You are on page 1of 17

Anthraquinone

Anthraquinones (AQs) are the group of secondary metabolites produced by plants,


which are structurally related to 9,10-dioxoanthracene (also known as anthracene
9,10-diones) [1].

From: Advances in Molecular Toxicology, 2017

Related terms:

Pigment, Emodin, Dye, Solvent, Molecule, Ring Formation, Blue, Red, Antimicrobial
Agent

View all Topics

Learn more about Anthraquinone

A Primer on Colorful Additives


Ronald M. Harris, in Coloring Technology for Plastics, 1999

Anthraquinone Dyes
The anthraquinone dyes are based on the structure shown in Figure 4. Key properties
are summarized in Table 3. Most commercial anthraquinone dyes have sufficient
heat stability to be used in polycarbonate and thermoplastic polyesters. However, as
indicated in Table 3, only a handful of these dyes are suitable for polyamide resins
and their blends and alloys. Even in these cases caution must be applied. Polyamide
materials colored with red anthraquinone dyes have been observed to slowly shift
bluer over time. One hypothesis states the color shift is due to the slow absorption
of moisture.

Figure 4. The anthraquinone ring system.


Table 3. Dyes

Dye Strength Heat stability Weather re- FDA status Cost $/lbs
sistance
Antraquinone dyes
Solvent Red G E E in mass F in no 25.0
111 tint
Solvent Red 52 G E G no 100.0
Solvent Blue G F-G F no 35.0
59
Solvent Blue G E F no 55.0
97
Solvent Yellow G E F no 30.0
163
Solvent Green G E F no 65.0
28
Solvent Green G G F no 40.0
3
Solvent Violet G F F no 45.0
13
Perinone dyes
Solvent Red G G E in mass F in yes, PET 35.0
135 tint
Solvent Red G-E E F no 43.0
179
Solvent Or- G G E in mass F in no 40.0
ange 60 tint
Fluorescent
dyes
Solvent Yellow E G-E F no 100.0
160
Vat Red 41 E G F no 85.0
Solvent Or- E G F no 115.0
ange 63

> Read full chapter

Barrier textiles for protection against


microbes
Y. Zhao, ... T. Lin, in Antimicrobial Textiles, 2016

12.2.3 Photoactive dyes


Commercial anthraquinone vat dyes (eg, C.I. 67300, C.I. 59100, and C.I. 60515) were
found to have photoreactivity and can generate reactive oxygen species under UV or
visible light irradiation (Gee et al., 1973; Diaz, 1990). Because of their photoreactivity,
they have been employed to develop colored textiles with light-induced antimicro-
bial property (Zhuo and Sun, 2013; Liu et al., 2011; Zhu and Sun, 2014). These an-
thraquinone dyes can be incorporated into fabrics by using a vat dyeing process (Fig.
12.1) (Zhuo and Sun, 2013). Anthraquinone dyes are water insoluble. For vat dyeing,
they are converted to water-soluble sodium phenolate of anthrahydroquinone (also
called leuco salt) by reaction with sodium dithionite. The leuco salts are easy to
be exhausted into cotton, and they will then be oxidized back to water insoluble
anthraquinone to finish the vat dyeing process. Other anthraquinone derivatives that
bear reactive groups can be directly linked onto cotton by forming covalent bonds.
For instance, anthraquinone carboxylic acid has been bonded onto cotton surfaces
by using N,N -carbonyldiimidazole through a mild esterification reaction (Liu et al.,
2011).

Figure 12.1. Fabric treatment with anthraquinone dyes through a vat dyeing process.

Reprinted with permission from Zhuo, J., Sun, G., 2013. Antimicrobial functions on
cellulose materials introduced by anthraquinone vat dyes. ACS Applied Materials
& Interfaces 5, 10830. Copyright (2013) American Chemical Society.

> Read full chapter

Acid dyes
N. Sekar, in Handbook of Textile and Industrial Dyeing, 2011

15.4 Anthraquinone dyes


Anthraquinone acid dyes are rich in violet through blue to green complementing
the azo dyes. They have very good light fastness. They may be subdivided into the
following types.

1. 1-Amino-4-(substituted)amino anthraquinone-2-sulphonic acids


2. Diamino dihydroxy anthraquinone sulphonic acids

3. 1,4-Diaminoanthraquinones with external sulphonic acids

4. 1-Amino-4-hydroxy anthraquinones with external sulphonic acid groups

5. Other acid anthraquinone dyes.

Anthraquinone acid dyes offer bright blue shades not obtainable in azo dyes. The
red and yellow anthraquinone dyes are of little importance. Green dyes obtained
by the combination of yellow and blue dyes have mostly inferior wash fastness.
The uniformly dyeing green dyes of the anthraquinone series have proved their
special value in this context. When synthetic polyamides were known some selected
dyes from this series were used. Special acid dyes of this class have been recently
developed for this purpose.

15.4.1 1-Amino-4-(substituted)amino anthraquinone-2-sulphonic acids


These are obtained by the condensation of bromamine acid with aromatic or cy-
cloaliphatic amines, by an Ullmann reaction catalysed by copper or copper salts. This
group of blue acid dyes has important usage value and there are several reported
commercial structures (Chengqin, 1979; Chunlong, 1989; p. 189; Foris, 1977; p.
229; p. 57; Mengzheng, 1984; Venkataraman, 1952; Zollinger, 1987; p. 169). The
shade, levelling characteristics and light fastness can be varied over a wide range
by selecting particular amines. Cycloaliphatic amines give the same brightness as
aliphatic amines but give greater light fastness. Arylamines substituted with alkyl,
halogen, aryl, aryloxy or sulphonic ester group provide dyes with better wash fastness
and affinity in neutral media but levelling is improved. The brilliance is increased by
the substituents at the o-position accompanied by hypsochromic shift. The influence
of methyl groups on the colour and dyeing characteristics of acid dyes derived from
1-amino-4-bromoanthraquinone-2-sulphonic acid and arylamines has been studied
(Zhang and Hou, 1996). A new method for the preparation of bromamine acid
directly from bromamine acid has been reported (Ghaieni et al., 2007). A convenient
and efficient process for the manufacture of C.I. Acid Blue 78 from anthraquinone
has been reported (Ghaieni et al., 2008).

Some examples are C.I. Acid Blue 25, C.I. 62055 [6408-78-2] (52), C.I. Acid Blue 62,
C.I. 62045 [4368-56-3] (53), C.I. Acid Blue 129, C.I. 62058 [6397-02-0] (54), C.I. Acid
Blue 40, C.I. 62124 [6424-85-7] (55), and 56, and 57.
15.7. Anthraquinone acid dyes.

15.4.2 1-Amino-2-aryloxy-4-(substituted)amino anthraquinones


These are the violet acid dyes with a brilliant hue and excellent fastness introduced
between 1960 and 1970 (Peter and Gunthart 1951; Swiss Patent, 1966). The violet
acid dyes of the copper complex azo type gives dull shades while those of the
triphenylmethane type have very poor light fastness. The anthraquinone violet dyes
of this type can be used for dyeing natural and synthetic polyamide fibres in bright
shades of high light fastness.

They are obtained by condensation of bromamine acid witha di- or tri-alkylaniline


and subsequent condensation with an alkylphenol followed by sulphonation. While
the general formula of such dyes is known, the exact number of sulphonic acid
groups and their orientation are not clearly elucidated (Ciba-Geigy, 1980; Sandoz,
1979; Schwander et al., 1970). The structure and dyeing properties of such an-
thraquinone violet acid have been reported (Zhang et al., 1997).

15.4.3 Diamino dihydroxy anthraquinone sulphonic acids


These are the oldest synthetic dyes for wool, but their importance has decreased
considerably. An example is C.I. Acid Blue 43, C.I. 63000 [2150-60-9] (58).

15.4.4 1,4-Diaminoanthraquinones with external sulphonic acid groups


These are obtained by sulphonation of the corresponding dye base derived from
quinizarin or halo amino anthraquinones. The reaction products of quinizarin with
aromatic or araliphatic amines predominate in number and importance. The intro-
duction of hydroxy function into the 5- or the 5,8 positions causes the expected
red shift. Substituents influence the wash fastness and levelling. The shade can
be varied from brilliant blue to green by appropriate amines. Sterically hindered
amines, araliphatic and cycloaliphatic amines lead to brilliant blue shades. Dyes
in this series exhibit more light fastness than the derivatives of bromamine acid.
Among the unsymmetrically substituted dyes derived from 1-amino or 1-alky-
lamino-4-haloaminoanthraquinones, 1-alkylamino anthraquinones have inferior
light fastness. Exceptions are the 1-sec-alkylamino-4-halo anthraquinones. Intro-
duction of alkoxy or aryloxy groups into the 2 position introduces a blue shift giving
bright violet dyes.

Examples are C.I. Acid Green 25, C.I. 61570 [4403-90-1] (59), C.I. Acid Green 41, C.I.
62560 [4430-16-4] (60), greenish blue dyes 61, 62 and C.I. Acid Violet 42, C.I. 62026
[64026 [6408-73-7] (63).

15.4.5 1-Amino-4-hydroxy anthraquinones with external sulphonic acid


groups
These are obtained by the partial reaction of quinizarin with arylamines, followed
by sulphonation. They are the violet level dyeing ones. Derivatives of 1-amino-4-hy-
droxy-2-phenoxy anthraquinones were developed for polyamides of synthetic origin.
15.8. Anthraquinone acid dyes.

Examples are C.I. Acid Violet 43, C.I. 60730 [4430-18-6] (64) and the bluish brilliant
red dye 65.

15.4.6 Other acid anthraquinone dyes


In addition to the classes mentioned above a whole series of specially developed dyes
are available. For example, derivatives of the anthrimide and carbazole series are
known to be very light-fast grey and brown wool dyes. The sulphonation products
of 1,5- and 1,8-diarylamino anthraquinones are violet dyes commonly used as
mixtures.

An example is C.I. Acid Black 48, C.I. 65005 [1328-24-1] (66).

A novel azo-anthraquinone dye made through innovative enzymatic process has


been the subject of a recent publication (Enaud et al., 2010).

> Read full chapter

Natural dyes
B.H. Patel, in Handbook of Textile and Industrial Dyeing, 2011

11.3.1 Anthracenes
The two major groups of the anthracenes contain several well-known dyes.

Anthraquinones (yellow, pink and red pigments) (Saidman et al., 2002) include
Morinda citrifolia (Morindone), Alkanna tinctoria, Dactylopius coccus (cochineal)
(carminic acid), Laccifer lacca (Lac insect) (Laccaic acid), Kermes ilicis (shield louse)
(Kermisic acid), Alnus glutinosa, Rubia cordifolia, Rubia tinctoria, etc. (Alizarin) (Ru-
berethric acid). The structures of some anthraquinone dyes are shown in Fig. 11.1.

11.1. Structures of anthraquinone dyes.

Naphthoquinones (brown, pink or purple pigments) include Juglans nigra, Juglans


regia (Juglone), Lawsonia innermis (Lawsone), Woodfordia fruticosa, Lawsonia alba
(Lawsone), A. tinctoria (Alkannin) and Lithospermum erythorhizon (Shikonin). Natural
chromogens of these dyes are shown in Fig. 11.2 (Gulrajani et al., 1992c).
11.2. Structures of some naphthoquinone dyes.

> Read full chapter

Heterocyclic Dyes and Pigments


D.R. Waring, in Comprehensive Heterocyclic Chemistry, 1984

1.12.4.3 Anthraquinone Dyes


Anthraquinone dyes are very important generally in dye chemistry and depending
upon the nature of the substitution provide bright red to greenish-blue dyes. There
are, however, few heterocyclic analogues of importance in disperse dye chemistry.
However, an interesting exception is provided by the series of dyes derived by annel-
lation of heterocyclic rings in the 2,3-position of the anthraquinone nucleus as in the
N-alkylated imides (58) which are beautiful bright greenish-blues 56USP2753356
hitherto unmatched by any azo dye structure.

The dyes may be prepared by treatment of 1,4-diamino-2,3-dicyanoanthraquinone


with sulfuric acid followed by alkylation of the imide by primary alkyl amine ex-
change. Alternatively, the sulfuric acid treatment may be carried out in the presence
of a secondary alcohol (Scheme 11).
Scheme 11. Synthesis of anthraquinoneimide dyes

> Read full chapter

Dyeing of synthetic fibres


A.K. Roy Choudhury, in Handbook of Textile and Industrial Dyeing, 2011

Anthraquinone dyes
Anthraquinone dyes are a large and important group covering the hue range from
greenish yellow to bluish green. However, these dyes are particularly valued for
the production of bluish red, violet, blue and bluish green dyeing. The simple
anthraquinone dyes are bright and stable under dyeing conditions. Dyes of this class
will probably not increase their market share in the near future because:

1. They are somewhat weaker in tinctorial value than many azo dyes.

2. Several intermediate production stages increase capital investment with con-


sequent increase in dye cost.
3. Many dye intermediates required for their production are derived from an-
thraquinone- -sulphonic acid,which requires a mercury catalyst. This may
create pollution problems and alternative methods are to be established.

Examples are C.I. Disperse Violet 1 (Fig. 2.2) and C.I. Disperse Blue 56 (Fig. 2.3). A
few other anthraquinone disperse dyes are:

1. C.I. Disperse Yellow 77

2. C.I. Disperse Oranges 5, 6

3. C.I. Disperse Red 86

4. C.I. Disperse Blue 73.


Anthraquinone dyes containing primary or secondary amino groups are susceptible
to fading by burnt gas fumes, e.g. when 1, 4, 6, 8 tetra-aminoan-thraquinone dye
is used for dyeing and printing on secondary acetate and triacetate. Introduction of
electronegative groups (e.g. –OH) ortho to the amino groups, arylation of the amino
groups (as in the case of C.I. Disperse Blue 27), and/or omission of amino groups
altogether, increases resistance to burnt gas fumes. The sublimation fastness of the
dye is also improved due to the hydroxyethyl group attached to the pendant phenyl
ring in the above dye. Similarly, C.I. Disperse Blue 73 has very good fastness to light
and adequate sublimation fastness to most of the end-use. However, if the molecule
becomes weakly acidic, it becomes unstable to alkali with consequent deterioration
in dyeing properties.

Quinizarin is a traditional and readily available anthraquinone intermediate, which


condenses readily with arylamines particularly in the presence of a proportion of
leuco-quinizarin to produce violet to green dyes for polyester, e.g. C.I. Disperse
Violet 27.

> Read full chapter

Photoactive chemicals for antimicro-


bial textiles
J. Zhuo, in Antimicrobial Textiles, 2016

11.3.2 Anthraquinone and benzophenone derivatives


Anthraquinone and its derivatives have been used as photo-initiators in photopoly-
merizations of vinyl monomers for years owing to their high photoactivity. Recently,
Liu and Sun (2011b) indicated that photoactive anthraquinone derivatives were
able to generate considerable amounts of ROS, especially hydroxyl radicals and
hydrogen peroxide, through type I photosensitization under UVA light (365 nm).
The photoactivity of anthraquinone derivatives and the efficiency of generat-
ing ROS in the system mostly depend on the structures of the compounds and
the properties of the polymer substrate they are applied onto. Studies show that
anthraquinone derivatives with electron-withdrawing groups (eg, carboxylic acid
and sulfate groups) or weak electron-donating groups result in higher ROS yield
compared with those possessing strong electron-donating substitutions, such as
hydroxyl or amino groups (Zhuo and Sun, 2015). It is known that the frontier
orbital, the carbonyl group, in the structure will go through n to π or π to π
transition to the excited singlet state upon light excitation and quickly transfer to
the triplet state through intersystem crossing. The excited molecules in the triplet
state show a biradical feature, which is active to abstract vulnerable hydrogens in
the environment and form semiquinone radicals. Under aerobic conditions, the
semiquinone radicals would be trapped by oxygen molecules to produce the original
anthraquinone structure with the accompanying formation of hydrogen peroxide
and hydroxyl radicals. The photosensitization reactions and the generation of ROS
are shown in Fig. 11.6.

Figure 11.6. Photochemical reaction mechanisms of anthraquinone in the presence


of oxygen to generate reactive oxygen species. ISC, intersystem crossing.

Furthermore, compared with H2O moisture trapped on the textile surface, polymer
substrates containing weak C–H or N–H bonds in structures can serve as better
hydrogen donors. Materials with such structural features include cellulose-based
polymers like cotton, polyamides such as nylon 6 and nylon 6,6, and protein fibers
like wool and silk (Hong and Sun, 2008a; Liu and Sun, 2011a; Zhuo and Sun, 2013,
2014). In the case of cellulose-based materials, the bond dissociation energy of the
position-1 C–H bond in the glucose ring is low so that it can be easily cleaved
by semiquinone radicals, and the formed radicals are stabilized by nearby oxygen
molecules (Hong et al., 2009; Zhuo and Sun, 2014). The vulnerable hydrogens in
polyamides and protein fibers are possibly hydrogen donors in nylons and wool.

The photoactive derivative 2-anthraquinone carboxylic acid has been successfully


bonded onto cotton fabrics through esterification reactions (Liu et al., 2011). In addi-
tion, treated materials exhibited excellent photo-induced antibacterial effects against
both gram-negative E. coli and gram-positive S. aureus deposited on the surface.
Significant reductions in both bacteria were detected on treated cotton fabrics after
1 h UVA light (365 nm) exposure. Most interestingly, anthraquinone derivatives are
the second most popular class of colorants based on chemical structures, providing
a full spectrum of colors. Many natural colorants contain the anthraquinone group,
particularly those produced from aloe latex, fungi, senna, and some insects. Such
feature could be advantageous for the preparation of antibacterial textile fabrics
because the potential coloration and antimicrobial function could be achieved in a
one-step conventional textile dyeing process without any additional treatment. In
addition, the strong chemical interactions between dyed anthraquinone photosen-
sitizers and polymers can contribute to the high stability of the antibacterial agent
on the textile surface compared with photocatalyst NPs and can further improve the
reusability of the treated materials.

Photoactive 2-ethyl-anthraquinone and a real anthraquinone dye, Vat Yellow GCN,


have been efficiently applied onto cotton fabrics by the vat dyeing method (Zhuo
and Sun, 2013). The anthraquinone compounds were first reduced to the water-sol-
uble leuco form, which can be easily exhausted onto cellulose materials through
the vatting process. After undergoing exhaustion, the leuco salts were oxidized back
into the initial water-insoluble structure inside the fibers. Both modified cotton
fabrics exhibited light-triggered antibacterial function against E. coli and S. aureus,
and the bacterial inactivation efficiency depended on the concentration and type
of photosensitizers applied on the surface. No biocidal effects were detected on
pristine cotton fabrics or the treated fabric without UVA irradiation. Liu and Sun
(2011a) also successfully treated nylon 6 with photoactive 2-anthraquinone sulfonate
sodium, 2,6-anthraquinone sulfonate, and 2,7-anthraquinone sulfonate through an
acid dyeing process. Although the antibacterial properties of those treated textiles
were not indicated in this research, considerable amounts of hydrogen peroxide and
hydroxyl radicals have been detected on the fabrics with the assistance of UVA light,
which may also render the materials antimicrobial.

Benzophenone derivatives, which possess a chemical structure similar to that of


anthraquinones, also exhibit photoactivity and provide potential application as an-
tibacterial agents owing to the same photochemistry mechanism but most under UV
light. This class has been widely used for synthetic perfumes and as a precursor of
dyes, pesticides, and medicines (Knowland et al., 1993; Jeon et al., 2008). Research
has achieved excellent bacterial inactivation on textile materials with activated ben-
zophenone groups. 4-Hydroxylbenzophenone was grafted onto cotton fabrics with
1,2,3,4-butanetetracarboxylic acid as a cross-linker (Hong and Sun, 2008a, 2011).
Significant colony reductions of E. coli and S. aureus were observed on treated
cotton fabrics under UVA light. Similar results were obtained on benzophenone
derivative-grafted polystyrene film (poly(styrene-co-vinylbenzophenone)) (Hong and
Sun, 2008b).

> Read full chapter


Anthraquinone Dyes
Ghodsi Mohammadi Ziarani, ... Hendrik G. Kruger, in Metal-Free Synthetic Organic
Dyes, 2018

Abstract
Anthraquinone (AQ) derivatives are the largest group of natural quinones. Other
natural quinones are naphthoquinones and benzoquinones. AQs also constitute
the largest group of natural pigments with about 700 compounds described. AQ
1 as a building block of many dyes is attractive because of its superior stability
and light-fastness properties. Synthetic dyes such as alizarin are often derived
from 9,10-anthraquinone (Bien et al., 2000) [1]. Important derivatives include 1-ni-
troanthraquinone, anthraquinone-1-sulfonic acid, and dinitroanthraquinone (Vo-
gel, 2000) [2]. The potential of AQs to inhibit or prevent fungal and bacterial growth
are two interesting biological properties, but many other properties are attributed to
them, such as antioxidant, anticancer, antiinflammatory, laxative, and many others.
This chapter describes some synthetic AQ dyes.

> Read full chapter

Antibacterial colorants for textiles


F. Alihosseini, G. Sun, in Functional Textiles for Improved Performance, Protection
and Health, 2011

Anthraquinones
Anthraquinones are the largest class of naturally occurring quinones and contain
some of the most important natural colorants such as alizarin, purpurin, munjistin,
emodin, chrysophanol, aloe-emodin, physcion, rhein, etc. They exist in the form of
hydroxyanthraquinones and usually have 1–3 hydroxyl groups. These quinone dyes
could form complexes with different metal salts resulting in very good color fastness
(Cai et al., 2006; Huang and Cai, 2009).

Alizarin and purpurin (Fig. 17.5) are two main anthraquinone-type colorants found
in the root and tubers of Rubia tinctorum (Common Madder), R. peregrine (Wild
Madder), R. cordifolia (Indian Madder), and R. munjista. Madder has been cultivated
and used as a source of red dyes in Asia, Europe and America. In fact the oldest
sample of madder dyed cotton found in Egypt belongs to the third millinium bc
(De Santis and Moresi, 2007; Puchalska et al., 2003). Other colorants found in these
plants include munjistin, pseudopurpurin, xanthopurpurin, ruberythric acid and
rubiadin (Sanyova and Reisse, 2006).

17.5. Chemical structure of (a) alizarin and (b) purpurin.

Alizarin and purpurin were found in glycoside or monosaccharide form in the plants.
Part of alizarin can be extracted directly from the root of three-year-old plants while
the higher yield of colorants especially purpurin could be obtained from the plants
after two years’ storage. During this storage time fermentation occurs and enzymes
hydrolyze the glycoside and yield the free anthraquinone red dyes. In fermentation
process particles stick together and form a solid mass called lake. The enzymatic
process is considered the best since acidic or basic hydrolysis leads to the formation
of a mutagenic compound called lucidin (Bechtold and Mussak, 2009; Chenciner,
2000). Madder has been used to dye wool, leather and cotton and can also be used
as a food colorant. Alizarin known as Pigment Red 83 was one of the first colorants
synthesized and commercialized in 1863 (Chenghaiah et al., 2010; De Santis and
Moresi, 2007).

Alizarin and purpurin have antimicrobial and antifungal activity against different
pathogenic bacteria (Bechtold and Mussak, 2009; Chenciner, 2000). Wool fibers dyed
with madder have shown higher insect resistance against carpet beetles (Park et al.,
2005). Purpurin even reveals an antigenotoxic effect against a range of different
environmental carcinogens such as hetrocyclic amines (Marczylo et al., 1999). In
medicinal application, madder has been used in skin care products with astringent,
tonic, vulnerary and antiseptic functions. It can be used to clean open wounds and
treat skin diseases, especially tubercular conditions of the skin and mucous tissues.
Several studies on the safety of these compounds as phytopharmaceutical and food
colorants have shown that only lucidin among these structures is mutagenic (Dweck,
2002).

Carminic acid (Fig. 17.6) is another important hydroxyanthraquinone-based col-


orant, which is derived from cochineal and has been used to dye wool, silk and
cotton fabrics. Its brilliant red color with great light fastness could suppress many
similar synthetic ones used in textiles (Taylor, 1986). Carminic acid consists of an
anthraquinone with a sugar moiety attached to it. Treatment of carminic acid with an
aluminum salt produces aluminum lake or carmine. Carminic acid is water soluble,
and an acid stable colorant. Its aluminum lake, carmine, is soluble in alkaline media
and has higher stability to heat, light and oxygen. In alkaline conditions carmine
provides a blue-red shade (Cosentino et al., 2005).

17.6. Chemical structures of (a) carminic acid and (b) kermesic acid.

Although cochineal has shown antimicrobial activity against S. aureus, the dyed wool
fabric did not provide any antimicrobial activity. However, the wool fabrics dyed with
cochineal and some metal mordant revealed up to 100% antimicrobial functions
depends on the type of metals (Bae and Huh, 2006).

> Read full chapter

Separation of Natural Products


M. Ganzera, A. Murauer, in Supercritical Fluid Chromatography, 2017

15.2.4 Miscellaneous
Anthraquinones represent an important class of natural products, because they
are the active ingredients in popular laxatives like senna (Cassia angustifolia, Cassia
acutifolia), aloe (Aloe ferrox, Aloe barbadensis) and rhubarb (Rheum officinale, Rheum
palmatum). The major aglyca (chrysophanol, physcion, emodin, aloeemodin, and
rhein) in rhubarb could be resolved in less than 5 minutes on a SFC-specific
stationary phase (Acquity UPC2 HSS C18 SB, Waters) with a small particle size
(1.8 µm), in combination with a mobile phase containing CO2, methanol, and 0.05%
diethylamine [29]. Gradient elution at a flow rate of 2 mL/min was used for the
separation. The method was selective, precise, and accurate (recovery rates from
95.4% to 103.1%), with an LOD below 0.5 ng on-column, and was suitable for the
analysis of plant material without any specific sample clean-up. Quantitative results
were reproducible ( rel≤2%) and in agreement with published data (total content in
samples before hydrolysis 0.32–0.73%).

Destruxins are cyclic hexadepsipeptides with insecticidal and phytotoxic activities.


They are produced as secondary metabolites by the entomopathogenic fungus
Metarhizium brunneum, which is used as biological pest control agent against
cockchafers and garden-chafers. To facilitate the analysis of destruxins SFC was
evaluated and this approach permitted the identification of 13 derivatives in purified
M. brunneum culture broth [30]. Five compounds (destruxins A, B, D, and E, as
well as destruxin E-diol) were assigned using reference compounds, the tentative
identification of other substances was possible by SFC-DAD-MS/MS. For the latter
an Acquity UPC2 was coupled to a Xevo TQD triple quadrupole mass spectrometer
operated in the positive ESI mode; methanol was added as make-up solvent. The
mobile phase was the same as for stand-alone SFE experiments (CO2 and an 8/2
mixture of methanol and acetonitrile), only for SFE-MS 0.02% formic acid was added
to the modifier instead of 0.02% TFA. Method validation and quantification of the
compounds was achieved using a diode-array detector. Compared to the established
UPLC method the authors noted many advantages (shorter analysis time, rapid
equilibration, higher throughput), but also a slightly lower sensitivity (LOD by UPLC:
0.1–1.4 µg/mL; LOD by SFC: 4.4–6.5 µg/mL). They explained this observation by the
low injection volume of 1 µL and the diminished sensitivity of SFC-DAD.

The fact that chirality may determine pharmacological activity is relevant for natural
products as well. One out of many examples is Isatis indigotica, a plant also known
as woad. A blue dye can be produced from its leaves and the roots (“Ban Lan Gen”
in Chinese) are an important remedy in Asian traditional medicine to treat fever.
They contain two sulfur-containing oxazolidine enantiomers, R- and S-goitrin, as
active ingredients, which differ significantly in their biological activity. The R-variant
(epigoitrin) is antiviral, whereas S-goitrin is goitrogen (i.e. it increases thyroid tissue
due to disrupted production of thyroid hormones). For quality control purposes,
it is important to quantify both enantiomers. SFC with a chiral stationary phase, a
(S,S)-Whelk-O1 column (4.6×50 mm, 10 µm) and a simple CO2-methanol gradient
enabled their separation in less than 6 minutes [31]. Detection was performed at
244 nm (DAD) and also by a single quadrupole MS with APCI source (positive
mode). Not surprisingly, the sensitivity of the latter was 50-times higher, but other
validation parameters (correlation coefficients, repeatability, precision, etc.) were
comparable. Three commercial formulations all labeled as “Ban Lan Gen” powder
were quantitatively analyzed. One of them contained none of the goitrins, in the
remaining R-goitrin was dominant. However, with a maximum content of 0.03% its
concentration was rather low.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like