, xanthones, and flavonoids. Therefore, further studies are needed to evaluate the pharmacological effects and elucidate the pharmacological mechanisms. In addition, pharmacological mechanisms and safety evaluation of PPAPs on animal models need to be clarified. Yet, further comprehensive studies are required to elucidate the phytochemical constituents, pharmacological mechanisms, structure-activity relationships, safety evaluation, and quality standards of this plant. Takentogether, this review highlights the potential of H. sampsonii for medical application and drug development.
Article
Full-text available
Hypericin, a polycyclic naphthodianthrone and active plant pigment with the molecular formula C30H16O8, is a crucial phytochemical extracted from the dark-colored glands present on the aerial parts of the genus Hypericum. It is biosynthesized through the polyketide pathway by plant-specific type III polyketide synthases (PKSs). In addition to hypericin, the genus Hypericum is rich in various classes of phytochemicals. Alongside other bioactive compounds like hyperforin and flavonoids, hypericin exhibits antidepressant activity. Recently, hypericin has gained increased importance in the research due to its unique properties. Its photodynamic nature makes it an effective natural photosensitizer, extending its use in investigating skin disorders. Moreover, hypericin demonstrates antiviral and antitumoral properties. Despite its effectiveness in treating cancers and neurological disorders, hypericin production faces challenges due to its site-specific nature. Conventional methods struggle to meet the growing demand for hypericin. Biotechnological approaches, including plant tissue culture and bioreactor-based large-scale production, offer promising solutions to address this demand. This review focuses on various plant tissue culture techniques, such as cell and organ culture, and elucidates their biosynthetic pathways. It also discusses hypericin production using elicitation strategies involving biotic and abiotic components, as well as genetic engineering approaches to enhance hypericin yields. Bioreactor-scale production presents significant potential for sustainable hypericin production. Further advancements in understanding and engineering biosynthetic pathways hold promise for unlocking new avenues in hypericin production.
Article
Full-text available
Traditional Chinese Medicines (TCMs) have been used for centuries for the treatment and management of various diseases. However, their effective delivery to targeted sites may be a major challenge due to their poor water solubility, low bioavailability, and potential toxicity. Nanocarriers, such as liposomes, polymeric nanoparticles, inorganic nanoparticles and organic/inorganic nanohybrids based on active constituents from TCMs have been extensively studied as a promising strategy to improve the delivery of active constituents from TCMs to achieve a higher therapeutic effect with fewer side effects compared to conventional formulations. This review summarizes the recent advances in nanocarrier-based delivery systems for various types of active constituents of TCMs, including terpenoids, polyphenols, alkaloids, flavonoids, and quinones, from different natural sources. This review covers the design and preparation of nanocarriers, their characterization, and in vitro/vivo evaluations. Additionally, this review highlights the challenges and opportunities in the field and suggests future directions for research. Nanocarrier-based delivery systems have shown great potential in improving the therapeutic efficacy of TCMs, and this review may serve as a comprehensive resource to researchers in this field. Graphical abstract
Article
Hypericin can be derived from St. John's wort, which is widely spread around the world. As a natural product, it has been put into clinical practice such as wound healing and depression for a long time. In this article, we review the pharmacology, pharmacokinetics, and safety of hypericin, aiming to introduce the research advances and provide a full evaluation of it. Turns out hypericin, as a natural photosensitizer, exhibits an excellent capacity for anticancer, neuroprotection, and elimination of microorganisms, especially when activated by light, potent anticancer and antimicrobial effects are obtained after photodynamic therapy. The mechanisms of its therapeutic effects involve the induction of cell death, inhibition of cell cycle progression, inhibition of the reuptake of amines, and inhibition of virus replication. The pharmacokinetics properties indicate that hypericin has poor water solubility and bioavailability. The distribution and excretion are fast, and it is metabolized in bile. The toxicity of hypericin is rarely reported and the conventional use of it rarely causes adverse effects except for photosensitization. Therefore, we may conclude that hypericin can be used safely and effectively against a variety of diseases. We hope to provide researchers with detailed guidance and enlighten the development of it.
Article
Background: Hypericin (HYP) is a natural compound widely used as a food supplement. The encapsulation of HYP into nanosystems, such as nanostructured lipid carriers (NLC), is a promising strategy for delivering this lipophilic molecule and protecting it from degradation. Objective: The present study aims to develop and validate an analytical method to quantify the encapsulation efficiency of HYP in NLC. Methods: A reverse-phase high-performance liquid chromatography (HPLC) method was developed and validated according to ICH guide Q2(R1). NLC was prepared through the ultrasonication method and HYP encapsulation efficiency was evaluated using the validated method. Results: Separation was achieved using an isocratic mobile phase composed of acetonitrile, methanol, and ammonium acetate buffer (10 mM, pH 5.0) (54:36:10 v/v/v) and a reverse stationary phase. The specificity, linearity, precision, accuracy, and robustness of the method were assessed and confirmed during the validation. Furthermore, the validated method was able to determine the encapsulation efficiency of HYP in NLC. Conclusion: the HPLC method was validated, and the results indicated the ability of NLC to deliver HYP compounds for further application as a food supplement. Highlights: Hypericin is used as a food supplement and for photodynamic therapy.The developed method was specific, linear, precise, accurate, and robust Nanostructured lipid carriers showed a high ability to encapsulate hypericin.
Article
Full-text available
Avian infectious bronchitis virus (IBV), a coronavirus, causes infectious bronchitis leading to enormous economic loss in the poultry industry worldwide. Hypericin (HY) is an excellent compound that has been investigated in antiviral, antineoplastic, and antidepressant. To investigate the inhibition effect of HY on IBV infection in chicken embryo kidney (CEK) cells, 3 different experimental designs: pre-treatment of cells prior to IBV infection, direct treatment of IBV-infected cells, and pre-treatment of IBV prior to cell infection were used. Quantitative real-time PCR (qRT-PCR), immunofluorescence assay (IFA), flow cytometry, and fluorescence microscopy were performed and virus titer was determined by TCID50. The results revealed that HY had a good anti-IBV effect when HY directly treated the IBV-infected cells, and virus infectivity decreased in a dose-dependent manner. Furthermore, HY inhibited IBV-induced apoptosis in CEK cells, and significantly reduced the mRNA expression levels of Fas, FasL, JNK, Bax, Caspase 3, and Caspase 8, and significantly increased Bcl-2 mRNA expression level in CEK cells. In addition, HY treatment could decrease IBV-induced reactive oxygen species (ROS) generation in CEK cells. These results suggested that HY showed potential antiviral activities against IBV infection involving the inhibition of apoptosis and ROS generation in CEK cells.
Article
Full-text available
The core-shell structure molecularly imprinted magnetic nanospheres towards hypericin (Fe3O4@MIPs) were prepared by mercapto-alkyne click polymerization. The shape and size of nanospheres were characterized by dynamic light scattering (DLS) and transmission electron microscope (TEM). The nanospheres were analyzed by FTIR spectroscopy to verify the thiol-yne click reaction in the presence or absence of hypericin. The Brunauer–Emmet–Teller (BET) method was used for measuring the average pore size, pore volume and surface area. The Fe3O4@MIPs synthesized displayed a good adsorption capacity (Q = 6.80 µmol·g−1). In addition, so-prepared Fe3O4@MIPs showed fast mass transfer rates and good reusability. The method established for fabrication of Fe3O4@MIPs showed excellent reproducibility and has broad potential for the fabrication of other core-shell molecularly imprinted polymers (MIPs).
Article
Full-text available
A wide range of compounds that occur in the genus Hypericum are listed as effective drugs of natural origin. The main biological activities of several Hypericum representatives are due to the presence of naphthodianthrones, phloroglucinols, and other diverse groups of secondary metabolites that synergistically contribute to their therapeutic effects. The regulation of biosynthesis of hypericin as the key bioactive naphthodianthrone remains uncertain. Here, we present liquid chromatography mass spectrometry-based phenotyping of 17 Hypericum species, the results of which suggest an important role for skyrin and its derivatives in the polyketide pathway that leads to hypericin formation. Moreover, we report for the first time the presence of new metabolites in the genus Hypericum that are related to classes of anthraquinones, their derivatives, and phloroglucinols. As skyrin and other species of anthraquinones are rarely found in higher plants but frequently occur in fungal microorganisms, the obtained results suggest that further research on the synthesis pathways of hypericin and the role of anthraquinone derivatives in plant metabolism should be carried out. The fact that these compounds are commonly synthesized in endophytic fungi and perhaps there is some similarity in the metabolic pathways between these organisms should also be investigated. Electronic supplementary material The online version of this article (10.1007/s00216-018-1384-0) contains supplementary material, which is available to authorized users.
Article
Within Sicilian flora, the genus Hypericum (Guttiferae) includes 10 native species, the most popular of which is H. perforatum. Hypericum’s most investigated active compounds belong to naphtodianthrones (hypericin, pseudohypericin) and phloroglucynols (hyperforin, adhyperforin), and the commercial value of the drug is graded according to its total hypericins content.Ethnobotanical sources attribute the therapeutic properties recognized for H. perforatum, also to other Hypericum species. However, their smaller distribution inside the territory suggests that an industrial use of such species, when collected from the wild, would result in an unacceptable depletion of their natural stands. This study investigated about the potential pharmacological properties of 48 accessions from six native species of Hypericum, including H. perforatum and five "minor" species, also comparing, when possible, wild and cultivated sources.The variability in the content of active metabolites was very high, and the differences within the species were often comparable to the differences among species. No difference was enlightened between wild and cultivated plants. A properly planned cultivation of Hypericum seems the best option to achieve high and steady biomass yields, but there is a need for phytochemical studies, aimed to identify for multiplication the genotypes with the highest content of the active metabolites.
Article
Novel duck reovirus (NDRV) disease is a serious infectious disease for poultry, for which no effective therapy has been established. Therefore, development of novel antivirals against NDRV is urgently needed. In present study, we developed a complex wherein hypericin (HY), which shows broad-spectrum antiviral activity, was loaded onto graphene oxide (GO), which has a high drug-loading capacity and low cytotoxicity. The antiviral activity of the complex (GO/HY) was studied in DF-1 cells and in ducklings infected with the NDRV TH11 strain. GO/HY showed a dose-dependent inhibition of NDRV replication, which may be attributed to direct virus inactivation or inhibition of virus attachment. Western blotting and indirect immunofluorescence assay (IFA) showed markedly suppressed protein expression in GO/HY-treated NDRV-infected DF-1 cells. Moreover, GO/HY prolonged the survival time of the ducklings by reducing pathological lesions caused by the infection and inhibiting viral replication in the liver and lungs. These results suggest that GO/HY has antiviral activity against NDRV both in vitro and in vivo.
Article
Hypericin is considered to be the most biologically active substance in the crude extract of Hypericum perforatum L. (also known as St. John's wort) and has a wide range of pharmacological effects. In this study, a high resolution high performance liquid chromatography method for determining hypericin was established by comparing different chromatographic conditions. Xylanase-assisted extraction and microwave-assisted extraction can improve the extraction yield of hypericin significantly. And the coupling strategy between two methods resulted in a significant difference on the extraction yield. Microwave-assisted extraction after xylanase-assisted extraction was found to be the most efficient strategy for extracting hypericin. The yield was 0.319 ± 0.006 mg g −1 , which was a 209.7% increase over unassisted extraction. The combination of enzyme assisted extraction followed by microwave assisted extraction can be more commonly applied to improve extraction efficiency of bioactive compounds from plants.
Article
Hypericin is one of the most important bioactive substances present in Hypericum perforatum (Saint John’s Wort), which exhibits various biological activities such as inhibition of neuronal processes coupled with transmission of serotonin, dopamine, gamma-amino butyric acid and L-glutamine. This red-colored derivative of anthraquinone contains six benzene cycles and four hydroxy groups that polarize electron density in the conjugated double bond system and enable the molecule to act as a Lewis base in charge-transfer complexes. In this study, four stationary phases modified by the pentafluorophenyl group were selected to investigate the contribution of π-π interactions to the improvement of hypericin separation in comparison to the separation provided with a classical C18 based chromatographic column. In order to develop a suitable gradient chromatographic method for the quantification of hypericin, normal, polar organic and reverse elution modes were evaluated using high-performance liquid chromatography with ultraviolet-visible and mass spectrometry detection. It was disclosed in the analyses that the pentafluorophenyl stationary phase can separate hypericin from hyperforin, which was not possible to achieve with the C18 based stationary phase. The best analytical method found, employing the pentafluorophenyl stationary phase, showed sufficient linearity, accuracy and precision and was used for the determination of hypericin in Saint John’s Wort.
Article
Background/aim: We report the incorporation of prospective anticancer agent hypericin into DNA and bovine serum albumin (BSA), respectively, with emphasis on comparison of the differences in interaction mode between hypericin and its model compound emodin. Materials and methods: Spectrophotometric methods were used for determination of the binding constants of the drug complex with biomacromolecules. Differential scanning calorimetry was applied for evaluation of drug-macromolecule complex thermal stability. Results: The strength of interaction expressed by binding constants was found to be 4.0×104 l/mol for hypericin-DNA and 8.1×104 l/mol for emodin-DNA complex. Both molecules stabilize bovine serum albumin macromolecule and bind into the hydrophobic cavity in IIA subunit but their localization within the molecule is different. Conclusion: Anticancer agent hypericin and its derivative emodin interact with DNA with medium strength and are probably incorporated into the groove of DNA by hydrogen bonds. Bovine serum albumin can serve as a transport protein for hypericin since the binding force between both molecules is adequate.
Article
Background: Hypericin (HYP) extracted from St. John's wort ( Hypericum perforatum L.) is a natural photosensitizer in clinical photodynamic therapy (PDT). The PDT is one of the powerful methods for cancer treatments because of its excellent tumoritropic characteristics and photosensitizing properties. However, limited reports on the efficacy of PDT on the anaplastic thyroid cancer (ATC) have been published. Especially HYP-associated PDT has not been investigated in vitro and in vivo. In this study, we evaluated the effect of HYP for PDT against FRO ATC cells. Methods: The activities of HYP-assisted PDT were investigated in ATC cells. The ATC FRO cells were treated with a combination of HYP dose and laser power. The viability of FRO cells was measured by MTT assay, and Trypan blue staining was performed to monitor cell death. Detection reactive oxygen species (ROS) and mitochondrial membrane potential after HYP-assisted PDT were analyzed by confocal microscopy. For in vivo study, FRO cells were injected into nude mice. After intravenous injection of HYP, Laser was irradiated and nude mice were monitored in Day 4, 7, 14. Results and conclusions: The rate of FRO cell death was increased by applying HYP dose and laser power dependent. Moreover, HYP and laser irradiation induced FRO cell death was mediated by the intracellular ROS generation and mitochondrial damage. Finally, the HYP-assisted PDT eliminated FRO cell tumor from the mouse in vivo. These data demonstrate that HYP could be an effective photosensitizer for human ATC therapy.
Article
Effective targeting of mitochondria has emerged as a beneficial strategy in cancer therapy. However, the development of mitochondria-targeting ligands is difficult because of the low permeability of the mitochondrial double membrane. We found that hypericin (HY), a natural product isolated from Hypericum perforatum L., is an effective mitochondria-targeting ligand. HY-functionalized graphene oxide (GO) loaded with doxorubicin (GO-PEG-SS-HY/DOX) increased the synergistic anticancer efficacy of phototherapy and chemotherapy in the absence of apparent adverse side effects. In vitro and in vivo assays suggested GO-PEG-SS-HY/DOX induced the expression of the key proteins of the mitochondria-mediated apoptosis pathway and caused apoptosis of breast carcinoma cells. In addition, GO vehicle exhibited low toxicity toward normal cells, indicating high safety of functionalized GO preparations in antitumor therapy. Therefore, HY-functionalized GO can be successfully used as a platform technology to target mitochondria in cancer cells and improve the therapeutic efficacy of chemotherapeutic drugs. Statement of significance: Induction of mitochondria-mediated apoptosis is a promising approach in cancer therapy. However, mitochondria are difficult to access and permeate because of their negative membrane potential and highly dense double membrane. Mitochondria-targeting ligands can be conjugated to nanoparticles or small-molecule drugs to enhance their antitumor effect. Here, we showed that the natural photosensitizer hypericin is a novel mitochondria-targeting ligand and that graphene oxide particles co-loaded with hypericin and the chemotherapeutic agent doxorubicin exhibited a synergistic antitumor effect mediated by the mitochondrial-mediated apoptosis. Treatment with such particles in combination with laser irradiation led to apoptosis of the tumor MDA-MB-231 and MCF-7 cells in vitro and in vivo. Furthermore, treatment with hypericin/doxorubicin-functionalized graphene oxide had low cellular toxicity.