You are on page 1of 15

Algal Research 56 (2021) 102330

Contents lists available at ScienceDirect

Algal Research
journal homepage: www.elsevier.com/locate/algal

Update on sesquiterpenes from red macroalgae of the Laurencia genus and


their biological activities (2015–2020)
Ana-Marija Cikoš a, Mladenka Jurin b, Rozelindra Čož-Rakovac b, Dajana Gašo-Sokač a,
Stela Jokić a, Igor Jerković c, *
a
Faculty of Food Technology Osijek, University of Josip Juraj Strossmayer in Osijek, Franje Kuhača 18, 31000 Osijek, Croatia
b
Ruđer Bošković Institute, Bijenička 54, 10000 Zagreb, Croatia
c
Department of Organic Chemistry, Faculty of Chemistry and Technology, University of Split, Ruđera Boškovića 35, 21000 Split, Croatia

A R T I C L E I N F O A B S T R A C T

Keywords: Laurencia genus has been the most promising source of sesquiterpenes (among all macroalgae) including those
Laurencia genus with high degree of halogenation, especially bromination. Related published reviews do not report novel re­
Macroalgae searches included in this study and are not focused exclusively on this genus or sesquiterpenes. The present
Marine natural products (different
review reports novel structures of Laurencia derived sesquiterpenes exclusively in the period of 2015–2020, with
sesquiterpene skeletons)
emphasis on their structural diversity and different skeleton types including chemical biomarkers along with the
Bioactive compounds (cytotoxic, anti-
inflammatory, antibacterial, antifungal, antidi­ methods for their structural analysis. Their cytotoxic, anti-inflammatory, antibacterial, antifungal, antidiabetic,
abetic, anthelmintic, others) anthelmintic activities and others are summarised indicating determined structure – activity relationships along
with ecological role of targeted sesquiterpenes. Furthermore, possible stages of sesquiterpene biosynthesis are
presented for better understanding of their chemodiversity.

1. Introduction miscellaneous triterpenes), around 240 C15 acetogenins (acetogenins


containing tetrahydrofuran ring and tetrahydropyran ring; acetogenins
The genus Laurencia (Rhodophyta, Ceramiales, and Rhodomelaceae) containing oxepane ring; acetogenins containing oxocane ring; aceto­
comprise around 140 species distributed throughout the warm sea wa­ genins containing nine- or ten-membered cyclic ether ring; acetogenins
ters, particularly from temperate to tropical shores [1]. It has been rich containing twelve-membered cyclic ether ring; linear acetogenins),
source of unique specialised metabolites with high degree of haloge­ around 30 indols, 5 aromatic compounds, 15 steroids, and 35 miscel­
nation, especially bromination, and wide range of the skeleton types [2]. laneous metabolites. Among them, biologically active metabolites have
Within more than 60 species investigated, 700 compounds with unique been isolated with anticancer [6,7,8], antidiabetic [9], antibacterial
structures have been isolated from this genus [3] with sesquiterpenes as [10], antimalarial [11], anthelmintic [12] and other biological activ­
the most abundant terpenes along with diterpenes, triterpenes, steroids, ities. Harizani et al. [5] reported their cytotoxic activity, antibacterial
alkaloids and C15 acetogenins [4]. The comprehensive review of Har­ and antifungal activity, antiviral activity, activity against parasites and
izani et al. [5] presents the chemical diversity and the bioactivity of the their vectors, anti-inflammatory activity, miscellaneous biological ac­
secondary metabolites isolated from species of the genus Laurencia tivities, ecological functions, antifeedant activity and toxicity to marine
including more than 500 sesquiterpenes, about 100 diterpenes (par­ organismsm, brine shrimp toxicity, and antifouling activity. The pro­
gueranes, isopargueranes, and related diterpenes; labdanes, pimaranes, duction of specialised metabolites observed in this genus can be
and related diterpenes; irieanes and neoirieanes; dactylomelanes; explained in general as an ecological adaptive response to the environ­
obtusanes, 15,14-friedoobtusanes, and related diterpenes; prevezanes mental changes and oscillations. The specific metabolites are produced
and neorogiolanes; miscellaneous diterpenes), about 50 triterpenes only by certain Laurencia species indicating them as chemical markers.
(triterpenes possessing 2,7-dioxabicyclo[4.4.0]decane ring system and Several reviews provide information about sesquiterpenes found in
related metabolites; triterpenes possessing 2,8-dioxabicyclo[5.4.0] different macroalgae [13,14,15]. Phytochemical and biological prop­
undecane ring system; triterpenes possessing symmetry elements; erties of sesquiterpenes isolated from Laurencia genus were presented by

* Corresponding author.
E-mail address: igor@ktf-split.hr (I. Jerković).

https://doi.org/10.1016/j.algal.2021.102330
Received 6 February 2021; Received in revised form 21 April 2021; Accepted 24 April 2021
Available online 12 May 2021
2211-9264/© 2021 Elsevier B.V. All rights reserved.
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

Al-Massarani [13], while Jesus et al. [15] reviewed haloaryl specialised metabolites: elatol and obtusol belonging to the first type of metabolites,
metabolites with their biological applications including sesquiterpenes and [1(15)Z,2Z,4R,8S,9R]-8,15-dibromochamigra-1(15),2,11(12)-trien-
isolated from different macroalgae. Modzelewska et al. [16] reviewed 9-ol and [1(15)E,2Z,4R,8S,9R]-8,15-dibromochamigra-1(15),2,11(12)-
anticancer activity of naturally occurring sesquiterpenes. Le Bideau trien-9-ol belonging to the second type. The authors discovered the
et al. [14] systematically represented tricyclic sesquiterpenes from ma­ presence of chamigrane metabolites with ketone functionality in the
rine origin. The biosynthesis of sesquiterpenes was investigated in detail compounds 1, 2, 3, laurencenone B and 2-chloro-3-hydroxy-α-chamigran-
mainly for the compounds occurring in plants with wide variety of 9-one. They suggested that newly discovered metabolites can serve as
skeleton types [17,18], while Miller and Allemann [19] reviewed intermediates in the biosynthetic pathway of specialised metabolite pro­
sesquiterpene synthases for the better understanding of chemical shifts duction. Recently, undescribed oxygenated sesquiterpenes lau­
occurring during the biosynthesis. remantanones A (4) and B (5) were determined in L. majuscula along with
Nevertheless, the discovery of novel and interesting structures has known compounds (+)-elatol [29], obtusol [23], (+)-laurencenone B and
been occurring in new studies of Laurencia genus and therefore present 2-chloro-3-hydroxy-α-chamigran-9-one, as well as dendroidiol [30] and
review is providing update on Laurencia derived sesquiterpenes. The cartilagineol [31]. The compounds 4 and 5 were formed from α-chami­
present study addresses the following topics: a) exclusively sesquiter­ grane backbone, while the compounds (+)-elatol, obtusol, dendroidiol,
penes found in the genus Laurencia (as the main producer of sesquiter­ cartilagineol were formed from β-chamigrane backbone [10]. In the same
penes among all macroalgae) are targeted in the period 2015–2020 and year, Kamada et al. [32] discovered (− )-laurencenone D (6) and
presented by different skeleton types; b) the methods of their isolation (+)-deschloroelatol (7) which were not detected before in the same spe­
(including different extraction solvents) and analysis are systematically cies, along with the previous known compounds (+)-elatol, (+)-lau­
presented; c) their cytotoxic, anti-inflammatory, antibacterial, anti­ rencenone B, 2-chloro-3-hydroxy-α-chamigran-9-one and cartilagineol
fungal and other activites are addressed as well as their ecological roles; suggesting the unique diversity of halogenated chamigrane sesquiterpenes
d) the structure influence on bioactivity with the emphasis on cytotox­ and their rearranged analogues in L. majuscula. Red alga L. composita
icity of targeted sesquiterpenes is presented; e) probable stages of produced eleven highly halogenated chamigrane sesquiterpenes (com­
sesquiterpene biosynthesis are indicated for better understanding of positacins A-K (8–18)) and all of the metabolites contained unusual
their chemodiversity as well as general enzymes included in the for­ rearranged compositacin A (8) with specific composition including an
mation of different sesquiterpene classes (since in other reviews this ether bridge involving C-5/C-9 and C-3/C-5 in compositacins B (9) and D
topic is not elaborated in detail or not presented, this study covers (16) while compositacins B (9) and C (10) were the first chamigranes with
broader range of available references); f) the literature references are C-10 carbonyl group [33]. The absolute configuration of compositacin B
updated. (9) was determined by ECD calculation, and the absolute configurations of
compositacins A (8) and C–K (10–18) were proposed on the biosynthetic
2. Chemical biodiversity of sesquiterpenes from genus Laurencia grounds by comparison to compositacin B (9) and related known sesqui­
terpenoids johnstonol [34] and yicterpene A [35]. It was also suggested
Laurencia genus has been the most attractive source of sesquiterpenes that the structure of the previously reported sesquiterpenoid laurokamin A
among all marine macroalgae due to its capability of biosynthesizing a should be revised. Hence, the study of Hu et al. [36] revealed the presence
wide variety of structurally diverse sesquiterpenes which are divided of another compositacins in L. composita, including compositacin L (19),
according to their skeleton types to chamigrane, cuparane, laurane, compositacin M (20) and compositacin N (21), along with known com­
bisabolane, aristolane, brasilane, eudesmane, snyderane, perforanes and pounds 2,10β-dibromochamigra-2,7-dien-9α-ol [37], 2,10-dibromo-3-
other skeletons (Table 1). Only new structures reported from 2015 are chloro-α-chamigrene [38], obtusane [39], deoxyprepacifenol [40], pre­
further presented in the following figures. In novel papers, sesquiterpene pacifenol epoxide [41], pacifidiene, pacifenediol, pacifenol [42], and
structures were elucidated on the basis of extensive spectroscopic yicterpene A. Compositacin L (19) represents the third example of cha­
analysis including FT-IR, H1-NMR, C13-NMR, H1-H1-COSY, HSQC, migranes having C-10 carbonyl group. The study of Chen et al. [43] re­
HMBC, and NOESY. The absolute configurations were determined or ported the isolation of 18 compounds from the red alga L. tristicha
proposed based on the biosynthetic grounds by the comparison to including eight new chamigrane-type sesquiterpenes: tristichone A (22),
related known sesquiterpenes. ristichone B (23), tristichol A (24), tristichol B (25), tristichol C (26),
zristichol D (27), tristichone C (28) and tristichone D (29), along with nine
2.1. Chamigrane skeleton sesquiterpenes known compounds, halogenated chamigrene derivatives: 9-(E)-bromo­
methylidene-1,5,5-trimethylspiro[5.5]undeca-1,7-dien-3-one and 9-(Z)-
Chamigrane-type sesquiterpenes are the most widespread sesqui­ bromomethylidene-1,5,5-trimethylspiro[5.5]undeca-1,7-dien-3-one [44],
terpenes from the genus Laurencia. Chamigrane skeleton can serve as an ma'ilione, [1(15)Z,2Z,4S,8R,9S]-8,15-dibromochamagra-1(15),2,11(12)-
intermediate for formation of novel carbon skeletons. The presence of trien-9-ol [44], [1(15)E,2Z,4S,8R,9S]-8,15-dibromochamagra-1(15),2,11
both α- and β-chamigrane in marine macroalgae suggests that derived (12)-trien-9-ol [44], ma'iliohydrin [45], isorigidol [12], allo-isoobtusol
carbon skeletons are formed by intramolecular cyclization of both cha­ [46], and majusculone [47]. Philippus et al. [48] used molecular
migrane backbones. However, for terrestrial plants related skeletons are networking for determination of sesquiterpenes of some algae from Lau­
usually formed only from α-chamigrane backbone [20]. rencia genus (L. catarinensis, L. dendroidea and L. intricata). An interesting
Even though there are many reports for chamigrane-type sesquiter­ chemodiversity was observed in L. catarinensis collected from two nearby
penes [12,21,22,23,24], recent studies are revealing new undescribed islands: L. catarinensis collected in Xavier Island contained three chami­
compounds pointing out the diversity of sesquiterpenes produced by grane sesquiterpenes including prepacifenol epoxide, johnstonol and
Laurencia genus (Fig. 1). Kamada et al. [7] reported three undescribed pacifenol along with some C15-acetogenins; L. catarinensis collected in
α-chamigrane sesquiterpenes with ketone functionality from L. majuscula Arvoredo Island did not contain C15-acetogenins, but contained four
(7-aldehydelaurencenone B (1), 2-chloro-3-methoxy-α-chamigran-9-one different chamigrane sesquiterpenes including 9-hydroxy-4,10-dibromo-
(2), laureborneone (3)) along with five other known metabolites [1(15) 3-chloro-α-chamigrene [49], 2,10-dibromo-3-chloro-8-hydroxy-β-chami­
Z,2Z,4R,8S,9R]-8,15-dibromochamigra-1(15),2,11(12)-trien-9-ol, [1(15) grene [49], 4,10-dibromo-3-chloro-7,8-epoxychamigrane [50] and 4,10-
E,2Z,4R,8S,9R]-8,15-dibromochamigra-1(15),2,11(12)-trien-9-ol, ma'i­ dibromo-3-chloro-7,8-epoxy-5-hydroxychamigrane [51]. These results
lion [12], laurencenone B [25] and 2-chloro-3-hydroxy-α-chamigran-9- indicate that a different chemical composition could be present in the alga
one [26]. Prior to this study, they reported [27,28] that the population of in relation to the collection sites.
L. majuscula produces two major types of halogenated specialised

2
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

Table 1
Sesquiterpenes isolated from the species of genus Laurencia.
Laurencia species Determined compounds Extraction Analytical method Bioactivity References
solvent

Chamigrane skeleton sesquiterpenes


L. majuscula 7-Aldehyde-laurencenone B, 2-chloro-3-methoxy- CH3OH FT-IR, 1H and 13C NMR, Cytotoxic activity [7]
1 –1
α-chamigran-9-one, laureborneone H H COSY, HSQC, HMBC,
NOESY
1
Lauremantanones A and B H and 13C NMR, HRESIMS [10]
1
(− )-Laurencenone D and (+)-deschloroelatol H and 13C NMR, HRESIMS Antibacterial activity [32]
1
L. composita Compositacins A-K C2H5OH/ H and 13C NMR, HSQC, – [33]
1 –1
H2O H H COSY, ROESY, HMBC,
HREIMS, LREIMS
1
L. composita Compositacins L-N CH3OH H and 13C NMR, EIMS, Antibacterial and [36]
HREIMS, 1H–1H COSY, DEPT, antifungal activities
HSQC, HMBC, ROESY
1
L. tristicha Tristichone A, ristichone B, tristichol A, tristichol B, tristichol C6H14 H and 13C NMR, HRESIMS, Cytotoxicity, antibacterial [43]
C, zristichol D, tristichone C and tristichone D NOESY, COSY, HMBC and anti-inflammatory
activities

Bisabolane skeleton sesquiterpenes


1
L. composita Laurecomposins A and B CH3OH H and 13C NMR, EIMS, Antibacterial and [36]
HREIMS, 1H–1H COSY, DEPT, antifungal activities
HSQC, HMBC, ROESY

Cuparane skeleton sesquiterpenes


13
L. obtusa α-Bromocuparene and α-isobromocuparene C4H8O2 C NMR, HMBC, ESI-IT – [54]
1
L. johnstonii C2H5OH H and 13C NMR, HREIMS, Anti-acanth-amoeba [55]
HMBC, NOE, COSY activity
1
L. tristicha α-Bromocuparene C2H5OH H and 13C NMR, EIMS – [56]
1
L. nidifica α-Bromocuparene CH3OH H NMR, MS Repellent and insecticidal [74]
activities
L. majuscula Cuparen-3-ol C6H14 NMR and MS Cytoxic activity [57]
1
L. tristicha 4α-Hydroxy-bromocuparene C6H14 H and 13C NMR, HRESIMS, Cytotoxicity, antibacterial [43]
NOESY, COSY, HMBC and anti-inflammatory
activities
1
L. natalensis 8-Deoxyalgoane, 1-deacetoxyalgoane and algoane CH2Cl2/ H and 13C NMR, 1H–1H Anti-inflammatory and [9]
CH3OH COSY, ROESY, HSQC, HMBC, antidiabetic activity
ESIMS, APCIMS
1
L. obtusa 10-hydroxy-cuparaldehyde, 3-hydroxy-15-nor-cuparan-10β- CH2Cl2/ H and 13C NMR, EIMS, Antibacterial and [58]
ol, 2-bromo-3-hydroxy-15-nor-cuparan-10β-ol CH3OH HRESIMS anticandidal activities

Laurane skeleton sesquiterpenes


1
L. okamurai Nanji A CH2Cl2/ H and 13C NMR, LREIMS, – [59]
CH3OH HREIMS, 1H–1H COSY,
HMQC, HMBC
1
L. tristicha Laurinterol C2H5OH H and 13C NMR, EIMS – [56]
L. tristicha and L. CH3OH HPLC-DAD – [68]
okamurai
13
L. obtusa Laurinterol, iso-laurenisol, laurene, 3,7-dihydroxy-dihydro­ C4H8O2 C NMR, HMBC, ESI-IT – [54]
laurene, 11-iodolaurinterol
1
L. nidifica Laurinterol, isolaurinterol, aplysin CH3OH H NMR, MS Repellent and insecticidal [74]
activities
1 13
L. johnstonii C2H5OH H and C NMR, HREIMS, Cytotoxic and anti-acanth- [55]
HMBC, NOE, COSY amoeba activity
1
L. okamurai 3α-Hydroperoxy-3-epiaplysin, debromo-3β-hydroperoxy- C3H6O H and 13C NMR, COSY, PTPB inhibitory activity [75]
aplysin, 3β-hydroperoxyaplysin, isoaplysin, 10-bromoisoalpy­ HMBC, NOESY, HREIMS and antibacterial activities
sin, laurepoxyene, laurokamurene A, isolaurinterol
1
L. heteroclada 2-Bromo-3,5,6-trihydro-1,4-dihydroxy isolaurene CH3OH H and 13C NMR, HREIMS, – [76]
HMQC, 1H–1H COSY, HMBC,
NOESY
1
L. mariannensis Teanol CH3OH H and 13C NMR, HRESIMS, [77]
HMBC, NOESY
1
L. okamurai Laurokamurols A − C C3H6O H and 13C NMR, HREIMS, PTP1B inhibitory activity [78]
HMBC, NOESY, HMQC,
TDDFT-ECD
1
L. nangii Neolaurene CH3OH H and 13C NMR, HRESIMS, Cytotoxic and [79]
NOESY, HBMC antibacterial activities
1
L. tristicha 10-Hydroxyepiaplysinol, epiaplysinol C2H5OH H and 13C NMR, HRESIMS, Antioxidant activity [81]
HSQC, HMBC, NOESY

Aristolane skeleton sesquiterpenes


1
L. complanata Debilone MeOH H and 13C NMR, COSY, – [82]
NOESY, HMBC, HSQC, ESIMS

Brasilane skeleton sesquiterpenes


13
L. obtusa 4-Hydroxy-5-brasilene C4H8O2 C NMR, HMBC, ESI-IT – [54]

Eudesmane skeleton sesquiterpenes


L. obtusa Antifungal and cytotoxic [84]
(continued on next page)

3
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

Table 1 (continued )
Laurencia species Determined compounds Extraction Analytical method Bioactivity References
solvent
1
Eudesma-4(15),7-diene-5,11-diol, teuhetenone, chabrolidione CH2Cl2/ H and 13C NMR, EIMS,
B CH3OH HRESIMS, DEPT, HSQC,
1 –1
H H COESY
1
L. pinnata 1β-B -11-en-4α-ol, 1β-bromo-4α,5α-epoxyselinane, and 1β- CHCl3 H and 13C NMR, EIMS, – [85]
bromoselin-3,11-diene, 1β-bromo-6,8-cycloselin-4(15)-ene, HRAPPIMS, DEPT, HSQC,
1β-bromoselin-4(15),11-diene HMBC, NOE, 1H–1H COSY
1
L. obtusa Eudesma-4(15),11-diene-5,7-diol CH2Cl2/ H and 13C NMR, EIMS, Antibacterial and [58]
CH3OH HRESIMS anticandidal activities

Snyderane skeleton sesquiterpenes


13
L. obtusa β-Snyderol C4H8O2 C NMR, HMBC, ESI-IT – [54]
1
L. obtusa 8-Keto-10-dehydrobromo-γ-snyderol, α-snyderol, β-snyderol,, C4H8O2 H and 13C NMR, HRESIMS, – [89]
8-keto-10-dehydrobromo-β-snyderol, 8-hydroxy-β-snyderol, HSQC, COSY, HMBC
1
L. intermedia Aplysistatin, palisadin B, 3,4-epoxypalisadin A, 2-hydroxylu­ C6H14 H and 13C NMR, HRESIMS, – [93]
zofuranone, 2-hydroxyluzofuranone B HMBC, HSQC,

Other skeletons sesquiterpenes


1
L. johnstonii 3α-Bromojohnstone C2H5OH H and 13C NMR, HREIMS, Anti-acanth-amoeba [55]
HMBC, NOE, COSY activity
1
L. intricata (+)-Cyclocolorenone CH3OH H and 13C NMR, ESIMS Repellent activity [91]
1
L. majuscula Omphalaurediol, rhodolaurenones B and C CH3OH H and 13C NMR, HRESIMS, – [32]
HSQC, 1H–1H COSY, HMBC,
NOESY
L. dendroidea Silphiperfolan-7β-ol CH2Cl2 – – [98]
1
L. snackeyi Snakeol, snakediol CH3OH H and 13C NMR, HRESIMS, Antibacterial activity [99]
1 –1
H H COSY, NOESY

2.2. Bisabolane and cuparane skeleton sesquiterpenes okamurai [59] including debromolaurinterol [63], laurinterol acetate
[60], laurinterol [64], 3β-Hydroperoxyaplysin, 3α-hydroperoxy-3-epi­
Bisabolane skeleton includes monocyclic ring structure resulted from aplysin [65], isolaurinterol [63] and laurokomurenene A [61]. It was
the cyclization of geranyl cation [52]. The recent investigation on L. suggested that the compounds debromolaurinterol, laurinterol and iso­
composita resulted in two new bisabolane-type sesquiterpenes, named laurinterol can be characteristic specialised metabolites of Laurencia
laurecomposins A (30) and B (31) (Fig. 2). The relative configurations genus because they were already detected in L. tristicha [66] and L.
for both compounds were elucidated, while for laurecomposin A abso­ intermedia [63] along with the compounds 3β-hydroperoxyaplysin, 3α-
lute configuration was also given by the modified Mosher's method [36]. hydroperoxy-3-epiaplysin and laurokomurenene A which can act as
Cuparane structure of sesquiterpenes is formed by the cyclization of chemotaxonomic markers because they were only detected in L. oka­
bisabolane skeleton between C-6 and C-11. Three methyl groups are in murai [61,65]. Laurinterol acetate was found the second time in nature
the aliphatic backbone, one is located at the position 1 and the other two and it was recorded in L. okamurai so it can be also used as the marker
methyls are at the position 2 [53]. α-Bromocuparene and α-iso­ [60]. L. okamurai and L. tristicha are very similar in morpohology as well
bromocuparene were found in different Laurencia species such as L. as in the content of specialised metabolites. It can be seen that some
obtusa [54] and L. johnstonii [55], while α-bromocuparene was obtained compounds already detected in L. okamurai such as debromolaurinterol,
from L. tristicha for the first time by Zhang et al. [56]. Later, α-bromo­ debromolaurinterol acetate, laurinterol acetate and laurinterol were
cuparene was also found in L. nidifica (Ishii et al. 2017) and L. majuscula detected in L. tristicha as well, along with some previously described
[57]. L. majuscula also showed the presence of cuparen-3-ol [57]. In compounds detected only in L. tristicha aplysin [67], aplysinol [60],
2016, Chen et al. [43] reported one new bromocuparane-type sesqui­ debromoaplysinol and aplysinal [67]. Therefore, it can be suggested that
terpene, 4α-hydroxybromocuparene (32) (Fig. 2) isolated from L. tris­ these two algae share the same metabolic pathway and possess closer
ticha. L. natalensis contained earlier undescribed cuparane genetic affinity. Moreover, laurinterol as the main compound found in
sesquiterpenes including 8-deoxyalgoane (33), 1-deacetoxyalgoane (34) both of the algae can serve as chemotaxonomic marker for these two
and algoane (35) [9]. New cuparane-type sesquiterpene, 10-hydroxycu­ species [56]. Zhang et al. [68] provided the overall chemical profiles of
paraldehyde (36), and two nor-cuparanes, 3-hydroxy-15-nor-cuparan- L. tristicha and L. okamurai obtained by high-performance liquid chro­
10β-ol (37) and 2-bromo-3-hydroxy-15-nor-cuparan-10β-ol (38), were matography (HPLC). The results showed that both algae are important
isolated for the first time from L. obtusa. Nor-cuparanes, 37 and 38, sources of laurane-type sequiterpenes due to the presence of laurinterol
isolated in that study are rarely found in marine organisms [58]. as the main sesquiterpene in both species and debromolaurinterol as the
second abundant compound. L. okamurai contained higher content of
2.3. Laurane skeleton sesquiterpenes laurinterol (14.96–15.42 mg/g), while L. tristicha contained higher
concentration of debromolaurinterol (1.15–1.38 mg/g). Consequently,
Contrary to cuparane-type sesquiterpenes, in laurane-type com­ laurane-type sesquiterpenes were once again shown as important
pounds methyls in the aliphatic portion are placed at the positions 1, 2 chemotaxonomic markers for L. tristicha and L. okamurai. Hence, an
and 3. Among all marine organisms, macroalgae belonging to the genus effective new workflow for absolute structure elucidation, in which
Laurencia are considered as the main producers of laurane-type sesqui­ crystallographic analysis is iteratively applied together with NMR
terpenes [53]. analysis was established by Wada et al. [69]. In their study they eluci­
Recently, nanji A was found for the first time in L. okamurai [59] and dated the structure of laurinterol by NMR-coupled crystalline-sponge
it represents rare dimeric sesquiterpene of the cyclolaurane-type but it (CS) method. In the research by Esselin et al. [54] a wide variety of
was already detected in L. microcladia [8]. Even though this alga was specialised metabolites were identified from L. obtusa extract, among
studied more times [60,61,62] there was no other report of this com­ them several compounds of laurane-type including laurinterol, iso-
pound and possible reason can be the collection time influencing the laurenisol [70], laurene [71], 3,7-dihydroxydihydrolaurene [72] and
metabolism. Previously detected compounds were also found in L. 11-iodolaurinterol [73]. Hence, L. nidifica [74] and L. johnstonii [55]

4
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

Fig. 1. Chamigrane skeleton sesquiterpenes from Laurencia.

revealed the presence of laurinterol, isolaurinterol and aplysin. New was determined by the NOESY spectrum [77]. Three novel hetero­
compounds including, debromo-3α-hydroperoxy-3-epiaplysin (39) and dimeric laurane-type sesquiterpenoids, laurokamurols A − C (43–45),
debromo-3β-hydroperoxyaplysin (40) (Fig. 3) were isolated from L. along with eight known related monomeric ones were isolated from the
okamurai along with already known compounds 3β-hydroperoxyaplysin, East China Sea red alga Laurencia okamurai Yamada [78]. The absolute
isoaplysin, 10-bromoisoalpysin [60], laurepoxyene [65], laurokamur­ configurations of new bis-sesquitepenoids, especially their axial
ene A and isolaurinterol. The compound 39 was elucidated as debro­ chirality, were determined by extensive spectroscopic analyses and
minated analog of 3α-hydroperoxy-3-epiaplysin, while compound 40 TDDFT-ECD method. L. nangii was shown as a source of neolaurene (46),
was suggested as an epimer of compound 39 at C-3 position, as well a new 3,4-olefinic laurene type carbon skeleton isolated from a marine
debromo analog of 3β-hydroperoxyaplysin [75]. Undescribed bromi­ source. Its stereochemistry was determined based on NOESY data [79].
nated nonaromatic isolaurene type sesquiterpene identified as 2-bromo- One new laurane-type sesquiterpene, 10-hydroxyepiaplysinol (47) and
3,5,6-trihydro-1,4-dihydroxy isolaurene (41) was found in L. heteroclada known epiaplysinol [80] were isolated from L. tristicha [81].
[76]. Recently, the structure of teanol (42), undescribed rearranged
brominated cyclolaurane-type sesquiterpene, was reported from 2.4. Aristolane and brasilane skeleton sesquiterpenes
Thailand red alga L. mariannensis (the first report of the genus Laurencia
from Thailand). The relative stereochemistry on the oxolane ring of 42 The aristolane-type sequiterpenes are derivatives of 6,11-

5
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

Fig. 2. Bisabolane and cuparane skeletons sesquiterpenes from Laurencia.

Fig. 3. Laurane skeleton sesquiterpenes from Laurencia.

cycloeremophilanes. For the first time, debilone (48) (Fig. 4) was iso­ 2.5. Eudesmane skeleton sesquiterpenes
lated from L. complanata that represents the first marine organism as a
source of this compound [82]. Brasilane-type sesquiterpenes possess the L. obtusa was the source of eudesmane-type sesquiterpenes that were
basic structure of octahydro-1,6,6-trimethyl-4-(1-methyl)-1H-indene reported for the first time and they included undescribed eudesmane
[53]. Esselin et al. [54] found one undescribed compound 4-hydroxy-5- sesquiterpene named eudesma-4(15),7-diene-5,11-diol (50), along with
brasilene (49) (Fig. 4), along with known brasilenol and epibrasilenol trinor-sesquiterpene teuhetenone (51) and seco-eudesmane sesquiter­
[83] in L. obtusa. pene, chabrolidione B (52) (Fig. 5) [84]. Chemical investigation of the
red alga L. pinnata collected from Nanji Island of China [85] resulted in
the isolation and identification of three new eudesmane sesquiterpenes,
1β-bromoselin-11-en-4α-ol (53), 1β-bromo-4α,5α-epoxyselinane (54),
and 1β-bromoselin-3,11-diene (55), and one new 6,8-cycloeudesmane
sesquiterpene, 1β-bromo-6,8-cycloselin-4(15)-ene (56) (Fig. 5)
together with one known eudesmane sesquiterpene, 1β-bromoselin-4
(15),11-diene [86]. Eudesmane derivatives reported in that study are
predominant and rarely reported from this genus, and they may be taken
as chemotaxonomic markers for L. pinnata. One new eudesmane-type
sesquiterpene was isolated from L. obtusa [58] and it was named
eudesma-4(15),11-diene-5,7-diol (57).

Fig. 4. Aristolane and brasilane skeletons sesquiterpenes from Laurencia.

6
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

Fig. 5. Eudesmane skeleton sesquiterpenes from Laurencia.

2.6. Snyderane skeleton sesquiterpenes aplysistatin, palisadin B, 3,4-epoxypalisadin A, 2-hydroxyluzofuranone


and 2-hydroxyluzofuranone B were for the first time isolated from L.
Snyderane-type sesquiterpenes are bromo monocyclo-nerodiol de­ intermedia [93] even though they were already reported for some other
rivatives. In 2017, Esselin et al. [54] analyzed the crude extract obtained species such as L. similis [94], L. saitoi [95] and L. luzonensis [96].
from L. obtusa and discovered that its chemical composition is domi­
nated by β-snyderol [87] that was also observed in the study of Sutour 2.7. Other skeletons of sesquiterpenes
et al. [88]. Hence, Esselin et al. [89] detected an unusual new γ-sny­
derane skeleton in 8-keto-10-dehydrobromo-γ-snyderol (58) (Fig. 5) In addition to all mentioned skeletons, certain found sesquiterpenes
that was isolated from L. obtusa. Six known compounds including cannot be easily classified into the mentioned groups (Fig. 6, Table 1). A
α-snyderol and β-snyderol [87], two rearranged derivatives of α-sny­ novel rearranged carbon skeleton, johnstane, was found in L. johnstonii
derol, 8-keto-10-dehydrobromo-β-snyderol and 8-hydroxy-β-snyderol and represents the first natural sesquiterpene metabolite with that kind
[90] were also determined. While α- and β-snyderanes are characteristic of structure isolated from Laurencia species so far. It was identified as 3α-
for Laurencia genus [5], γ-snyderanes are rare. The authors suggested bromojohnstone (59) [55]. Aromadendrane skeleton was discovered in
that γ-snyderol can be considered as a chemotaxonomic marker in L. the compound identified as (+)-cyclocolorenone (60) isolated from L.
obtuse, but a systematic study of other L. obtusa populations is necessary intricata [91]. Omphalaurediol (61) represents the second example of
[89]. L. snackeyi collected from three different locations (Minna Island, marine-derived omphalane sesquiterpenes, while rhodolaurenones B
Sesoko Island and Kayo Coast) revealed the presence of palisadin A, (62) and C (63) were found as new rhodolaurane-type sesquiterpenes
palisadin B and aplysistatin [91] that were already reported in the [32], along with known rhodolaurenone A [97], rhodolaureol and iso­
literature [92]. Five halogenated snyderane-type sesquiterpenes, rhodolaureol [45]. da Silva Machado et al. [98] examined six native

Fig. 6. Other skeletons sesquiterpenes from Laurencia.

7
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

populations of L. dendroidea at Southeast Brazilian coast in order to enhancement of the cytotoxicity against various cancer-derived cell
evaluate the chemical diversity and they reported that a triquinane lines (colon, glioblastoma, breast, ovarian, lung, skin, prostate, neuro­
sesquiterpene, named silphiperfolan-7β-ol (64), was observed in all blastoma, pancreas, and murine glioblastoma) [8,100,101]. Barcellos
investigated populations. Two new non-halogenated sesquiterpenes, Marini et al. [6] showed that (− )-elatol and obtusol exhibited the most
snakeol (65) and snakediol (66), were isolated from L. snackeyi. The promising activity against Colo-205 cell line with IC50 values of 1.2 ±
structure of snakeol (65) is elucidated as the unusual monocyclofarnesol 1.4 μg/mL and 2.5 ± 1.3 μg/mL. Due to the multidrug resistance in
due to the E configuration of double bond at C-2/C-3, while the most tumors after successful chemotherapy, new antitumor substances should
Laurencia species possess the Z configuration double bond at C-3/C-4 be found with high tumor cell selectivity resulting in apoptosis. Hence,
[99]. elatol and obtusol showed a remarkable apoptotic process in the treat­
ment against Colo-205 cell line, but authors suggested that other
3. Bioactivity of sesquiterpenes isolated from Laurencia species methodologies should be used for the confirmation of this process [6].
Not only the structure of sesquiterpenes influences the cytotoxic
Sesquiterpenes from Laurencia species showed different activities: activity intensity, but also the applied concentration against cancer cell
cytotoxic, anti-inflammatory, antibacterial, antifungal, antidiabetic, lines. It was observed that laurinterol showed dose-dependent inhibition
anthelmintic activities and others as well as different ecological roles. of the metabolic activity of breast cancer cell lines. When crude extract
The most actual is cytotoxic activity (expressed as IC50 values) that can of L. johnstonii was applied against the same cells, the reduced metabolic
be even compared to the standard compounds such as paclitaxel that has viability was also observed in breast cancer explants. The cytological
been used in clinical trials for chemotherapy. However, only a few effects of the crude extract (IC50 value for Vero cells was 26.18 μg/mL
studies are available containing comparison and potency of isolated and 28.05 μg/mL for MCF-7 cells) were similar to those induced by
sesquiterpenes with reference drugs. Targeted sesquiterpenes could be paclitaxel which served as positive control, while activity of laurinterol
effective anti-inflammatory agents in comparison to e.g. well-known (IC50 = 15.68 μg/mL for Vero cells; IC50 = 16.07 μg/mL for MCF-7 cells)
anti-inflammatory agent dexamethasone and exhibited moderate to was similar to cisplatin, suggesting different molecular damage mech­
potent antibacterial and antifungal activity compared to e.g. benzylpe­ anisms that should be studied in detail [102]. Eudesma-4(15),7-diene-
nicillin. Further investigations should provide better understanding of 5,11-diol (49), teuhetenone (50) and chabrolidione B (51) were tested
how chemical structure of the compound influences the bioactivity against MCF-7 cells and the results were compared to cisplatin. Teuhe­
mechanism. However, many studies have been reporting a variety of tenone (50) exhibited the highest cytotoxic activity against MCF-7 (IC50
sesquiterpenes from Laurencia species, but without performing bioac­ = 22.8 ± 0.18 μM) when compared to cisplatin (59 ± 0.045 μM) [84].
tivity tests. Furthermore, the bioactivity of Laurecia species has been Neolaurene (45) showed the strongest cytotoxic activity among all
often presented for different total extracts as the compounds mixture, tested compounds against HeLa, MCF-7 and P-388 cells with MIC values
not providing information about specific compound activity and there­ of 125.0, 175.0 and 125.0 μg/mL, respectively [79]. 4,10-Dibromo-3-
fore this gap is opening new research potential, especially for the most chloro-7(14)-chamigrene, α-isobromocuparene and cuparen-3-ol were
represented sesquiterpenes. In addition, sesquiterpenes are extremely isolated from L. majuscula and tested against three cancer cell lines HCT-
important for the chemical defences of macroalgae. However, there is 116 (human colon cancer cells), PC-3 (human prostate cancer cells) and
lack of bioassays in an ecologically relevant manner for instance field HepG2 (human hepatocellular carcinoma cells). α-Isobromocuparene
and laboratory assays with natural herbivores that are necessary to showed the best activity against HCT-116 cell lines (IC50 value of 46.70
examine the ecological interactions. Their exact role in marine μg/mL), cuparen-3-ol against PC-3 cells (IC50 value of 60 μg/mL), while
ecosystem and the mechanisms of macroalgal chemical defences against 4,10-dibromo-3-chloro-7(14)-chamigrene was best active against
herbivores and competitors should be studied more, since very little data HepG2 cells (IC50 52 μg/mL), but when compared with cisplatin which
are available. served as positive control they did not exhibit significant activities (IC50
values for cisplatin were 3.78, 1.50 and 1.65 μg/mL, respectively) [57].
3.1. Cytotoxic activity
3.2. Anti-inflammatory activity
Cytotoxic activity was the most studied among Laurencia-derived
sesquiterpenes which resulted with new findings of potential antitumor The studies have shown that marine-derived halogenated specialised
agents. The structure of tested sesquiterpenes exhibited a significant metabolites can be considered as effective anti-inflammatory agents
influence on the promotion of activity against wide variety of tested cell [103,104]. Chen et al. [43] evaluated anti-inflammatory activities of
lines. The composition of chamigrane-type sesquiterpenes revealed that compounds isolated from L. tristicha by measuring their ability to sup­
β-chamigrane type sesquiterpenes (elatol, obtusol, [1(15) press N-formylmethionyl-leucyl-phenylalanine/cytochalasin B (fMLP/
Z,2Z,4R,8S,9R]-8,15-dibromochamigra-1(15),2,11(12)-trien-9-ol, [1 CB)-induced superoxide anion (O−2 •) generation, as well as elastase
(15)E,2Z,4R,8S,9R]-8,15-dibromochamigra-1(15),2,11(12)-trien-9-ol, release in human neutrophils. When the concentration of 20 μM was
dendroidiol and cartilagineol) showed much stronger activity against applied, tristichol A 24 and allo-isoobtusol exhibited significant inhibi­
three tested cancer cell lines (HeLa, MCF-7 and P-388; IC50 ≤ 1.0 μg/ tion towards generation of fMLP/CB, while elastase release was signif­
mL), while α-chamigrane sesquiterpenes (laureborneone, lau­ icantly inhibited when tristichol A 24, [1(15)Z,2Z,4S,8R,9S]-8,15-
remantanones A and B) showed activity against only two types of cell dibromochamagra-1(15),2,11(12)-trien-9-ol and allo-isoobtusol were
lines (HeLa and P-388; IC50 ≤ 5.0 μg/mL) [7,10]. Among all tested used. (− )-Elatol (IC50 16.5 ± 1.1 μg/mL) was the most active for the
isolated sesquiterpenes, 7-aldehydelaurencenone B showed the stron­ inhibition of production of nitric oxide among other isolated com­
gest activity against HeLa cell line (IC50 = 7.83 μg/mL) and P-388 cell pounds, obtusol (IC50 33.2 ± 1.0 μg/mL) and cartilagineol (IC50 > 100
line (IC50 = 8.98 μg/mL) that can be attributed to the presence of μg/mL), from L. dendroidea. Also, (− )-elatol reduced tumor necrosis
aldehyde group when compared to laurencenone B (17) that did not factor-α at 100 μg/mL (53.01 ± 2.49%) and 20 μg/mL (35.05 ± 1.92%)
show any activity. Therefore, the authors suggested that the cytotoxicity with an IC50 of 189.8 ± 3.6 μM.
of laureborneone, [1(15)Z,2Z,4R,8S,9R]-8,15-dibromochamigra-1
(15),2,11(12)-trien-9-ol, [1(15)E,2Z,4R,8S,9R]-8,15-dibromochamigra- 3.3. Antimicrobial activity
1(15),2,11(12)-trien-9-ol might be due to the presence of the vinyl
bromide [7]. Kamada et al. [32] tested a wide range of halogenated chamigrane-
It was shown that the presence of hydroxyl and exo-methylene type sesquiterpenes and their rearranged analogues, as well as some
moieties are important for biologically active sesquiterpenes due to the metabolites belonging to omphalane and rhodolaurane skeletons (elatol,

8
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

laurencenone B and 2-chloro-3-hydroxy-α-chamigran-9-one, cartilagi­ activity against M. gypseum with MIC80 value of 4 μg/mL, while com­
neol, (− )-laurencenone D (6), and (+)-deschloroelatol (7), omphalaur­ positacins B–D (9–11), I–K (16–18) and deoxyprepacifenol displayed
ediol (60), rhodolaurenones B (61) and C (62), rhodolaurenone A, moderate antifungal activity [33]. Four bisabolane sesquiterpenes iso­
rhodolaureol and isorhodolaureol). The compounds elatol, cartilagineo, lated from L. composita exhibited significant inhibitory activity against
rhodolaurenones B 61 and C 62, showed bactericidal activity against all Microsporum gypseum with MIC values of 4.0, 8.0, 8.0 and 4.0 μg/mL for
three clinical pathogens, Escherichia coli, Salmonella typhi and Vibrio laurecomposins A (29) and B (30), preintricatol and gossonorol,
cholera (MIC value 100 μg/mL; MBC value 250 μg/mL). Neolaurene (45) respectively. The results were compared with positive controls including
isolated from L. nangii showed good antibacterial activity against E. coli, itraconazole, ketoconazole, fluconazole and voriconazole that inhibited
S. typhi, Staphylococcus aureus and V. cholera with LC50 values of 12.5, M. gypseum with MICs values of 2.0, 1.0, 8.0 and 0.125 μg/mL, respec­
7.5, 7.5 and 12.5 μg/mL, respectively [79]. The compounds isolated tively [36].
from L. tristicha in the study by Chen et al. [43] were evaluated against
Enterobacter aerogenes, Serratia marcescens and Yersinia enterocolitica. 3.4. Other reported bioactivities
When compared to the standard antibiotic ampicillin at the dosage of
100 μg/disk (inhibition zone 5 mm), only 9-(Z)-bromomethylidene- As it was stated that the halogenation of sesquiterpenes leads to more
1,5,5-trimethylspiro[5.5]undeca-1,7-dien-3-one showed a good anti­ active compounds in terms of cytotoxicity and anti-inflammatory ac­
bacterial activity against S. marcescens at the dosage of 100 μg/disk tivity, Garcia-Davis et al. [55] conducted bromination of inactive laur­
(inhibition zone 8 mm), while other compounds (tristichol A 24, ma'i­ interol and isolaurinterol to observe the influence on antiparasitic
lione, [1(15)E,2Z,4S,8R,9S]-8,15-dibromochamagra-1(15),2,11(12)- activity against Acanthamoeba castellanii. The bromination resulted in
trien-9-ol, ma'iliohydrin and allo-isoobtusol) displayed lower activities. four halogenated derivatives 3α-bromojohnstone (58), 8-bromoaplysin,
When tristichone A (22) and ristichone B (23) were tested against 8,10-dibromoisoaplysin and 13-hydroxy derivative which exhibited
multridrug-resistant bacteria such as E. coli, Klebsiella pneumoniae, Pro­ better activity than the natural sesquiterpenes with IC50 values of
teus mirabilis, Pseudomonas aeruginosa, Enterococcus faecalis and 18.804 ± 0.198, 24.559 ± 1.105, 22.818 ± 1.896 and 29.937 ± 2.918
S. aureus, only ristichone B (23) exhibited activities with MICs in the μg/mL, but showed lower activity when compared to the reference
range of 0.08–0.15 mM. Hence, tristichone A (22) displayed significant compounds chlorhexidine (IC50 = 1.526 ± 0.45 μg/mL) and vor­
anticandidal activity against Candida albicans and Candida tropicalis iconazole (IC50 = 0.33 ± 0.1 μg/mL) which are used for the treatment of
(MICs = 8.27 and 10.13 μM, respectively) when compared to positive Acanthamoeba infections. Among them, 3α-bromojohnstone (58) was
control amphotericin B (MICs = 4.63 and 5.27 μM, respectively) [58]. the most active compound with IC50 value of 18.804 ± 0.198 μg/mL.
Snakeol (64) and snakediol (65) isolated from L. snackeyi were tested for Arberas-Jimenez et al. [107] aimed to investigate the potential of
their antibacterial activity and it was showed that they exhibited strong laurinterol isolated from L. johnstonii against Negleria floweri that is
activity against S. typhi with MIC/MBC ratio of 2.79, as well against known as brain eating amoebae and it can cause infection called primary
E. coli with MIC/MBC ratio of 3.02 and 2.76, respectively. It was sug­ amoebic meningoencephalitis (PAM). The results showed that laur­
gested that these antibacterial activities could be due to the presence of interol induced cell death in the treated amoebae and it was able to
hydroxyl group at C-1 position in the monocyclofarnesol-type structure eliminate it with the concentration IC50 of 13.42 μM, but it was not as
[99]. Laurecomposins A (29) and B (30), preintricatol and gossonorol active as the commonly used therapeutic agent against N. floweri known
were evaluated for antibacterial activity against Gram-positive bacteria as amphotericin B (IC50 = 0.12 μM).
S. aureus and Gram-negative bacteria P. aeruginosa. Tested compounds Debromo-3α-hydroperoxy-3-epiaplysin (38) and debromo-3β-
were not active towards Gram-negative P. aeruginosa, but they were hydroperoxyaplysin (39) were evaluated for the protein tyrosine phos­
active against Gram-positive S. aureus with MICs ranging from 10.9 to phatase 1B (PTP1B) enzyme inhibitory activity. The compound 39
26.8 μg/mL. The results were compared with MICs obtained for positive exhibited significant PTP1B inhibitory activity with IC50 value of 13.0
controls including tobramycin, kanamycin sulfate, chloramphenicol and μg/mL with oleanolic acid as the positive control (IC50 = 1.3 μg/mL),
erythromycin (2.0, 16.0, 8.0 and 0.5 μg/mL, respectively) [36]. The while its epimer 38 was inactive. Authors suggested that the R-config­
tested halogenated sesquiterpenes, elatol and obtusol, isolated from L. uration at C-3 position may be significant for the activity [75]. Later, Li
dendroidea were tested for antimycobacterial activity of Mycobacterium et al. [78] also evaluated dimeric bis-lauranes laurokamurols A − C
bovis and M. tuberculosis H37 Rv culture. Obtusol displayed higher ac­ (42–44) for their PTP1B activity and they exhibited significant inhibi­
tivity with IC50 of 31.4 ± 0.8 μg/mL for M. bovis and IC50 of 97.1 ± 1.4 tory activity with IC50 values of 8.1, 12.5 and 6.1 μM, respectively,
μM, whereas elatol was not able to inhibit mycobacterial growth [105]. compared with oleanolic acid (IC50 = 3.3 μM). The potent PTP1B
Laurinterol and aplysin isolated from L. johnstonii were evaluated inhibitory activities of these compounds could provide a direction for
against nine strains of M. tuberculosis that causes tuberculosis. Laur­ further function-oriented synthesis of such compounds towards new
interol inhibited eight out of nine strains of M. tuberculosis with MICs antiobesic or antidiabetic drug discovery.
≤100 μg/mL, but when compared with rifampicin, the most common Three tested compounds, 8-deoxyalgoane (32), 1-deacetoxyalgoane
drug for tuberculosis, MICs were much lower than for laureinterol and (33) and algoane (34) isolated from L. natalensis showed greater anti-
aplysin (0.1–32 μg/mL). The strain M. tuberculosis CIPTIR-F296 was the inflammatory activity against xanthine oxidase than the commercial
most susceptible for both compounds and when compared with rifam­ inhibitor allopurinol, especially algoane (34) with IC50 value of 3.091
picin (MIC = 32 μg/mL), MIC value for laurinterol was the lowest (MIC μM when compared with IC50 value of 3.543 μM for allopurinol. 1-
= 25 μg/mL). Authors suggested that for this activity the hydroxyl group Deacetoxyalgoane (33) and algoane (34) possess two hydroxyl groups
in laurinterol was more relevant than the bromide atom and the for­ and they were more active for the inhibition of α-glucosidase (AG) and
mation of an ether bond between the aromatic ring and the cyclopentane dipeptidyl peptidase IV (DPP IV) with IC50 values of 36.167 ± 2.48 μM
fragment in aplysin [106]. and 79.023 ± 3.84 μM for the inhibition of AG and 43.365 ± 3.27 μM
Eudesma-4(15),7-diene-5,11-diol (49) and chabrolidione B (51) and 23.941 ± 1.70 μM for the inhibition of DPP IV. 8-Deoxyalgoane (32)
exhibited a good antifungal activity against C. albicans, C. tropical, contains only one hydroxyl group and it was the least active compound
Aspergillus flavus and A. niger with MIC values ranging from 2.10 μM to (IC50 values of 82.843 ± 1.98 μM for the inhibition of AG and 58.523 ±
6.7 μM, while positive control amphotericin B showed MIC 4.6–5.4 μM 3.32 μM for the inhibition of DPP IV). Hence, all three compounds
[84]. Compositacins A-K (8–18) were evaluated against eight microor­ exhibited higher activity than acarbose (IC50 value of 306.84 ± 2.15 μM
ganisms including Microsporum gypseum, Trichophyton rubrum, for the inhibition of AG), known as commercial α-glucosidase inhibitor.
C. albicans, C. glabrata, C. parapsilosis, Cryptococcus neoformans and Also, algoane (34) was noted as the most promising antidiabetic agent
A. fumigatus. Only compositacin G (14) exhibited significant antifungal due to its activity close to the commercial inhibitor of

9
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

dipeptidylpeptidase IV (IC50 value of 2.923 ± 0.18 μM for the inhibition A. brasiliana induced the anti-herbivory chemical defences in L. den­
of DPP IV). These findings were confirmed when all three compounds droidea and showed significant changes in the sesquiterpene metabo­
were subjected to molecular property analysis for the prediction of po­ limic profile by increasing its content. Hence, they showed that L.
tency and bioavailability. All three molecules displayed good results for dendroidea can perceive waterborne chemical signals released by
LogP and Topological Surface Area (TPSA), known as two main in­ neighboring grazed conspecifics [118].
dications for the prediction of oral bioavailability of drugs, also the L. intricata can be promising source of bioactive metabolites that can
positive results against the enzyme inhibitory targets and nuclear re­ be used as potential biocontrol agents for stored-product insects. It was
ceptor ligands were obtained [9]. Although stereochemistry of the established that (+)-cyclocolorenone (59) can act as natural insecticide
compounds did not show any significant effect on cytotoxic activity due to its strong repellent activity against the maize weevil Sitophilus
[108], it was observed that the enantiomers of elatol exhibited differ­ zeamais (EC50 2.0 μg/mL), and it showed the same activity as pyrethrin
ences in acetylcholinesterase (AChE) inhibitory activity. While (EC501.7 μg/mL) known as natural insecticide [91]. Laurinterol isolated
(+)-elatol did not exert significant effect on the enzyme activity, from L. nidifica also showed a strong repellency against the maize weevil
(− )-elatol were the most active tested sesquiterpene with 78.5% of AChE with ED50 value of 12.65 μg/cm2 and it was showed as good insecticide
inhibition. Other compounds that exhibited significant activity were against termite Reticulitermes speratus (LD50 2.2 μg/insect) even though
dendroidiol with inhibition of 72.9% and cartilagineol with 61.3% of it exhibited lower activity than the commercial insecticides (rotenone,
inhibition [109]. Authors also presented a molecular docking of fenitrothion and pyrethrins with LD50 values of 0.110, 0.025 and 0.013
(− )-elatol that showed the importance of interactions with specific μg/insect, respectively) [74]. These are promising results that can help
residues, thus corroborating the inhibitory effect of AChE. In the study for the development of environmentally-friendly natural pesticides
where 11 laurane-type sesquiterpenes obtained from L. tristicha were without toxic residues and disturbances in the environment, also lethal
evaluated for their antioxidant activity by scavenging effect on hydroxyl effects on non-targeted organisms will be avoided but further research
free radical using electron paramagnetic resonance (EPR), it was shown and evaluation of structure-activity relationship are needed. Anti­
that the 1,10-epoxy moiety was crucial for the activity since the com­ lesihmanial activity of (− )-elatol and obtusol was tested against the
pounds having this moiety exhibited better activity. The compounds insect-stage promastigotes of Leishmania mazonensis and they exhibited
that possess Br at phenyl ring, as well as the compounds with acetyl in vitro and in vivo leishmanicidal activity with IC50 values of 6.2 μg/mL
group connecting to 1-OH showed less activity, but the phenolic hy­ and 9.7 μg/mL, respectively. The values were lower than the reference
droxy group significantly enhanced the activity [81]. drug potassium antimony (III) tartrate hydrate (IC50 0.9 ± 0.2 μg/mL)
[23].
3.5. Ecological role of sesquiterpenes
4. Biosynthesis of sesquiterpenes
Even though the literature has been reporting mostly bioactivities of
sesquiterpenes, several authors reviewed and investigated the ecological Sesquiterpenes represent a large group of natural organic compounds
role of sesquiterpenes [110,111,112]. Namely, macroalgae have been whose identification is still expanding due to broad structural diversity
producing sesquiterpenes as the protection against enemies or for (ca. 300 specific C15-hydrocarbon skeletons) that represent a challenge
feeding and reproduction. They have developed a range of defensive for the structure elucidation and synthesis [19]. For instance, only
mechanisms, both physical and chemical, as an answer to the extreme among tricyclic sesquiterpenes, 54 different skeletal types (284 com­
conditions and competition for space, light and nutrients [113]. How­ pounds) were reported by Le Bideau et al. [14] pointing out large che­
ever, the predictions of toxic or deterrent effects are difficult to make modiversity. Several of these compounds were specific only for the
based only upon the chemical structures of particular metabolites. The marine organisms and they could not be found in the terrestrial plants.
effects can vary greatly among the compounds that differ only slightly in In distinction from monoterpenes which were mostly halogenated
the chemical structure as it was noticed that halogenation in red algal [119], tricyclic sesquiterpenes showed limited halogenation, but the
metabolites did not enhance their feeding deterrent effects. Debro­ most of halogenated molecules were found in the genus Laurencia.
moaplysinal, known as one of common laurane-type sesquiterpene, Ružička and Stoll [120] first pointed out that the sesquiterpenes can
exhibited moderate antifouling activity (EC50 = 1.00 μg/mL), but be formed from a common intermediate farnesyl pyrophosphate (FPP),
pacifenol (EC50 = 2.36 μg/mL) and 2,10-dibromo-3-chloro-9-hydroxy- and previously it was considered that the intermediate was an activated
α-chamigrene (EC50 = 2.51 μg/mL) exhibited weak activity. These re­ derivative of acyclic alcohol farnesol. While geranyl pyrophosphate
sults were compared to the positive control CuSO4 with EC50 = 0.18 μg/ (GPP) is synthesized in the plastids, FPP is synthesized in the plant
mL [114]. Also, Al-Lihaibi et al. [115] compared their results for anti­ cytosol from acetyl-CoA [121], but this observation has not been
fouling activity with CuSO4 (EC50 = 0.34 μg/mL) and the results indi­ experimentally determined in red algae where some of methylerythritol
cated that 2,10-dibromo-3-chloro-7-chamigrane isolated from L. obtusa phosphate (MEP) pathway enzymes were plastidial, but others appeared
showed potent activity against larvae of Balanus amphitrite with EC50 = to be localized in the cytosol or other organelles [122]. de Oliviera et al.
0.14 μg/mL. Protopapa et al. [116] evaluated several sesquiterpenes [123] reported the presence of MEP pathway in Laurencia dendroidea,
isolated from different species of Laurencia genus and among them only but there are no studies reporting the presence of mevalonate (MVA)
perforenol exhibited promising antifouling potential (IC50 value 0.5 μM) pathway in this specie. In further research, de Oliviera et al. [124]
with bromosphaerol serving as a benchmark (IC50 value 0.6 μM) against indicated high diversity of chemical modifications of sesquiterpene
the barnacle Amphibalanus amphitrite and at the same time it did not precursors resulting with a wide variety of carbon skeletons. However,
exhibit ecotoxicity against non-target organisms. Liposoluble terpenes the biogenetic routes can be different when vascular plants and algae are
can mediate different alga-herbivore interactions as it was shown in the compared due to their phylogenetic distance.
study by Nocchi et al. [117] where elatol (9) isolated from L. dendroidea
attracted Aplysia brasiliana and revealed its potential as a chemical 4.1. Sesquiterpenes biosynthesis stages
mediator in the interaction between Laurencia and Aplysia. They showed
for the first time with the combination of experiments that elato can The enzymes known as prenyltransferases are responsible for the
serve as chemical cue attracting A. brasiliana which is very important for elongation of hydrocarbon chains by adding isopentenyl pyrophosphate
the maintenance of associations and biodiversity. Another study exam­ (IPP) units to dimethylallyl pyrophosphate (DMAPP). IPP units can be
ined the chemical defense in L. dendroidea by direct grazing by formed by MVA and/or MEP pathways. After the initial condensation of
A. brasiliana, as well as water-borne chemical cues from conspecific IPP and DMAPP, GPP is formed, while another 1′ -4 condensation of IPP
neighbors grazed by A. brasiliana. The results revealed that grazing by yields FPP, crucial for sesquiterpene biosynthesis [125].

10
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

The conversion into the cyclic terpene carbon skeletons is stimulated one of the carbon-carbon double bonds in the substrate [125]. Specif­
by terpene synthases, also known as terpene cyclases and the pyro­ ically, for sesquiterpenes, FPP is converted to nerolidyl pyrophosphate
phosphate group is eliminated and the carbocations are generated. (NPP). This initial isomerization is crucial for generating the interme­
Hence, formed carbocations can undergo the cyclization, but first the diate with the proper reactivity and conformational flexibility to over­
isomerization of the initial carbocation to an intermediate capable of come the stereochemical barrier for direct cyclization and allow the
cyclization is needed due to the impossibility of direct cyclization of (E) formation of cyclic products. During the isomerization, ionization of FPP
configuration of double bond of terpene synthases substrates. The to the transoid cation-pyrophosphate anion occurs that is subjected to
cyclization can occur by the addition of the stabilized cationic centre to several transformations [126]. As can be seen in Fig. 7, the initial

Fig. 7. Proposed mechanisms for the initial cyclizations of FPP and formation of typical Laurencia sesquiterpene skeletons.

11
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

cyclization of (E,E)-FPP (66) to sesquiterpenes can occur by six alter­ binding sites in all sesquiterpene synthases. The farnesyl chain in the
native pathways. Without the initial isomerization of FPP, two ways of active site accepts reactive conformation by binding three Mg ions. The
ionization of farnesyl cation (67) are possible, either 1,10-cyclization to second phase includes the reactions of sesquiterpene synthases leading
(E,E)-germacradienyl cation (68) or 1,11-ring closure to (E,E)-humulyl the carbocation through complex cyclization cascades that include
cation (69). As mentioned before, the isomerization of 66 to both en­ additional carbocationic and neutral intermediates. After the initial
antiomers (3S)- or (3R)-NPP (70), depending on the specific sesquiter­ ionization, the enzymes act as passive catalysts that simply guide the
pene synthase, via 71 is obligatory for common 1,6-cyclization to intermediates through the reaction cascades [19].
bisabolyl cation 72 or a less common cycloheptenyl cation 73 by 1,7- There are specific sesquiterpene synthases responsible for each
cyclization that requires (Z)-configured olefinic double bond. The in­ sesquiterpene class. This great variety is well represented in the work by
termediate cation (71) created by diphosphate extrusion from (70) can Cane [126] where he reviewed the enzymatic formation of sesquiter­
serve as the precursor for 72 and 73, (E,Z)-germacradienyl cation (74) penes including the description of wide variety of sesquiterpenes cy­
obtained by 1,10-cyclization and (E,Z)-humulyl cation (75) formed by clases, such as bisabolane, trichodiene, bergamotene, humulene,
1,11-ring closure that further generates the stereoisomers of 68 and 69 caryophyllene, pentalenene, aristolochene, and patchoulol synthases.
[127]. Sesquiterpenes are present with wide range of skeleton types, but Also, the mechanisms of formation of cyclic sesquiterpenes and their
chamigrane, bisabolane, laurane and cuparane are the most represented precursors along with their stereochemistry were represented. Sesqui­
among macroalgae belonging to Laurencia genus. Thus, their probable terpene cyclases are among the most naturally occurring versatile cat­
formation is presented in Fig. 7 from bisabolyl cation which is the main alysts responsible for the formation of more than 15 distinct
precursor of these skeleton types. Bisabolyl cation 7 leads to the for­ sesquiterpene carbon skeletons in Laurencia species [132]. de Oliviera
mation of chamigrane and bisabolane skeleton types of sesquiterpenes et al. [124] unveiled the genes related to the biosynthesis of α-bisabo­
including the conversion of α-bisabolane precursor via the oxidation and lane synthase which is responsible for the biosynthesis of bisabolane-
nucleophilic attack by bromide resulting in halidrine derivative which type sesquiterpenes. Moreover, the genes coding pentalenene synthase
yields bisabolane sesquiterpenes. On the other hand, the formation of and δ-cadinene synthase which are involved in humulene- and cadinyl-
γ-bisabolane precursor yields α- and β-chamigrane derivatives by the type sesquiterpenes biosynthesis were detected. The biosynthesis of
enzymatic cyclization of γ-bisabolane derivative [128]. Hence, laurane germacrene-type sesquiterpenes involves the enzymes germacrene-A
type sesquiterpenes are also derived from bisabolyl cation intermediate synthase and oxidase, aristolochene synthase [126] as well as 5-epi-aris­
followed by the ring closure, alkyl shift and then aromatization of the tolochene-1,3-dihydroxylase which use germacrene A as intermediate as
cyclohexane ring [65]. Cuparane type sesquiterpenes are derived via demonstrated by Rising et al. [133]. Also, Scheuer [134] suggested that
protonation of γ-bisabolane followed by the cyclization to cuparane the biosynthesis of elatol, the major specialised metabolite found in
cation which subsequently yields sesquiterpenes with cyclopropane ring Laurencia dendroidea, involved the enzymatic addition of bromochloride
[56,129]. Interestingly, cuparane cation can be the precursor for for­ at bisabolonium ion prior to the cyclization to chamigrane derivative. It
mation of laurane type sesquiterpenes by the methyl migration of has been suggested that red algal terpene synthase genes are evolu­
cuparane cation [130]. tionarily more closely related to microbial terpene synthase genes than
The final stage of sesquiterpene biosythesis includes complex re­ to typical plant terpene synthase genes [135].
actions and transformations including oxidation, alkyl and hydride
shifts, deprotonation, isomerization, conjugation and additional cycli­ 5. Conclusion
zations where the parent carbon skeletons are enzymatically mediated
to form carbocationic intermediates and at the end they are converted to The present review encompasses sesquiterpenes isolated exclusively
numerous distinct terpene metabolites. However, most of these metab­ from Laurencia genus in the period 2015–2020 and they were classified
olites are cyclic so the synthases utilize a very similar carbocationic into the groups by their skeleton structure to chamigrane, cuparane,
reaction mechanism which is employed by prenyltransferases. Depro­ laurane, bisabolane, aristolane, brasilane, eudesmane, snyderane and
tonation of the cation to an olefin or capturing by a nucleophile (H2O) other skeletons including those with high degree of halogenation,
marks termination of the reaction cascade [121,125]. especially bromination. Among them, chemical markers were found. It
was shown that Laurencia sesquiterpenes exhibit cytotoxic, anti-
4.2. The enzymes involved in sesquiterpene biosynthesis inflammatory, antibacterial, antifungal, antidiabetic and anthelmintic
activities and others including different ecological roles. The most actual
Sesquiterpene synthases are highly active enzymes which are is cytotoxic activity that can be even compared to e.g. paclitaxel used in
responsible for generating different structures of sesquiterpenes as it was clinical trials in chemotherapy. Targeted sesquiterpenes could also be
shown with only two sesquiterpene synthases from Abies grandis, effective anti-inflammatory agents. The explored structure-activity re­
δ-selinene and γ-humulene synthases, capable of synthesizing up to 52 lationships were presented, but further research in necessary. There is
different sesquiterpenes [131]. The formation of multiple products by lack of bioassays in an ecologically relevant manner for instance field
terpene synthases is the consequence of their reaction mechanism, and laboratory assays with natural herbivores that are necessary to
including the emerging of carbocationic intermediate from which many examine the ecological interactions. Possible general biosynthetic
products can be formed. It is necessary to consider their level of activity, pathways of targeted sesquiterpenes were presented, but more detail
as well as their subcellular location [125]. Sesquiterpene synthases need research of their formation is necessary since the biogenetic routes can
to eliminate the pyrophosphate group (PP) of FPP so that the carboca­ be different when vascular plants and algae are compared due to their
tion can be created and then neutralized to relatively simple structure or phylogenetic distance.
it can undergo to various rearrangements resulting in complex di- and Laurencia sesquiterpenes chemical diversity, along with their bio­
tricyclic structures [121]. activities, leads to more attention of scientists for further exploration of
The catalytic mechanism of synthases can be divided into two pha­ these interesting, easy accessible red algae that are distributed around
ses, the metal ion binding and pyrophosphate cleavage and carboca­ the world. Due to the fact that the novel sesquiterpene structures are still
tionic reaction cascades followed by the final reaction products. In the being detected, this might lead to the new discoveries that can be
first phase, divalent metal (Mg2+) initiates the formation of an ion pair directed towards investigation of biological potential (particularly for
between pyrophosphate and farnesyl cation due to the cleavage of C–O anticancer activity) as well as bioavailability of the most promising
bond in FPP with the attachment of three metal ions at the top of the sesquiterpenes.
active site. The binding of these metal cations, that coordinates
diphosphate group of the substrate, are controlled by two metal ion

12
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

CRediT authorship contribution statement [13] S.M. Al-Massarani, Phytochemical and biological properties of sesquiterpene
constituents from the marine red seaweed, Laurencia: a review. Nat. Prod. Chem.
Res. 2 (2014) 147, https://doi.org/10.4172/2329-6836.1000147.
Ana-Marija Cikoš: Methodology, Visualization, Writing-Original [14] F. Le Bideau, M. Kousara, L. Chen, L. Wei, F. Dumas, Tricyclic sesquiterpenes from
draft preparation. Mladenka Jurin: Writing-Original draft prepara­ marine origin, Chem. Rev. 117 (2017) 6110–6159, https://doi.org/10.1021/acs.
tion, Visualization. Rozelindra Čož-Rakovac: Project administration; chemrev.6b00502.
[15] A. Jesus, M. Correia da Silva, C. Agonso, M. Pinto, H. Cidade, Isolation and
Resources. Dajana Gašo-Sokač: Supervision. Stela Jokić: Conceptual­ potential biological applications of haloaryl secondary metabolites from
ization, Supervision. Igor Jerković: Conceptualization, Supervision, macroalgae, Mar. Drugs 17 (2019) 73, https://doi.org/10.3390/md17020073.
Writing-Original draft preparation. [16] A. Modzelewska, S. Sur, S.K. Kumar, S.R. Khan, Sesquiterpenes: natural products
that decrease cancer growth, Curr. Med. Chem. Anticancer Agents 5 (2005)
477–499, https://doi.org/10.2174/1568011054866973.
Funding [17] G.A. Cordell, Biosynthesis of Sesquiterpenes, Chem. Rev. 76 (1976) 426–460,
https://doi.org/10.1021/cr60302a002.
[18] B.M. Fraga, Natural sesquiterpenoids, Nat. Prod. Rep. 25 (2008) 1180–1209,
This work was supported by the Croatian Government and the Eu­ https://doi.org/10.1039/B806216C.
ropean Union (European Regional Development Fund—the Competi­ [19] D.J. Miller, R.K. Allemann, Sesquiterpene synthases: passive catalysts or active
tiveness and Cohesion Operational Programme - KK.01.1.1.01) through players? Nat. Prod. Rep. 29 (2012) 60–71, https://doi.org/10.1039/c1np00060h.
[20] A.R. Diaz-Marrero, I. Brito, J. de la Roda, J. Daia, M. Cueto, Gomerones A–C,
project Bioprospecting of the Adriatic Sea (KK.01.1.1.01.0002) granted halogenated sesquiterpenoids with a novel carbon skeleton from, Laurencia
to The Scientific Centre of Excellence for Marine Bioprospecting- majuscula. Tetrahedron 64 (2008) 10821–10824, https://doi.org/10.1016/j.
BioProCro. Therefore please add as the acknowledgement: "We would tet.2008.09.027.
[21] E. Dorta, A.R. Diaz-Marrero, M. Cueto, L. D'Croz, J.L. Mate, J. Darias,
like to thank the Croatian Government and the European Union (Euro­
Chamigrenelactone, a polyoxygenated sesquiterpene with a novel structural type
pean Regional Development Fund—the Competitiveness and Cohesion and devoid of halogen from, Laurencia obtusa. Tetrahedron Lett. 45 (2004)
Operational Programme - KK.01.1.1.01) through project Bioprospecting 7065–7068, https://doi.org/10.1016/j.tetlet.2004.07.125.
of the Adriatic Sea (KK.01.1.1.01.0002) granted to The Scientific Centre [22] A. Diaz Marrero, I. Brito, J. de la Rosa, L. D’Croz, O. Fabelo, C. Ruiz-Perez,
J. Darias, M. Cueto, Novel lactone chamigrene-derived metabolites from,
of Excellence for Marine Bioprospecting-BioProCro for funding of this Laurencia majuscula. Eur. J. Org. Chem. 9 (2009) 1407–1411, https://doi.org/
research. 10.1002/ejoc.200801006.
[23] F.L. Da Silva Machado, W. Pacienza-Lima, B. Rossi-Bergmann, L.M. de Souza
Gestinari, M.T. Fujii, J.C. de Paula, S. Soares Costa, N. Peporine Lopes, C.
Author contributions R. Kaiser, A. Ribeiro Soares, Antileishmanial sesquiterpenes from the brazilian
red alga, Laurencia dendroidea. Planta Med. 77 (2011) 733–735, https://doi.org/
All authors contributed to the conceptualization and design of the 10.1055/s-0030-1250526.
[24] X.-D. Li, W. Ding, F.-P. Miao, N.-Y. Ji, Halogenated chamigrane sesquiterpenes
review, as well as drafting and revising it. from, Laurencia okamurae. Mag. Recon. Chem. 50 (2012) 174–177, https://doi.
org/10.1002/mrc.2870.
Declaration of competing interest [25] D.E. White, I.C. Stewart, R.H. Grubbs, B.M. Stoltz, The catalytic asymmetric total
synthesis of elatol, J. Am. Chem. Soc. 130 (2008) 810–811, https://doi.org/
10.1021/ja710294k.
The authors declare no conflict of interest. [26] D.J. Faulkner, Marine natural products, Nat. Prod. Rep. 5 (1988) 613–663.
[27] C.S. Vairappan, Potent antibacterial activity of halogenated metabolites from
Malaysian red algae, Laurencia majuscula (Rhodomelaceae, Ceramiales), Biomol.
References
Eng. 20 (2003) 255–259, https://doi.org/10.1016/S1389-0344(03)00067-4.
[28] C.S. Vairappan, S.P. Anangdan, K.L. Tan, M. Suzuki, Role of secondary
[1] Guiry MD, Guiry GM (2020) AlgaeBase. World-wide Electronic Publication, metabolites as defense chemicals against ice-ice disease bacteria in biofouler at
National University of Ireland, Galway. http://www.algaebase. carrageenophyte farms, J. Appl. Phycol. 22 (2010) 305–311, https://doi.org/
org/search/genus/detail/?genus_id=32789. (accessed 20 Apr 2020). 10.1007/s10811-009-9460-7.
[2] M. Suzuki, C.S. Vairappan, Halogenated secondary metabolites from Japanese [29] M.R. Brennan, K.L. Erickson, D.A. Minott, K.O. Pascoe, Chamigrane metabolites
species of the red algal genus Laurencia (Rhodomelaceae, Ceramiales), Curr. Top. from a Jamaican variety of, Laurencia obtusa. Phytochem. 26 (1987) 1053–1057,
Phytochem. 7 (2005) 1–34. https://doi.org/10.1016/S0031-9422(00)82349-3.
[3] B.G. Wang, J.B. Gloer, N.Y. Ji, J.C. Zhao, Halogenated organic molecules of [30] F.L. da Silva Machado, T.L. Ventura, L.M. Gestinari, V. Cassano, J.A. Resende, C.
Rhodomelaceae origin: chemistry and biology, Chem. Rev. 113 (2013) R. Kaiser, E.B. Lasunskaia, M.F. Muzitano, A.R. Soares, Sesquiterpenes from the
3632–3685, https://doi.org/10.1021/cr9002215. Brazilian red alga, Laurencia dendroidea J. Agardh. Molecules 19 (2014)
[4] C.D. Amsler, Algal Chemical Ecology, Springer-Verlag, Berlin Heidelberg, 2008. 3181–3192, https://doi.org/10.3390/molecules19033181.
[5] M. Harizani, E. Ioannou, V. Roussis, The Laurencia paradox: an endless source of [31] M.E.Y. Francisco, M.M. Turnbull, K.L. Erickson, Cartilagineol, the fourth lineage
chemodiversity, Prog. Chem. Org. Nat. Prod. 102 (2016) 91–252, https://doi. of Laurencia-derived polyhalogenated chamigrene, Tetrahedron Lett. 39 (1998)
org/10.1007/978-3-319-33172-0_2. 5289–5292, https://doi.org/10.1016/S0040-4039(98)01011-9.
[6] M. Barcellos Marini, W. Rodrigues de Freitas, F.L. da Silva Machado, I.C. Ramof [32] T. Kamada, C.-S. Phan, C.S. Vairappan, New anti-bacterial halogenated tricyclic
Leal, A. Ribeiro Soares, M. Masahiko Kanashiro, M. Frazao Muzitano, Cytotoxic sesquiterpenes from Bornean Laurencia majuscula (Harvey) Lucas, Nat. Prod. Res.
activity of halogenated sesquiterpenes from, Laurencia dendroidea. Phytother. 33 (2019) 464–471, https://doi.org/10.1080/14786419.2017.1396593.
Res. 32 (2018) 1119–1125, https://doi.org/10.1002/ptr.6052. [33] X.-Q. Yu, C.-S. Jiang, Y. Zhang, P. Sun, T. Kurtan, A. Mandi, X.-L. Li, L.-G. Yao, A.-
[7] T. Kamada, C.-S. Phan, V.S.-T. Sien, C.S. Vairappan, Halogenated chamigrane H. Liu, B. Wang, Y.-W. Guo, S.-C. Mao, Compositacins A-K: bioactive chamigrane-
sesquiterpenes from Bornean, Laurencia majuscula. J. Appl. Phycol. 30 (2018) type halosesquiterpenoids from the red alga, Laurencia composita Yamada.
3373–3378, https://doi.org/10.1007/s10811-018-1452-z. Phytochem. 136 (2017) 81–93, https://doi.org/10.1016/j.
[8] M. Kladi, H. Xenaki, C. Vagias, P. Papazafiri, V. Roussis, New cytotoxic phytochem.2017.01.007.
sesquiterpenes from the red algae, Laurencia obtusa and Laurencia microcladia. [34] J.J. Sims, W. Fenical, R.M. Wing, P. Radlick, Marine natural products III.
Tetrahedron 62 (2006) 182–189, https://doi.org/10.1016/j.tet.2005.09.113. Johnstonol, an unusual halogenated epoxide from the red alga, Laurencia
[9] K.R.R. Rengasamy, L. Poštova Slavetinska, M.G. Kulkarni, W.A. Stirk, J. Van johnstonii. Tetrahedron Lett. (1972), 195e198, https://doi.org/10.1016/S0040-
Staden, Cuparane sesquiterpenes from Laurencia natalensis Kylin as inhibitors of 4039(01)84278-7.
alpha-glucosidase, dipeptidyl peptidase IV and xanthine oxidase, Algal Res. 25 [35] X.D. Li, F.P. Miao, X.R. Liang, B.G. Wang, N.Y. Ji, Two halosesquiterpenes from,
(2017) 178–183, https://doi.org/10.1016/j.algal.2017.05.008. Laurencia composita, RSC Adv. 3 (2013) 1953–1956, https://doi.org/10.1039/
[10] T. Kamada, C.S. Phan, T. Okino, C.S. Vairappan, Cytotoxicity and antibacterial c2ra23101h.
potential of halogenated chamigrenes from malaysian red alga, Laurencia [36] Hu, Z.-B., Yu, X.-Q., Wang, B., Liu, A.-H., Zhao, T.-S., Guo, Y.-W., Huang, H.-L.,
majuscula. Planta Med. Int. Open 6 (2019) 36–40, https://doi.org/10.1055/a- Mao, S.-C. 2020. Structurally diverse halosesquiterpenoids from the red alga
0977-4418. Laurencia composita Yamada. Fitoterapia 104, 104716. DOI:https://doi.org/10.10
[11] G. Topcu, Z. Aydogmus, S. Imre, A.C. Gӧren, J.M. Pezzuto, J.A. Clement, D.G. 16/j.fitote.2020.104716.
I. Kingston, Brominated Sesquiterpenes from the Red Alga, Laurencia obtusa. J. [37] Ji, N.-Y., Li, X.-M., Li, K., Wang, B.-G., 2009. Halogenated sesquiterpenes from the
Nat. Prod. 66 (2003) 1505–1508, https://doi.org/10.1021/np030176p. marine red alga Laurencia saitoi (Rhodomelaceae). Helv. Chim. Acta 92,
[12] D. Davyt, R. Fernandez, L. Suescun, A.W. Mombru, J. Saldana, L. Dominguez, 1873–179. DOI: https://doi.org/10.1002/hlca.200900073.
J. Coll, M.T. Fujii, E. Manta, New sesquiterpene derivatives from the red alga [38] B.M. Howard, W. Fenical, Structures and chemistry of two new halogen-
Laurencia scoparia, Isolation, structure determination, and anthelmintic activity. containing chamigrene derivatives from, Laurencia. Tetrahedron Lett. 16 (1975)
J. Nat. Prod. 64 (2001) 1552–1555, https://doi.org/10.1021/np0102307. 1687–1690, https://doi.org/10.1016/s0040-4039(00)72233-7.
[39] A.G. Gonzalez, J.D. Martin, V.S. Martin, M. Norte, Marine natural products from
the Atlantic zone. Part 22. Carbon-13 NMR application to, Laurencia

13
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

polyhalogenated sesquiterpenes, Tetrahedron Lett. 20 (1979) 2719–2722, from the red alga Laurencia okamurai Yamada, Phytochem. 103 (2014) 162–170,
https://doi.org/10.1016/S0040-4039(01)86397-8. https://doi.org/10.1016/j.phytochem.2014.03.021.
[40] C. Ireland, M.O. Stallard, D.J. Faulkner, J. Finer, J. Clardy, Some chemical [66] N.-Y. Ji, X.-M. Li, K. Li, L.-P. Ding, B.-G. Wang, Laurane-derived sesquiterpenes
constituents of the digestive gland of the sea hare, Aplysia californica. J. Org. from the marine red alga, Laurencia tristicha (Rhodomelaceae). Nat. Prod. Res. 22
Chem. 41 (1976) 2461–2465, https://doi.org/10.1021/jo00876a024. (2008) 715–718, https://doi.org/10.1080/14786410802019309.
[41] D.J. Faulkner, M.O. Stallard, C. Ireland, Prepacifenol epoxide, a halogenated [67] D.C. Harrowven, M.C. Lucas, P.D. Howes, The synthesis of a natural product
sesquiterpene diepoxide, Tetrahedron Lett. 95 (1974) 3571–3574, https://doi. family: from debromoisolaurinterol to the aplysins, Tetrahedron 57 (2001)
org/10.1016/S0040-4039(01)91969-0. 791–804, https://doi.org/10.1016/S0040-4020(00)01055-3.
[42] C.R. Kaiser, L.F. Pitombo, A.C. Pinto, Complete 1H and 13C NMR assignments of [68] J. Zhang, S.L. Lunyong, H.L. Ding, P. Tu, Q. Zhang, Comparative analysis of
chamigrenes from, Aplysia dactilomela. Magn. Reson. Chem. 39 (2001) 147–149, Laurencia tristicha and Laurencia okamurai based on qualitative and quantitative
https://doi.org/10.1002/mrc.800. HPLC determination, Biochem. Syst. Ecol. 93 (2020), 104170, https://doi.org/
[43] J.-Y. Chen, C.Y. Huang, Y.S. Lin, T.L. Hwang, W.L. Wang, S.F. Chiou, J.H. Sheu, 10.1016/j.bse.2020.104170.
Halogenated Sesquiterpenoids from the Red Alga, Laurencia tristicha Collected in [69] N. Wada, R.D. Kersten, T. Iwai, S. Lee, F. Sakurai, T. Kikuchi, D. Fujita, M. Fujita,
Taiwan. J. Nat. Prod. 79 (2016) 2315–2323, https://doi.org/10.1021/acs. J.-K. Weng, Crystalline-sponge-based structural analysis of crude natural product
jnatprod.6b00452. extracts, Angew. Chem. Int. Ed. 57 (2018) 3671–3675, https://doi.org/10.1002/
[44] M. Suzuki, A. Furusaki, N. Hashiba, E. Kurosawa, The structure and absolute anie.201713219.
stereochemistry of a halogenated chamigrene derivative from the red alga, [70] G.M. König, A.D. Wright, New C15-acetogenins and sesquiterpenes from the red
Laurencia Species. Tetrahedron Lett. 20 (1979) 879–882, https://doi.org/ alga Laurencia sp cf, L. gracilis. J. Nat. Prod. 57 (1994) 477–485, https://doi.org/
10.1246/bcsj.61.3371. 10.1021/np50106a006.
[45] J.C. Coll, A.D. Wright, Tropical marine algae. III. New sesquiterpenes from, [71] R. Kazlauskas, P.T. Murphy, R.J. Quinn, R.J. Wells, New laurene derivatives from,
Laurencia majuscula (Rhodophyta, Rhodophyceae, Ceramiales, Rhodomelaceae). Laurencia filiformis. Aust. J. Chem. 29 (1976) 2533–2539, https://doi.org/
Aust. J. Chem. 42,1591–1603. (1989), https://doi.org/10.1071/CH9891591. 10.1071/CH9762533.
[46] E.G. Juagdan, R. Kalidindi, P. Scheuer, Two new chamigrenes from Hawaiian red [72] M. Kladi, C. Vaigas, G. Furnari, D. Moreau, C. Roussakis, V. Roussis, Cytotoxic
alga, Laurencia cartilaginea, Tetrahedron 53 (1997) 521–528, https://doi.org/ cuparene sesquiterpenes from, Laurencia microcladia. Tetrahedron Lett. 46 (2005)
10.1016/S0040-4020(96)01002-2. 5723–5726, https://doi.org/10.1016/j.tetlet.2005.06.076.
[47] M. Suzuki, E. Kurosawa, K. Kurata, Majusculone, a novel norchamigrane-type [73] R.R. Izac, J.J. Sims, Marine natural products. 18. Iodinated sesquiterpenes from
metabolite from the red alga, Laurencia majuscula Harvey. Bull. Chem. Soc. Jpn. the red algal genus, Laurencia. J. Am. Chem. Soc. 101 (1979) 6136–6137,
60 (1987) 3795–3796, https://doi.org/10.1246/bcsj.60.3795. https://doi.org/10.1021/ja00514a054.
[48] A.C. Philippus, G.A. Zatelli, T. Wanke, A. de Gabriela, M. Barros, S.A. Kami, [74] T. Ishii, B.C. Nguyen Takumi Nagamine, S. Tawata, Insecticidal and repellent
C. Lhullier, L. Armstrong, L.P. Sandjo, M. Falkenberg, Molecular networking activities of laurinterol from the Okinawan red alga, Laurencia nidifica. Records
prospection and characterization of terpenoids and C15-acetogenins in Brazilian of Natural Products 11 (2017) 63–68.
seaweed extracts, RSC Adv. 8 (2018) 29654–29661, https://doi.org/10.1039/ [75] X.L. Li, W.F. He, J. Li, L.F. Lan, X.W. Li, Y.W. Guo, New laurane-type
c8ra02802h. sesquiterpenoids from the Chinese red, alga Laurencia okamurai Yamada, J. Asian
[49] Suzuki, M., Segawa, M., Suzuki T., Kurosawa, E. 1983. Structures of halogenated Nat. Prod. Res. 17 (2015) 1146–1152, https://doi.org/10.1080/
chamigrene derivatives, minor constituents from the red alga Laurencia nipponica 10286020.2015.1102135.
Yamada. Bull. Chem. Soc. Jpn., 56, 3824–3826. DOI: https://doi.org/10.1246/bc [76] H.M. Haniffa, H. Ranjith, W. Dharmaratne, M.Y. Mohammad, M.I. Choudhary,
sj.56.382. Allelopathic activity of some Sri Lankan seaweed extracts and the isolation of a
[50] J.F. Elsworth, R.H. Thomson, A new chamigrane from, Laurencia glomerata. J. new brominated nonaromatic isolaurene type sesquiterpene from red alga,
Nat. Prod. 52 (1989) 893–895, https://doi.org/10.1021/np50064a044. Laurencia heteroclada Harvey. Nat. Prod. Res. 20 (2019) 1–8, https://doi.org/
[51] M. Chang, J.T. Vazquez, K. Nakanishi, F. Cataldo, D.M. Estrada, J. Fernandez, 10.1080/14786419.2019.1655023.
A. Gallardo, J.D. Martin, M. Norte, R. Perez, Regular and irregular sesquiterpenes [77] Y. Minamida, H. Matsuura, C.S. Vairappan, S. Suttikarn, T. Abe, M. Suzuki,
containing a halogenated hydropyran from, Laurencia caespitosa. Phytochem. 28 Teanol, a new brominated sesquiterpene from the Thailand, Laurencia
(1989) 1417–1424, https://doi.org/10.1016/S0031-9422(00)97759-8. mariannensis. Biochem. System. Ecol. 92 (2020), 104093, https://doi.org/
[52] B.M. Fraga, Natural sesquiterpenoids, Nat. Prod. Rep. 14 (1997) 145–162, 10.1016/j.bse.2020.104093.
https://doi.org/10.1039/NP9971400145. [78] X.L. Li, T. Kurtán, J.C. Hu, A. Mándi, J. Li, X.W. Li, Y.W. Guo, Structural and
[53] Conolly, J.D., Hill, R.A., 1991. Dictionary of Terpenoids, Volume 1, mono and stereochemical studies of laurokamurols A-C, uncommon bis-sesquiterpenoids
sesquiterpenoids. Chapman and Hall, London, UK. from the Chinese Red Alga, Laurencia okamurai Yamada. J. Agric. Food Chem. 65
[54] H. Esselin, S. Sutour, J. Liberal, M.T. Cruz, L. Salgueiro, B. Siegler, I. Freuze, (2017) 1550–1555, https://doi.org/10.1021/acs.jafc.6b05238.
V. Castola, M. Paoli, A. Bighelli, F. Tomi, Chemical composition of Laurencia [79] Kamada, T., Vairappan, C.S. 2015. New Laurene-type Sesquiterpene from
obtusa extract and isolation of a new C15-acetogenin, Molecules 22 (2017) 779, Bornean Laurencia nangii. Nat. Prod. Comm. 10, 1934578X1501000. DOI:htt
https://doi.org/10.3390/molecules22050779. ps://doi.org/10.1177/1934578x1501000610.
[55] S. Garcia-Davis, I. Sifaoui, M. Reyes-Batlle, E. Viveros-Valdez, J.E. Piñero, [80] J. Sun, L.J. Han, R.Y. Yang, D.Y. Shi, Z.H. Yuan, J.G. Shi, Studies on chemical
J. Lorenzo-Morales, J.J. Fernández, A.R. Díaz-Marrero, Anti-Acanthamoeba constituents of Laurencia tristicha, Chin. J. Chin. Materia Medica. 32 (2007)
activity of brominated sesquiterpenes from Laurencia johnstonii, Mar. Drugs 16 120–123.
(2018) 443, https://doi.org/10.3390/md16110443. [81] J. Zhang, L.Y. Shi, L.P. Ding, H. Liang, P.F. Tu, Q.Y. Zhang, Antioxidant
[56] J. Zhang, L.-P. Ding, H. Liang, X.-Y. Guo, Q.-Y. Zhang, Sesquiterpenes from the terpenoids from the red alga, Laurencia tristicha, Nat. Prod. Res. 8 (2020) 1–7,
red alga Laurencia tristicha, Biochem. Syst. Ecol. 60 (2015) 116–119, https://doi. https://doi.org/10.1080/14786419.2020.1774762.
org/10.1021/np050096g. [82] M. Rahelivao, M. Gruner, H. Andriamanantoanina, B. Andriamihaja, I. Bauer, H.-
[57] H.S. Alorfi, M.A. Ghandourah, A.J. Turki, Cytotoxic effect of acetogenins and J. Knölker, Red algae (rhodophyta) from the coast of Madagascar: preliminary
sesquiterpenes obtained from the Red alga, Laurencia majuscula. Trop. J. Pharm. bioactivity studies and isolation of natural products, Mar. Drugs 13 (2015)
Res. 19 (2020) 583–586, https://doi.org/10.4314/tjpr.v19i3.18. 4197–4216, https://doi.org/10.3390/md13074197.
[58] N.O. Bawakid, W.M. Alarif, H.S. Alorfi, K.O. Al-Footy, N.A. Alburae, M. [83] M.O. Stallard, W. Fenical, The brasilenols, rearranged sesquiterpenes alcohols
A. Ghandourah, S.S. Al-Lihaibi, Z.H. Abdul-Hameed, Antimicrobial isolated from the marine opsthobranch, Aplysia brasilana. Tetrahedron 34 (1978)
sesquiterpenoids from, Laurencia obtusa Lamouroux. Open Chem. 15 (2017), 2077–2081.
https://doi.org/10.1515/chem-2017-0025. [84] W.M. Alarif, K.O. Al-Footy, M.S. Zubair, H.P.H. Mohamed, M.A. Ghandourah, S.
[59] X.-X. Yang, Y.-Z. Su, J. Tan, C.-E. Cai, P.-M. He, R. Jia, A new dimeric A. Basaif, S.S. Al-Lihaibi, S.-E.N. Ayyad, F.A. Badria, The role of new eudesmane-
sesquiterpene and other related derivatives from the marine red alga, Laurencia type sesquiterpeneoid and known eudesmane derivatives dro the red alga
okamurai. Biochem. System. Ecol. 79 (2018) 57–59, https://doi.org/10.1016/j. Laurencia obtusa as potential antifungal-antitumor agents, Nat. Prod. Res. 30
bse.2018.05.003. (2015) 1150–1155, https://doi.org/10.1080/14786419.2015.1046378.
[60] S.-C. Mao, Y.-W. Guo, Cuparane-derived sesquiterpenes from the Chinese red alga [85] N.Y. Ji, X.M. Li, L.P. Ding, B.G. Wang, Halogenated eudesmane derivatives and
Laurencia okamurai Yamada, Helv. Chim. Acta 88 (2005) 1034–1039, https://doi. other terpenes from the marine red alga, Laurencia pinnata and their
org/10.1002/hlca.200590074. chemotaxonomic significance, Biochem. System. Ecol. 64 (2016) 1–5, https://doi.
[61] S.C. Mao, Y.-W. Guo, A laurane sesquiterpene and rearranged derivatives from org/10.1016/j.bse.2015.11.010.
the chinese red alga, Laurencia okamurai Yamada. J. Nat. Prod. 69 (2006) [86] N.-Y. Ji, W. Wen, X.-M. Li, Q.-Z. Xue, H.-L. Xiao, B.-G. Wang, Brominated Selinane
1209–1211, https://doi.org/10.1021/np0503810. Sesquiterpenes from the marine Brown alga, Dictyopteris divaricata. Mar. Drugs 7
[62] Y. Liang, X.M. Li, C.M. Cui, C.S. Li, B.G. Wang, A new rearranged chamigrane (2009) 355–360, https://doi.org/10.3390/md7030355.
sesquiterpene from, Laurencia okamurai. Chinese Chem. Lett. 20 (2009) 190–192, [87] B.M. Howard, W. Fenical, α- and β-snyderol; new bromo-monocyclic
https://doi.org/10.1016/j.cclet.2008.11.001. sesquiterpenes from the seaweed, Laurencia. Tetrahedron Lett. 17 (1976) 41–44,
[63] T. Irie, M. Suzuki, E. Kurosawa, T. Masamune, Laurinterol, debromolaurinterol https://doi.org/10.1016/S0040-4039(00)71317-7.
and isolaurinterol, constituents of, Laurencia intermedia Yamada. Tetrahedron 26 [88] S. Sutour, H. Esselin, A. Bighelli, J. Casanova, L. Le Gall, F. Tomi, Discrimination
(1970) 3271–3277, https://doi.org/10.1016/S0040-4020(01)92906-0. and characterization of two mediterranean species from the Laurencia complex
[64] M. Suzuki, Y. Hayakawa, T. Irie, Constituents from marine plants. XIV. Acid- (Rhodomelacea) using an NMR-based metabolomic approach, Chem. Biodivers.
catalyzed rearrangement of laurinterol derivatives, Bull. Chem. Soc. Jpn. 42 14 (2017), e1700226, https://doi.org/10.1002/cbdv.201700226.
(1969) 3342–3344, https://doi.org/10.1246/bcsj.42.3342. [89] H. Esselin, S. Sutour, A. Bighelli, F. Tomi, Snyderol derivatives from Laurencia
[65] W.Q. Yu, W.-F. He, D.-Q. Liu, M.-T. Feng, Y. Fang, B. Wang, L.-H. Feng, Y.- obtusa collected in Corsica, Biochem. Syst. Ecol. 82 (2019) 24–26, https://doi.
W. Guo, S.-C. Mao, A seco-laurane sesquiterpene and related laurane derivatives org/10.1016/j.bse.2018.11.002.

14
A.-M. Cikoš et al. Algal Research 56 (2021) 102330

[90] M. Norte, R. Gonzalez, A. Padilla, J.J. Fernandez, J.T. Vazquez, New halogenated [112] R.C. Pereira, B.A.P. da Gama, V.L. Teixeira, Y. Yoneshigue-Valentin, Ecological
sesquiterpenes from the red alga Laurencia caespitosa, Can. J. Chem. 69 (1991) roles of natural products of the Brazilian red seaweed, Laurencia obtusa. Braz. J.
518–520, https://doi.org/10.1139/v91-077. Biol. 63 (2003) 665–672, https://doi.org/10.1590/s1519-69842003000400013.
[91] T. Ishii, Y. Shinjo, M. Miyagi, H. Matsuura, T. Abe, N. Kikuchi, M. Suzuki, [113] A. Ianora, M. Boersma, R. Gasotti, A. Fontana, J. Harder, F. Hoffman, H. Pavia,
Investigation of insect repellent activity of cyclocolorenone obtained from the red P. Potin, S.A. Poulet, G. Toth, New trends in marine chemical ecology, Estuaries
alga, Laurencia intricata. Rec. Nat. Prod. 13 (2019) 81–84, https://doi.org/ Coast. 29 (2006) 531–551, https://doi.org/10.1007/BF02784281.
10.25135/rnp.66.18.02.221. [114] Y. Oguri, M. Watanabe, T. Ishikawa, T. Kamada, C.S. Vairappan, H. Matsuura,
[92] M. Kuniyoshi, M.S. Marma, T. Higa, G. Bernardinelly, C.W. Jefford, New K. Kaneko, T. Ishii, M. Suzuki, E. Yoshimura, Y. Nogata, T. Okino, New marine
bromoterpenes from the red alga, Laurencia luzonensis. J. Nat. Prod. 64 (2001) antifouling compounds from the red alga, Laurencia sp. Mar. Drugs 15 (2017)
696–700, https://doi.org/10.1021/np000638o. 267, https://doi.org/10.3390/md15090267.
[93] Trung, D.V., Truc, N.T.T., Duy, C.N.H., Nhan, L.T.T., Phung, N.K.P., Duong, N.T. [115] S.S. Al-Lihaibi, A. Abdel-Lateff, W.M. Alarif, Y. Nogata, S.-E.N. Ayyad, T. Okino,
T. 2019. Halogenated sesquiterpenes from the red alga Laurencia intermedia Potent antifouling metabolites from red sea organisms, Asian J. Chem. 27 (2015)
Yamada. Vietnam J. Chem. 57, 723–727. DOI: vjch.2019000107. 2252–2256, https://doi.org/10.14233/ajchem.2015.18701.
[94] H. Su, D.-Y. Shi, J. Li, S.-J. Guo, L.-L. Li, Z.-H. Yuan, X.-B. Zhu, Sesquiterpenes [116] M. Protopapa, M. Kotsiri, S. Mouratidis, V. Roussis, E. Ioannou, S.G. Dedos,
from, Laurencia similis. Molecules 14 (2009) 1889–1897, https://doi.org/ Evaluation of antifouling potential and ecotoxicity of secondary metabolites
10.3390/molecules14051889. derived from red algae of the genus, Laurencia. Mar. Drugs 17 (2019) 646,
[95] H. Su, Z.-H. Yuan, S.-J. Guo, L.-P. Deng, L.-J. Han, X.-B. Zhu, D.-Y. Shi, https://doi.org/10.3390/md17110646.
Sesquiterpenes from the marine red alga, Laurencia saitoi. Helv. Chim. Acta 92 [117] N. Nocchi, A.R. Soares, M.L. Souto, J.J. Fernández, M.N. Martin, R.C. Pereira,
(2009) 1291–1297, https://doi.org/10.1002/hlca.200800437. Detection of a chemical cue from the host seaweed Laurencia dendroidea by the
[96] M. Kuniyoshi, P.G. Wahome, T. Miono, T. Hashimoto, M. Yokoyama, K. associated mollusc Aplysia brasiliana, PLoS One 12 (2017), e0187126, https://doi.
L. Shrestha, T. Higa, Terpenoids from, Laurencia luzonensis. J. Nat. Prod. 68 org/10.1371/journal.pone.0187126.
(2005) 1314–1317, https://doi.org/10.1021/np058004a. [118] R.C. Pereira, N. Nocchi, L. Marí, J.J. Fernández Souto, M. Norte, H.M. Duarte, A.
[97] Gonzalez, A.G., Martin, J.D., Martin, V.S., Perez, R., Tagle, B., Clardy, J. 1985. R. Soares, The sea-hare Aplysia brasiliana promotes induction in chemical defense
Rhodolaureol and rhodolauradiol, two new halogenated tricyclic sesquiterpenes in the seaweed Laurencia dendroidea and in their congeneric neighbors, Plant
from a marine alga. J. Chem. Soc. Chem. Commun. 260–261. DOI: https://doi. Physiol. Biochem. 154 (2020) 295–303, https://doi.org/10.1016/j.
org/10.1039/C39850000260. plaphy.2020.05.020.
[98] da Silva Machado, F.L.S., Duarte, H.M., Gestinari, L.M.S., Cassano, V., Kaiser, C. [119] Cikoš, A.-M., Jurin, M., Čož-Rakovac, R., Jokić, S., Jerković, I., 2019. Update on
R., Soares, A.R. 2016. Geographic distribution of natural products produced by monoterpenes from red macroalgae: isolation, analysis, and bioactivity. Mar.
the red alga Laurencia dendroidea J.Agardh. Chem.Biodivers., 13, 845–851. DOI: Drugs 17, 537. DOI: https://doi.org/10.3390/md17090537.
https://doi.org/10.1002/cbdv.201500246. [120] L. Ružička, M. Stoll, Höhere Terpenverbindungen VII. Konstitution des Eudalins,
[99] T. Kamada, C.S. Vairappan, Nonhalogenated new sesquiterpenes from Bornean, des Selinens und des α-Santalens. Das Kohlenstoffgerüst der Sesquiterpene, Helv.
Laurencia snackeyi, Nat. Prod. Res. 31 (2017) 333–340, https://doi.org/10.1080/ Chim. Acta 5 (1922) 923–936, https://doi.org/10.1002/hlca.19220050613.
14786419.2016.1241996. [121] P. Arendt, J. Pollier, N. Callewaert, A. Goossens, Synthetic biology for production
[100] T. Kamada, C.S. Vairappan, New bioactive secondary metabolites from bornean of natural and new-to-nature terpenoids in photosynthetic organisms, Plant J. 87
red alga, Laurencia similis (Ceramiales), Nat. Prod. Commun. 8 (2013) 287–288, (2016) 16–37, https://doi.org/10.1111/tpj.13138.
https://doi.org/10.1177/1934578X1300800302. [122] Bhattacharya, D., Price, D.C., Chan, C.X., Qiu, H., Rose, N., Ball, S., Weber, A.P.
[101] Zaleta-Pinet, D.A., Holland, I.P., Munoz-Ochoa, M., Murillo-Alvarez, J.I., Sakoff, M., Arias, M.C., Henrissat, B., Coutinho, P.M., Krishnan, A., Zäuner, S., Morath, S.,
J.A., van Altena, I.A., McCluskey, A., 2014. Cytotoxic compounds from Laurencia Hilliou, F., Egizi, A., Perrineau, M.-M., Yoon, H.S., 2013. Genome of the red alga
pacifica. Org. Med. Chem. Lett. 4, 8. DOI: https://doi.org/10.1186/s13588 Porphyridium purpureum. Nat. Commun. 4:1941 Biosynthesis Nat. Prod. Rep. 28,
-014-0008-8. 1917–1936. DOI: https://doi.org/10.1038/ncomms2931.
[102] S. Garcia-Davis, E. Viveros-Valdez, A.R. Díaz-Marrero, J.J. Fernández, [123] L.S. De Oliviera, G. Bueno Gregoracci, G.G. Zacarias Silva, L. Tavares Saldago, G.
D. Valencia-Mercado, O. Esquivel-Hernández, P. Carranza-Rosales, I.E. Carranza- A. Filhjo, M. Alves-Ferreira, R. Crespo Pereira, F.L. Thompson, Transcriptomic
Torres, N.E. Guzmán-Delgado, Antitumoral effect of laurinterol on 3D culture of analysis of the red seaweed Laurencia dendroidea (Florideophyceae, Rhodophyta)
breast cancer explants, Mar. Drugs 17 (2019) 201, https://doi.org/10.3390/ and its microbiome, BMC Genomics 13 (2012) 487, https://doi.org/10.1186/
md17040201. 1471-2164-13-487.
[103] W.A.J.P. Wijesinghe, E.-A. Kim, M.-C. Kang, W.-W. Lee, H.-S. Lee, C.S. Vairappan, [124] L.S. De Oliviera, D.A. Tschoeke, A.S. de Oliveira, L.J. Hill, W.C. Paradas, L.
Y.J. Jeon, Assessment of anti-inflammatory effect of 5β-hydroxypalisadin B T. Salgado, C.C. Thompson, R.C. Pereira, F.L. Thompson, New insights on the
isolated from red seaweed Laurencia snackeyi in zebrafish embryo in vivo model, terpenome of the red seaweed, Laurencia dendroidea (Florideophyceae,
Environ. Toxicol. Pharmacol. 37 (2014) 110–117, https://doi.org/10.1016/j. Rhodophyta). Mar. Drugs 13 (2015) 879–902, https://doi.org/10.3390/
etap.2013.11.006. md13020879.
[104] C.S. Vairappan, T. Kamada, W.-W. Lee, Y.-J. Jeon, Anti-inflammatory activity of [125] M. Ashour, M. Wink, J. Gershenzon, Biochemistry of terpenoids: monoterpenes,
halogenated secondary metabolites of Laurencia snackeyi (Weber-van Bosse) sesquiterpenes and diterpenes, Annu, Plant Rev 40 (2010) 258–303, https://doi.
Masuda in LPS-stimulated RAW 264.7 macrophages, J. Appl. Phycol. 25 (2013) org/10.1002/9781444320503.ch5.
1805–1813, https://doi.org/10.1007/s10811-013-0023-6. [126] D.E. Cane, Enzymatic formation of sesquiterpenes, Chem. Rev. 90 (1990)
[105] T.L. Biá Ventura, F.L. da Silva Machado, M.H. de Araujo, L.M. de Souza Gestinari, 1089–1103, https://doi.org/10.1021/cr00105a002.
C.R. Kaiser, F.D. Esteves, E.B. Lasunskaia, A. Ribeiro Soares, M.F. Muzitano, Nitric [127] J.S. Dickschat, Isoprenoids in three-dimensional space: the stereochemistry of
oxide production inhibition and anti-mycobacterial activity of extracts and terpene biosynthesis, Nat. Prod. Rep. 28 (2011) 1917–1936, https://doi.org/
halogenated sesquiterpenes from the brazilian red alga, Laurencia dendroidea J. 10.1039/c1np00063b.
Agardh. Pharmacogn. Mag. 11 (2015) 611–618, https://doi.org/10.4103/0973- [128] De Carvalho, L.R., Fujii, M.T., Roque, N.F., Kato, M.J., Lago, J.H.G., 2003.
1296.172972. Aldingenin A, new brominated sesquiterpene from red algae Laurencia aldingensis.
[106] S. Garcia-Davis, K. Leal-López, C.A. Molina-Torres, L. Vera-Cabrera, A.R. Díaz- Tetrahedron Lett. 44, 2637–264. DOI: https://doi.org/10.1002/chin.200326192.
Marrero, J.J. Fernández, P. Carranza-Rosales, E. Viveros-Valdez, [129] W. Parker, J.S. Roberts, R. Ramage, Sesquiterpenes Biogenesis, Q. Rev. Chem.
Antimycobacterial activity of Laurinterol and Aplysin from, Laurencia johnstonii. Soc. 21 (1967) 331–363, https://doi.org/10.1039/QR9672100331.
Mar. Drugs 18 (2020) 287, https://doi.org/10.3390/md18060287. [130] J. Sun, D. Shi, M. Ma, S. Li, S. Wang, L. Han, Y. Yang, X. Fan, J. Shi, L. He,
[107] I. Arberas-Jiménez, S. García-Davis, A. Rizo-Liendo, I. Sifaoui, M. Reyes-Batlle, Sesquiterpenes from the red alga, Laurencia tristicha. J. Nat. Prod. 68 (2005)
O. Chiboub, R. Rodriguez-Exposito, A.R. Diaz-Marrero, J.E. Pinero, J. 915–919, https://doi.org/10.1021/np050096g.
J. Fernandez, J. Lorenzo-Morales, Laurinterol from Laurencia johnstonii eliminates [131] C.L. Steele, J. Crock, J. Bohlmann, R. Croteau, Sesquiterpene synthases from
Naegleria fowleri triggering PCD by inhibition of ATPases, Sci. Rep. 10 (2020) grand fir (Abies grandis): comparison of constitutive and wound-induced
17731, https://doi.org/10.1038/s41598-020-74729-y. activities, and cDNA isolation, characterization and bacterial expression of delta-
[108] A. Campos, C. Borges Souzaa, C. Lhullierb, M. Falkenbergb, E.P. Schenkelb, R. selinene synthase and gamma-humulene synthase, J. Biol. Chem. 273 (1998)
M. Ribeiro-do-Vallea, J. Mota Siqueira, Anti-tumour effects of elatol, a marine 2078–2089, https://doi.org/10.1074/jbc.273.4.2078.
derivative compound obtained from red algae, Laurencia microcladia. J. Pharm. [132] Martın, J.D., Darias, J., 1978. In: Scheuer, P. (ed), Marine Natural Products:
Pharmacol. 64 (2012) 1146–1154, https://doi.org/10.1111/j.2042- Chemical and Biological Perspectives. Academic, New York, pp. 125–174.
7158.2012.01493.x. [133] K.A. Rising, C.M. Starks, J.P. Noel, J. Chappell, Demonstration of germacrene A as
[109] K.G. Gonçalves, L.L. da Silva, A.R. Soares, N.C. Romeiro, Acetylcholinesterase as a an intermediate in 5-epi-aristolochene synthase catalysis, J. Am. Chem. Soc. 122
target of halogenated marine natural products from, Laurencia dendroidea. Algal (2000) 1861–1866, https://doi.org/10.1021/ja993584h.
Res. 52 (2020), 102130, https://doi.org/10.1016/j.algal.2020.102130. [134] P. Scheuer, Marine Natural Products V1: Chemical and Biological Perspectives,
[110] Paul, V.J., Cruz-Rivera E., Thacker, R.W. 2001. Chemical mediation of Academic Press, New York, 2012.
macroalgal-herbivore interactions: ecological and evolutionary perspectives. In: [135] G. Wei, Q. Jia, X. Chen, T.G. Kӧllner, D. Bhattacharya, G.K.-S. Wong,
(McClintock, J.B., Baker, B.J., eds), Marine chemical ecology. CRC Press, Boca J. Gershenzon, F. Chen, Terpene biosynthesis in red algae is catalyzed by
Raton, FL. pp. 227–265. microbial type but not typical plant terpene synthases, Plant Physiol. 179 (2019)
[111] Pereira, R.C., da Gama, B.A.P. 2008. Macroalgal chemical defenses and their roles 382–390, https://doi.org/10.1104/pp.18.01413.
in structuring tropical marine communities. In: (C.D. Amsler, ed) Algal Chemical
Ecology. Springer-Verlag, Berlin/Heidelberg. pp. 25–55.

15

You might also like