Benzenediol lactones: a class of fungal metabolites with diverse structural features and biological activities

Benzenediol lactones: a class of fungal metabolites with diverse structural features and biological activities

Accepted Manuscript Benzenediol lactones: a class of fungal metabolites with diverse structural features and biological activities Weiyun Shen , Hongq...

2MB Sizes 107 Downloads 129 Views

Accepted Manuscript Benzenediol lactones: a class of fungal metabolites with diverse structural features and biological activities Weiyun Shen , Hongqiang Mao , Qian Huang , Jinyan Dong PII:

S0223-5234(14)01111-8

DOI:

10.1016/j.ejmech.2014.11.067

Reference:

EJMECH 7562

To appear in:

European Journal of Medicinal Chemistry

Received Date: 29 July 2014 Revised Date:

4 November 2014

Accepted Date: 26 November 2014

Please cite this article as: W. Shen, H. Mao, Q. Huang, J. Dong, Benzenediol lactones: a class of fungal metabolites with diverse structural features and biological activities, European Journal of Medicinal Chemistry (2015), doi: 10.1016/j.ejmech.2014.11.067. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

TITLE PAGE

2

Benzenediol lactones: a class of fungal metabolites with

3

diverse structural features and biological activities

4

Weiyun Shen, Hongqiang Mao, Qian Huang & Jinyan Dong*

Key Laboratory of Eco-environments in Three Gorges Reservoir Region (Ministry of Education), School of Life Science, Southwest University, Chongqing 400715, People's Republic of China

SC

*E-mail: [email protected], Tel: 86-023-68252400

Abstract: Benzenediol lactones are a structurally variable family of fungal polyketide metabolites possessing a macrolide core structure fused into a resorcinol aromatic ring. These compounds are

M AN U

widespread in fungi mainly in the genera such as Aigialus, Cochliobolus, Curvularia, Fusarium, Humicola, Lasiodiplodia, Penicillium and Pochonia etc. Most of these fungal metabolites were reported to possess several interesting biological activities, such as cytotoxicities, nematicidal properties, inhibition of various kinases, receptor agonists, anti-inflammatory activities, heat shock response and immune system modulatory activities etc. This review summarizes the research on the isolation, structure elucidation, and biological activities of the benzenediol lactones, along with some available structure–activity relationships, biosynthetic studies, first syntheses, and syntheses that lead to the revision of structure or stereochemistry, published up to the year of 2014. More than 190

TE D

benzenediol lactones are described, and over 300 references cited. Keywords: Secondary metabolites; Benzenediol lactones; Dihydroxyphenylacetic acid lactones;

EP

Resorcylic acid lactones; Bioactivities; Structure-activity relationships.

AC C

5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

RI PT

1

1

ACCEPTED MANUSCRIPT[Benzenediol lactones]

21

Fungi have been proved to be a fertile and important biosource of numerous secondary metabolites with a huge variety of chemical structures and diverse bioactivities. Fungal metabolites are of considerable synthetic interest and remarkable importance as new lead compounds for medicine as well as for plant protection. Importantly, fungal polyketides are one of the largest and most structurally diverse classes of naturally occurring compounds, ranging from simple aromatic metabolites to complex macrocyclic lactones [1-8].

RI PT

Among these polyketide secondary metabolites, benzenediol lactone (BDL) is a growing class which is defined by a 1, 3- benzenediol moiety bridged by a macrocyclic lactone ring [9]. They represent a rich scaffold for structural variation, which differ particularly in the degree and positioning of oxidation and unsaturation about the macrolactone ring. Meanwhile, the phenolic hydroxyl groups on the benzene ring can be methylated. The BDL family may be split into resorcylic acid lactones (RALs) and dihydroxyphenylacetic acid lactones (DALs) based on a substituted resorcinol fragment

SC

fused to α, β-, and β, γ-positions of the macrocyclic lactone ring, respectively (Fig. 1) [10, 11]. To better describe these structures, there exist different numbering systems. Take 14-membered RALs for example, the older system uses the numbers 1-6 for the aromatic ring and 1'-12' for the aliphatic

M AN U

macrocycle, whereas the more recent IUPAC system counts the C-atoms from 1-18 (Fig. 2). This situation, although unsatisfactory, did not cause major problems to date, because the number of BDLs was quite limited [12]. Considering the coherence of the previous review, here we retain both systems. Fig. 1. Basic scaffold of benzenediol lactones

Fig. 2. Different numbering systems of 14-membered RALs

23 24 25 26 27 28 29 30 31 32

AC C

EP

22

1. Introduction

TE D

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

To date, over 190 natural BDLs were found in numerous fungi mainly belonging to the genera

Aigialus, Cochliobolus, Curvularia, Fusarium, Humicola, Lasiodiplodia, Penicillium and Pochonia etc. Table 1 shows all the natural BDLs discovered including names, fungal sources and references published between the end of 1953 and July 2014. Although these BDLs are anabolited by many different fungal species, they are known to be produced via a pair of similar collaborating iterative polyketide synthases (iPKSs): a highly reducing iPKS (hrPKS) with product that is further elaborated by a nonreducing iPKS (nrPKS) to yield a 1, 3- benzenediol moiety bridged by a macrolactone [11, 13]. Of course, subtle differences exist between the biosynthetic pathways for different compounds. In recent years, these fungal metabolites were reported to exhibit a wide range of significant 2

ACCEPTED MANUSCRIPT[Benzenediol lactones] biological activities such as inhibition of heat shock protein 90 (Hsp90) and kinases [14], as well as antimicrobial [9], cytotoxic [15], antineoplastic [16], and anti-inflammatory activities [17]. In view of their promising biological activities and interesting structural characteristics, BDLs have increasingly received a great deal of research focus. Since the 1970s, a number of synthetic studies on BDLs have been disclosed. The general synthetic routes usually involve olefin metathesis together with some classical chemical reactions (such as the Heck coupling reaction, the Witting reaction, the Mitsunobu reaction, the Stille coupling reaction, Diels-Alder reaction etc) [18].

RI PT

Some BDLs have already been mentioned in a number of certain bioactivity reviews [19-23]. However, these authors cover only a fraction of all known BDLs, since their articles naturally omit the compounds with other activities not included in their reviews. In addition, the resorcylic acid lactones containing a cis-enone or with kinase inhibitory activities of this family have been frequently reviewed by several authors [24-26]. In 2007, Winssinger and Barluenga published a topical review article

SC

focused on the biosynthesis, chemical synthesis, and biological activity of RALs covered the literature up to 2007. Yet, it only described ten typical RALs in detail [9]. Recently, a review from Xu et al. [18] described 60 resorcinolic macrolides concerning their isolation, bioactivities, biosyntheses, and representative chemical syntheses in recent decades. According to their literature, over 60 different

M AN U

resorcinolic macrolides are produced by fungi. However, we found a significantly higher number of relevant fungal metabolites up to 2014: Around 190 compounds are presented in this review. In this review we survey the chemical and biological literatures regarding the isolation, structure elucidation, biological activities, biosyntheses, and chemical syntheses of BDL derivatives from nature. Additionally, those available structure-activity relationships and action mechanisms of some bioactive compounds will also be discussed. We focus on work that has appeared in the literatures up to July 2014. We also hope that we can provide some references for further development and utilization of this

TE D

kind of natural compounds. For ease of illustration and comparison, these natural products are divided into six subclasses according to their ring sizes and structural characteristics: 8-membered BDLs, 10-membered BDLs, 12-membered BDLs, 13-membered BDLs, 14-membered BDLs and other BDLs. Table 1. Benzenediol lactones (BDLs) Name (stereochemistry) (alternative name)

Producing Species

References

Coryoctalactone A (1)

Corynespora cassiicola

[27]

Coryoctalactone B (2)

Corynespora cassiicola

[27]

Coryoctalactone C (3)

Corynespora cassiicola

[27]

Coryoctalactone D (4)

Corynespora cassiicola

[27]

Coryoctalactone E (5)

Corynespora cassiicola

[27]

EP

Benzenediol octalactone

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27

Benzenediol decalactone Sporostatin (6)

Sporormiella sp.

[28]

Xestodecalactone A (7)

Penicillium cf. montanense

[31]

Xestodecalactone B (8)

Penicillium cf. montanense

[31]

Xestodecalactone C (9)

Penicillium cf. montanense

[31]

Xestodecalactone D (10)

Corynespora cassiicola

[36]

Xestodecalactone E (11)

Corynespora cassiicola

[36]

Xestodecalactone F (12)

Corynespora cassiicola

[36]

(3R,5R)-Sonnerlactone (13)

Zh6-B1

[37]

3

ACCEPTED MANUSCRIPT[Benzenediol lactones] (3R,5S)-Sonnerlactone (14)

Zh6-B1

[37]

Relgro (15)

Fusarium sp.

[40]

Lasiodiplodia theobromae

[56]

Sarocladium kiliense

[72]

Syncephalastrum racemosum

[70]

unidentified endophytic fungus

[27]

Lasiodiplodia theobromae

[56]

(3R)-Lasiodiplodin (16)

(R)-de-O-Methyllasiodiplodin (17)

RI PT

12-Membered benzenediol lactones

Syncephalastrum racemosum

[70]

unidentified endophytic fungus

[27]

5-oxo-Masiodiplodin (18)

Lasiodiplodia theobromae

[58]

(3R, 5R)-Hydroxylasiodiplodin (19)

Lasiodiplodia theobromae

SC

Syncephalastrum racemosum

[58] [70]

Lasiodiplodia theobromae

[58]

Sarocladium kiliense

[73]

(3R, 4S)-4-Hydroxylasiodiplodin (21)

Lasiodiplodia theobromae

[58]

(3R, 5R)-5-Hydroxyl-de-O-methyllasiodiplodin (22)

Lasiodiplodia theobromae

[58]

unidentified endophytic fungus

[27]

Lasiodiplodia theobromae

[58]

Lasiodiplodia theobromae

[59]

unidentified endophytic fungus

[27]

Syncephalastrum racemosum

[70]

unidentified endophytic fungus

[27]

Lasiodiplodia theobromae

[68]

(3R,4R)-4-Hydroxy-de-O-methyllasiodiplodin (28)

Lasiodiplodia theobromae

[69]

(E)-9-Etheno-de-O-methyllasiodiplodin (29)

Lasiodiplodia theobromae

[69]

(3R, 5S)-5-Hydroxyl-de-O-methyllasiodiplodin (30)

Syncephalastrum racemosum

[70]

(3S, 6R)-6-Hydroxylasiodiplodin (31)

Sarocladium kiliense

[73]

Curvularin (32)

Alternaria cinerariae

[77]

Alternaria zinniae

[78]

Ascochytula obiones

[79]

Beauveria bassiana

[80]

Chrysosporium lobatum

[84]

Cochliobolus spicifer

[81]

Curvularia eragrostidis

[83]

Curvularia inaequalis

[82]

Curvularia lunata

[76]

Curvularia pallescens

[83]

Drechslera australiensis

[85]

Eupenicillium sp.

[86]

Helminthosporium maydis

[87]

Penicillium baradicum

[89]

Penicillium gilmanii

[88]

Penicillium sumatrense

[92]

(3R, 6R)-6-Hydroxyl-de-O-methyllasiodiplodin (23) (3R, 6S)-6-Hydroxylasiodiplodin (24) 6-oxo-de-O-Methyllasiodiplodin (25)

(E)-9-Etheno-lasiodiplodin (26)

AC C

EP

TE D

Botryosphaeriodiplodin (27)

M AN U

(3R, 5S)-Hydroxylasiodiplodin (20)

4

ACCEPTED MANUSCRIPT[Benzenediol lactones] [91, 108, 121]

Ulocladium sp.

[94]

α, β-Dehydrocurvularin (33)

Alternaria macrospora

[78]

(also named trans-dehydrocurvularin, dehydrocurvularin)

Alternaria cucumerina

[105]

Alternaria cinerariae

[107]

Alternaria zinniae

[109]

Alternaria tomato

[106]

Aspergillus sp.

[110]

Aspergillus terreus Chrysosporium lobatum Curvularia eragrostidis Curvularia lunata

SC

Curvularia pallescens

RI PT

Penicillium citreo-viride

[111] [84] [83] [76]

[83];

[113]

Drechslera australiensis

[85]

Eupenicillium sp.

[86]

Helminthosporium maydis

[87]

Nectria galligena

[114]

Penicillium turbatum

[90]

Penicillium citreo-viride

[108]

Stemphylium radicinum

[115]

9F series marine fungi

[116]

Ulocladium sp.

[94]

Aspergillus terreus

[114]

Penicillium sp.

[122]

Penicillium citreo-viride

[108]

Penicillium sp.

[112]

Alternaria tomato

[106]

Penicillium citreo-viride

[108]

Penicillium sp.

[112]

Aspergillus terreus

[111]

Penicillium sp.

[122]

Penicillium citreo-viride

[108]

Penicillium sp.

[112]

Curvularia oryzae

[125]

Penicillium citreo-viride

[108]

Penicillium sp.

[112]

Cis-dehydrocurvularin (36)

Penicillium citreo-viride

[91, 108, 121]

12-oxo-Curvularin (37)

Aspergillus sp.

[110]

(also named 7-oxo-curvularin)

Penicillium citreo-viride

[91, 108, 121]

11-β-Hydroxy-12-oxo-curvularin (38)

Aspergillus sp.

[110]

(also named 8-β-Hydroxy-7-oxo-curvularin)

Penicillium citreo-viride

[91, 108, 121]

Penicillium sp.

[112]

Penicillium citreo

[91, 108, 121]

M AN U

Cercospora scirpicola

11-α- Hydroxycurvularin (34a)

EP

11-β- Hydroxycurvularin (34b)

TE D

11-Hydroxycurvularin (34)

AC C

11-Methoxycurvularin (35)

11-α-Methoxycurvularin (35a)

11-β-Methoxycurvularin (35b)

11, 12- Dihydroxycurvularin (39)

5

ACCEPTED MANUSCRIPT[Benzenediol lactones] Penicillium citreo

[91, 108, 121]

Citreofuran (41)

Penicillium citreo-viride

[91, 108, 121]

Penilactone (42)

Penicillium sp.

[126]

10,11-Epoxycurvularin (43)

Penicillium sp.

[126]

β,γ-Dehydrocurvularin (44)

Aspergillus sp.

[110]

E-6-Chloro-10, 11-dehydrocurvularin (45)

Cochliobolus spicifer

[81]

11-O-Acetyldehydrocurvularin (46)

Cercospora scirpicola

[113]

Curvularin-7-O-β-D-glucopyranoside (47)

Beauveria bassiana

[80]

Curvularin-4'-O-methyl-7-O-β-D-glucopyranoside (48)

Beauveria bassiana

6-Hdroxycurvularin-4'-O-methyl-6-O-β-D-glucopyranoside

Beauveria bassiana

(49) Oxacyclododecindione (50)

Exserohilum rostratum

(+)-(15R)-10,11-E-Dehydrocurvularin (51)

Curvularia sp.

(+)-(15R)-12-Hydroxy-10,11-E-dehydrocurvularin (52)

Curvularia sp.

(+)-(15R)-13-Hydroxy-10,11-E-dehydrocurvularin (53)

Curvularia sp.

(+)-(11S,15R)-11-Hydroxycurvularin (54)

Curvularia sp.

(+)-(11R,15R)-11-Hydroxycurvularin (55)

Curvularia sp.

RI PT

12-Hydroxy-10, 11- trans- dehydrocurvularin (40)

[80] [80]

[127]

M AN U

SC

[128] [128] [128] [128] [128]

Curvularia sp.

[128]

Curvularia sp.

[129]

Penicillium sumatrense

[130]

Penicillium sumatrense

[130]

Penicillium sumatrense

[130]

Drechslera phlei

[133]

Penicillium sp.

[131, 135]

Pyrenophora teres

[132]

Drechslera phlei

[133]

Penicillium roseopurpureum

[136]

Penicillium sp.

[131, 135]

Pyrenophora teres

[132]

Drechslera phlei

[133]

Penicillium sp.

[135]

Acremonium zeae

[134]

Drechslera phlei

[133]

Penicillium sp.

[135]

Acremonium zeae

[134]

(R)-7- Methoxydihydroresorcylide (65)

Penicillium sp.

[135]

(S)-7- Methoxydihydroresorcylide (66)

Penicillium sp.

[135]

Dihydroresorcylide (67)

Acremonium zeae

[134]

(68)

Aigialus parvus

[144]

(69)

Aigialus parvus

[144]

Aigialomycin A (70)

Aigialus parvus

[143]

Aigialomycin B (71)

Aigialus parvus

[143]

(+)-(15R)-12-Oxocurvularin (56) Curvulone A (57) Sumalarins A (58) Sumalarins B (59) Sumalarins C (60)

TE D

Trans- resorcylide (61)

EP

Cis-resorcylide (62)

AC C

(R)-7-Hydroxydihydroresorcylide (63)

(S)-7-Hydroxydihydroresorcylide (64)

13-Membered benzenediol lactones

14-Membered benzenediol lactones

6

ACCEPTED MANUSCRIPT[Benzenediol lactones] [157]

Cochliobolus lunatus

[158]

Aigialus parvus

[143]

Paecilomyces sp.

[157]

Aigialus parvus

[143]

Fusarium sp.

[156]

Paecilomyces sp.

[219]

Aigialomycin E (74)

Aigialus parvus

[143]

Aigialomycin F (75)

Aigialus parvus

Aigialomycin C (72)

Aigialomycin D (73)

Paecilomyces sp. Aigialomycin G (76)

Aigialus parvus

1', 2'-Epoxyaigialomycin D (77)

Hypomyces subiculosus

SC

Paecilomyces sp.

RI PT

Paecilomyces sp.

[144] [157] [144] [155] [157]

Cochliobolus lunatus

[158]

Caryospomycins A (79)

Caryospora callicarpa

[159]

Caryospomycins B (80)

Caryospora callicarpa

[159]

Caryospomycins C (81)

Caryospora callicarpa

[159]

Cochliobolus lunatus

[161]

Cochliobolus lunatus

[161]

Cochliobolus lunatus

[161]

Cochliobolus lunatus

[158]

Cochliobolus lunatus

[158]

Cochliobolus lunatus

[158]

Hamigera avellanea

[142]

Hamigera avellanea

[142]

Hamigera avellanea

[165]

Hamigera avellanea

[165]

Hamigera avellanea

[165]

Hamigera avellanea

[165]

Hamigeromycin G (94)

Hamigera avellanea

[165]

Hypothemycin (95)

Aigialus parvus

[143]

Coriolus versicolor

[168]

Hypomyces subiculosis

[155]

Hypomyces trichothecoides

[173, 174]

Phoma sp.

[171]

7', 8'-Dihydrohypothemycin (96)

Hypomyces trichothecoides

[174]

5'-O-Methylhypothemycin (97)

Phoma sp.

[171]

L-783,277 (98)

Phoma sp.

[182]

L-783,290 (99)

Phoma sp.

[182]

4-O-Demethylhypothemycin (100)

Aigialus parvus

[143]

Hypomyces subiculosus

[155]

(101)

Hypomyces subiculosus

[155]

(102)

Hypomyces subiculosus

[155]

Monorden (103)

Chaetomium chiversii

[193]

M AN U

Deoxyaigialomycin C (78)

Cochliomycin A (82) Cochliomycin B (83) Cochliomycin C (84) Cochliomycin D (85) Cochliomycin E (86) Cochliomycin F (87)

Hamigeromycin B (89) Hamigeromycin C (90) Hamigeromycin D (91) Hamigeromycin E (92)

AC C

EP

Hamigeromycin F (93)

TE D

Hamigeromycin A (88)

7

ACCEPTED MANUSCRIPT[Benzenediol lactones] Colletotrichum graminicola

[194]

Humicola fuscoatra,

[200]

Humicola sp.

[195, 196]

Monocillium nordinii

[183, 213]

Neocosmospora sp.

[218]

Neonectria radicicola

[184]

Neophaeosphaeria quadriseptata

[193, 197]

Paecilomyces sp.

[157]

Pochonia chlamydosporia

RI PT

(also named radicicol, or monorden A)

Penicillium luteo-aurantium

[222] [198] [199]

Humicola sp.

Tetrahydromonorden (104)

[196]

SC

(also named monorden B) Humicola sp.

Monorden D (106)

Humicola sp.

(also named pochonin D)

Pochonia chlamdosporia

[212]

Monorden E (107)

Humicola sp.

[196]

M AN U

Monorden C (105)

[196] [196]

Chaetomium chiversii

[193];

Colletotrichum graminicola

[194]

Monocillium nordinii

[213, 214]

Paraphaeosphaeria quadriseptata

[193, 197]

Colletotrichum graminicola

[194]

Monocillium nordinii

[213, 214]

Neocosmospora sp.

[218]

Pochonia chlamydosporia

[212]

Colletotrichum graminicola

[194]

Monocillium nordinii

[213, 214]

Pochonia chlamydosporia

[212]

Monocillium nordinii

[213, 214]

Neocosmospora sp.

[218]

Humicola fuscoatra

[205]

Penicillium sp.

[204]

Monocillium nordinii

[213, 214]

Nordinone (113)

Monocillium nordinii

[214]

Nordinonediol (114)

Monocillium nordinii

[214]

Monocillin II glycoside (115)

Pochonia chlamydosporia

[212]

Monorden analogue-1 (116)

Humicola fuscoatra

[200]

Pochonia chlamdosporia

[216]

Radicicol B (117)

Humicola fuscoatra

[217]

Radicicol C (118)

Humicola fuscoatra

[217]

Radicicol D (119)

Humicola fuscoatra

[217]

Neocosmosin A (120)

Neocosmospora sp.

[218]

Neocosmosin B (121)

Neocosmospora sp.

[218]

Monocillin I (108)

TE D

Monocillin II (109)

Monocillin III (110)

AC C

EP

Monocillin IV (111)

Monocillin V (112)

(also named tetrahydromonocillin I)

8

ACCEPTED MANUSCRIPT[Benzenediol lactones] Neocosmospora sp.

[218]

Paecilomycin A (123)

Paecilomyces sp.

[219]

Paecilomycin B (124)

Paecilomyces sp.

[219]

Paecilomycin C (125)

Paecilomyces sp.

[219]

Paecilomycin D (126)

Paecilomyces sp.

[219]

Paecilomycin E (127)

Paecilomyces sp.

[219]

Paecilomycin F (128)

Cochliobolus lunatus

[158]

Paecilomyces sp.

[219]

RI PT

Neocosmosin C (122)

Paecilomyces sp.

Paecilomycin H (130)

Paecilomyces sp.

Paecilomycin I (131)

Paecilomyces sp.

Paecilomycin J (132)

Paecilomyces sp.

Paecilomycin K (133)

Paecilomyces sp.

Paecilomycin L (134)

Paecilomyces sp.

[221]

Paecilomycin M (135)

Paecilomyces sp.

[221]

Pochonin A (136)

Pochonia chlamydosporia

[212]

Pochonin B (137)

Humicola fuscoatra

[217]

Pochonia chlamydosporia

[212]

Humicola fuscoatra

[217]

Pochonia chlamydosporia

[212]

Pochonia chlamdosporia

[212, 225]

Pochonia chlamdosporia

[212, 225]

Pochonia chlamdosporia

[216]

Pochonia chlamdosporia

[216]

Pochonia chlamdosporia

[216]

Pochonia chlamdosporia

[216]

Pochonia chlamdosporia

[216]

Pochonia chlamdosporia

[216]

Pochonia chlamdosporia

[216]

Fusarium sp.

[156]

Humicola fuscoatra

[217]

Pochonia chlamdosporia

[216]

Pochonin O (149)

Pochonia chlamdosporia

[216]

Pochonin P (150)

Pochonia chlamdosporia

[216]

LL-Z1640-1 (151)

Lederle Culture 21640

[231]

(also named L-783,279)

Cochliobolus lunatus;

[161]

Drechslera portulacae;

[234]

filamentous fungus(MSX63935)

[15]

Paecilomyces sp.

[157]

Sterile fungus (MF6280&6293)

[235]

LL-Z1640-2 (152)

Lederle Culture 21640

[231]

(also named C292, 5Z-7-oxo-zeaenol or L-783,278)

Cochliobolus lunatus

[161, 233]

M AN U

SC

Paecilomycin G (129)

Pochonin C (138)

Pochonin D (106) (also named Monorden D ) Pochonin E (139)

TE D

Pochonin F (140) Pochonin G (141) Pochonin H (142) Pochonin I (143) Pochonin J (144)

Pochonin L (146) Pochonin M (147)

AC C

Pochonin N (148)

EP

Pochonin K (145)

[157] [157] [157] [221] [221]

9

ACCEPTED MANUSCRIPT[Benzenediol lactones] [15, 233]

Sterile fungus (MF6280&6293)

[235]

LL-Z1640-3 (153)

Lederle Culture 21640

[231]

LL-Z1640-4 (154)

Lederle Culture 21640

[231]

Zeaenol (155)

Cochliobolus lunatus

[232]

Drechslera portulacae

[234]

filamentous fungus (MSX63935)

[15]

Paecilomyces sp.

[157]

(5E)-7-oxo-Zeaenol (156)

Cochliobolus lunatus

(also named (7'E)-6'-oxozeaenol)

Drechslera portulacae

RI PT

filamentous fungus (MSX63935)

[158, 233] [234] [15]

5,6-Dihydo-5-methoxy-7-oxo-zeaenol (157)

Cochliobolus lunatus

[233]

15-O-Desmethyl-(5Z)-7-oxozeaenol (158)

filamentous fungus (MSX63935)

Zearalenone (ZEA, ZEN) (159)

Fusarium sp.

SC

filamentous fungus (MSX63935)

[15]

[ 242, 277, 279] [220]

13-Formyl-zearalenone (5-formylzearalenone) (160)

Fusarium graminearum

[277]

5, 6-Dehydro-zearalenone (7'-dehydrozearalenone ) (161)

Fusarium graminearum

[277]

Cochliobolus lunatus

[40]

Paecilomyces sp.

[220]

Fusarium graminearum

[271, 272]

Fusarium graminearum

[277];

Streptomyces rimosus

[286]

Penicillium sp.

[300]

Fusarium graminearum

[277]

5-Hydroxy-zearalenol (6' 8'-Dihydroxyzearalene) (163)

Fusarium graminearum

[275]

10-Hydroxy-zearalenone (3'-Hydroxyzearalenones) (164)

Fusarium graminearum

[276]

Fusarium sp.

[273, 274, 277]

Zearalanone (166)

Fusarium sp.

[277]

α-Zearalanol (also named zeranol) (167a)

Fusarium sp.

[277]

Cis-zearalenone (168)

Fusarium sp.

[277]

Cis-α-zearalenol (169a)

Fusarium sp.

[277]

Cunninghamella bainieri

[286]

Cunninghamella bainieri

[286]

Mucor bainieri;

[291, 292, 287, 290]

M AN U

Paecilomyces sp. SC0924

5-Hydroxy-zearalenone (162)

5-Hydroxy-zearalenone (8'-hydroxyzearalenone, F-5-3) (162a)

TE D

5-Epi-hydroxy-zearalenone (8'-Epi-hydroxy-zearalenon, F-5-4) (162b)

10-α-Hydroxy-zearalenone (164a)

EP

10-β -Hydroxy-zearalenone (164b) Zearalenol (ZOL) (165)

α-Zearalenol (165a)

AC C

β-Zearalenol (165b)

β-Zearalanol (also named taleranol) (167b)

Cis-β-zearalenol (169b) 14, 16-Dimethoxyzearalenone (2,4-dimethoxyzearalenone) (170) 16-Methoxyzearalenone

(2-methoxyzearalenone) (171)

Zearalenone-14-β- D-glucopyranoside (172)

Rhizopus sp.; Thamnidium elegans

10

ACCEPTED MANUSCRIPT[Benzenediol lactones] Fusarium graminearum

[288]

5'-Hydroxyzearalenol (174)

Fusarium sp.

[40, 288]

8'-Hydroxyzearalanone (175)

Penicillium sp.

[300]

2'-Hydroxyzearalanol (176)

Penicillium sp.

[300]

Zearalenone-11, 12-oxide (177)

Fusarium graminearum

[301]

Zearalenone-11, 12-dihydrodiol (178)

Fusarium graminearum

[301]

10-Oxo-zearalenone (10-keto-ZEA) (179)

Fusarium graminearum

[301]

5'-Hydroxyzearalenone (180)

Fusarium sp.

[40]

Trans-7', 8'-dehydrozearalenol (181)

Paecilomyces sp.

Radicicol A (182)

Cochliobolus lunatus

(also named 89-250904-F1)

Hamigera avellanea

Ro 09-2210 (183)

Curvularia sp.

Queenslandon (184)

Chrysosporium queenslandicum

Cryptosporiopsin A (185)

Cryptosporiopsis sp.

[309]

Y5-02-B (186)

Unknown

[310]

Y5-02-C (187)

Unknown

Apralactone A (188)

Curvularia sp.

[128]

Menisporopsis theobromae

[311]

2. Overview of BDLs

[220] [233] [142] [306] [307]

[310]

2.1 8-Membered benzenediol lactones (Benzenediol octalactones)

TE D

Fig. 3. The structures of 8-membered benzenediol lactones (benzenediol octalactones)

To date, coryoctalactones A–E (1-5) (Table 1 and Fig. 3), isolated from Corynespora cassiicola

EP

(JCM 23.3), an endophyte of the mangrove plant Laguncularia racemosa (Combretacaeae) from the island of Hainan, China are the only members of benzenediol octalactones [27]. The relative stereochemistry of coryoctalactones were determined on the basis of one and two-dimensional NMR spectroscopy as well as by high resolution mass spectrometry. The absolute configuration of the side

AC C

5 6 7 8 9 10 11 12 13 14 15 16 17 18

SC

M AN U

Menisporopsin A (189)

1 2 3 4

RI PT

Zearalenone-14-sulfate (ZEN-4-sulfate) (173)

chain hydroxyl C in 1–3 and 5 were tentatively assigned as “R” based on biogenetic consideration in comparison with xestodecalactones D–F. Moreover, the comparison of the optical rotation value of 3 with those of 1 and 2 as well as comparison of NMR data of C-9 and H-9 indicated that 2 and 3 share identical configuration at C-9, and 1 and 2 are epimers at C-9. All isolated compounds were evaluated for their antimicrobial, cytotoxic, and antitrypanosomal activities, but no significant results were obtained. 2.2 10-Membered benzenediol lactones (benzenediol decalactones) Fig. 4. The structures of 10-membered benzenediol lactones (benzenediol decalactones)

11

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

To data, only ten examples of 10-membered benzenediol lactones have been reported, from marine or terrestrial fungi, namely, sporostatin (6), xestodecalactones A-F (7-12), sonnerlactones (13-14), relgro (15) (Table 1 and Fig. 4). Among them relgro and sonnerlactones belong to

M AN U

10-membered RALs and the other compounds belong to 10-membered DALs.

The first natural 10-membered benzenediol lactone, sporostatin (6), was isolated from the fermentation filtrate of the fungus, Sporormiella sp. M5032 in 1997 by Kinoshita et al [28]. The structure of 6 was secured by an X-ray diffraction study [28], but the absolute configuration was reported by its first total synthesis in 2009 [29]. Sporostatin has been found to have an inhibitory effect on cyclic adenosine 3', 5'-monophosphate phosphodiesterase (cAMP-PDE) with an IC50 value of 41 µg/mL. Further biological evaluation indicated that 6 was a specific kinase inhibitor in vitro whose IC50

TE D

values were 0.1 µg/mL for EGF receptor kinase, 3 µg/mL for ErbB-2, and 100 µg/mL or greater for other kinases, including the platelet derived growth factor (PDGF) receptor, ν-src and protein kinase. Kinetic analyses revealed that inhibition of EGF receptor kinase and cAMP-PDE by sporostatin was noncompetitive either with substrate or with ATP [28, 30]. The xestodecalactones A-C (7-9) were secondary metabolites of a fungus Penicillium cf.

EP

montanense obtained from the marine sponge Xestospongia exigua, and the constitutions, stereostructures including the relative configurations of 8 and 9 were initially established using online HPLC-NMR, ESI-MS/MS, and CD spectra analysis after isolating these compounds [31]. The absolute configurations at C-11 of 7-9 were initially reported as S; however, they were later revised as R until

AC C

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32

SC

1

their total syntheses were accomplished and reported during the period 2004 to 2007 [32-34]. Xestodecalactone B produced 25, 12, 7 mm inhibition zones against Candida albicans at 100, 50, 20 µmol, respectively [31]. It was also noted that xestodecalactone A has been patented for antitumor activity [35]. In 2012, the chemical investigation on the ethyl acetate extract of C. cassiicola from leaf tissues of the mangrove plant L. racemosa collected at Hainan Island in China, led to the isolation of xestodecalactones D–F (10-12) [36]. The structures of 10-12 were determined by analysis of NMR and MS data. The relative configuration of xestodecalactone D was obtained from a careful analysis of the coupling constants observed in the well resolved 1D 1H NMR spectrum, as well as from ROESY correlations. Their absolute configurations were assigned by TDDFT ECD calculations of their solution conformers, proving that they belong to the (11S) series of xestodecalactones, opposite to the (11R) configuration of the known xestodecalactones A–C. In this paper, the cytotoxic, antibacterial, antifungal and antitrypanosomal activity of these compounds were investigated, but no positive results 12

ACCEPTED MANUSCRIPT[Benzenediol lactones] were obtained. In 2010, two new metabolites, (3R, 5R)-sonnerlactone (13) and (3R, 5S)-sonnerlactone (14), were isolated from the mangrove endophytic fungus strain Zh6-B1 obtained from the bark of Sonneratia apetala growing in Zhuhai, Guangdong, China [37]. The absolute configuration of 13 was determined by single-crystal X-ray analysis and spectroscopic data. While 14 was deduced by NOESY analysis and comparing circular dichroism spectroscopy with compound 13. Both sonnerlactones (13, 14) exhibited weak anti-proliferative activity against multidrug-resistant human oral floor carcinoma cell lines

RI PT

(KV/MDR) (42.4% and 41.6%, respectively, at 100 µM). The stereoselective synthesis of 13 and 14 has been accomplished starting from L-aspartic acid [38]. The key steps involved asymmetric allylation, Alder–Rickert reaction and Mitsunobu macrolactonization.

Apart from the natural benzenediol decalactones as described above, a novel compound, relgro (15), was mentioned incidentally as an artificial product for the purpose of active test in 1973 [39]. Up

SC

to 2011, it was isolated and elucidated together with other 14-membered RALs from the seagrass-derived fungus Fusarium sp. PSU-ES73 [40]. Antimicrobial activity test showed that it was inactive against S. aureus ATCC25923, methicillin-resistant S. aureus SK1 and Cryptococcus 14-membered RALs together isolated. 2.3 12-Membered benzenediol lactones

M AN U

neoformans ATCC90113. We presumed that the structure may be related to the rearrangement from the

The number of natural 12-membered benzenediol lactones reported by July of 2014 is 53. This group is distributed in more than 30 species and can be divided into three subgroups: lasiodiplodin macrolides (RAL12), curvularin macrolides (DAL12) and other 12-membered benzenediol lactones. Lasiodiplodin macrolides (RAL12)

Fig. 5. The structures of lasiodiplodin macrolides (RAL12)

AC C

EP

TE D

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

24 25 26

To date, 16 lasiodiplodins, 16-31 (Table 1 and Fig. 5), have been produced by fungi Lasiodiplodia theobromae (syns. Botryodiplodia theobromae and Diplodia gossypina, teleomorph Botryosphaeria 13

ACCEPTED MANUSCRIPT[Benzenediol lactones] rodina), Sarocladium kiliense and Syncephalastrum racemosum, as well as various plants, Euphorbia splendens [41], Arnebia euchroma [42], E. fidjian [43], Annona dioica [44], Dioclea violacea [45], Durio zibethinus [46], Pholidota yunnanensis [47], Cibotium barometz [48], Osbeckia opipara [49], Ficus nervosa[50], Cyphostemma greveana [51], Ampelopsis japonica [52], Macroptilium lathyroides [53], and Caesalpinia minax [54]. Among them, L. theobromae, the common pathogenic fungi isolated in the tropics and subtropics on a variety of host plants, was found to be a remarkable producer of lasiodiplodins. Additionally, it should be noted that several authors assumed that lasiodiplodins may

RI PT

not be plant metabolites but may be from fungal epiphytes or endophytes [55].

(3R)-Lasiodiplodin (16) and (R)-de-O-methyllasiodiplodin (17) were the first members of 12-membered resorcylic acid lactone (RAL12) subclass of the benzenediol lactone (BDL) family to be discovered and were isolated from the cultural filtrate of L. theobromae S22L by Aldridge et al. in 1971 [56]. Later, another seven metabolites, 5-oxo-lasiodiplodin (18), (3R, 5R)-hydroxylasiodiplodin (3R,

5S)-hydroxylasiodiplodin

(20),

(3R,

4S)-4-hydroxylasiodiplodin

(21),

(3R,

SC

(19),

5R)-5-hydroxyl-de-O-methyllasiodiplodin (22), (3R, 6R)-6-hydroxyl-de-O-methyllasiodiplodin (23) and (3R, 6S)-6-hydroxylasiodiplodin (24) were also isolated from the fermentation broth and mycelium of L. theobromae IFO 31059 [57-59]. The relative configuration of 16 was assigned by single-crystal

M AN U

X-ray, whereas the absolute stereochemistry at C-3, initially deduced to be S from chemical reduction to afford phenyl-2-nonanol, was later revised to R on the basis of the synthetic studies [41, 56, 60-63]. The structures including absolute configurations, of the other compounds from L. theobromae were determined by means of spectroscopic analyses, mainly 1D and 2D NMR, the modified Mosher’s method, and chemical derivation. Especially, the S-configuration of C-3 in compounds 18-20 deduced from the structure of 16 based on the comparisons of spectral data in the previous report [57] also needs to be revised to R. Compounds 18-24 were found be active in potato micro-tuber induction at a

TE D

concentration of 10-3-10-4 M and the activity of 21 was stronger than that of the others [57-59]. We presumed that the hydroxyl was essential for potato micro-tuber inducing activity, and S-configuration has higher activity. The biosynthetic pathways of 16 and 20 in L. theobromae have been studied by Kashima with 13C-labeled acetate tracer experiments [64, 65], and the biosynthetic gene clusters was first identified by Xu et al. [66].

EP

In 2006, two new 12-membered resorcylic acid lactones, namely 6-oxo-de-O-methyllasiodiplodin (25) and E-9-etheno-lasiodiplodin (26), together with 16, 17 and 22, were isolated from the mycelium extracts of an unidentified endophytic fungus (No. ZZF36) obtained from a brown alga (Sargassum sp.) collected from Zhanjiang sea area, China [67]. The structure of 25 was confirmed by X-ray

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

crystallographic analysis. In this paper, (3R)-lasiodiplodin (16), de-O-methyllasiodiplodin (17) showed in vitro antimicrobial active against Staphylococcus aureus, Bacillus subtilis, and F. oxysporum. The bioassay result implied that C-13 and C-15 hydroxyls of lasiodiplodins probably may enhance their antibiotic activities. In 2009, one new lasiodiplodin derivative, named botryosphaeriodiplodin (27) together with 16, 19 and 20, were isolated from an endophytic fungus B. rhodina PSU-M114, which was obtained from the leaves of Garcinia mangostana collected in Suratthani Province, Thailand [68]. The absolute configuration at C-3 in 27 was assumed to analogous to that of 16, while the C-7 stereochemistry was not performed as 27 was obtained in low quantity. In this paper, (3R)-lasiodiplodin (16) exhibited antibacterial activity against S. aureus and methicillin-resistant S. aureus with the respective MIC values of 64 and 128 mg/mL. In 2014, two further lasiodiplodins, (3R, 4R)-4-hydroxy-de-O-methyl-lasiodiplodin (28) and (E)-9-etheno-de-O-methyl-lasiodiplodin (29), together with 17, 23 and 25, from a cytotoxic extract obtained from a culture of L. theobromae, an 14

ACCEPTED MANUSCRIPT[Benzenediol lactones] endophyte from the root tissues of Mapania kurzii (Cyperaceae) from the Malaysian rain forest, were characterized on microgram scale (capillary NMR probe) [69]. The structures including absolute configurations, of the other compounds were determined by means of spectroscopic analyses by analogy with the data for the known lasiodiplodins. The cultural extract was found to biologically active against the P388 murine leukaemia cell line (>80% growth inhibition) [69]. Interestingly, lasiodiplodins isolated from the plant have been previously reported to have antileukemic activities [41].

RI PT

The crude EtOAc extract of the strain S. racemosum from soil in tropical and subtropical regions was found to show cytotoxicity against cholangiocarcinoma cell line KKU-M156 with an IC50 value of 18.02 µg/mL. Bioassay-guided fractionation of this fungal extract led to the isolation of five lasiodiplodins,

(3R,

5S)-5-hydroxy-de-O-methyllasiodiplodin

de-O-methyllasiodiplodin

(17)

(25)

(3R)-lasiodiplodin

5R)-5-hydroxy-de-O-methyllasiodiplodin [70].

Compound

30

showed

(16),

(22)

cytotoxicity

and

against

SC

6-oxo-de-O-methyllasiodiplodin

(3R,

(30),

cholangiocarcinoma, KKU-M139, KKU-M156, and KKU-M213 cell lines with IC50 values in the range of 14.30-19.04 µg/mL, while 17 showed cytotoxicity against KB, BC1, and NCI-H187 cell lines with IC50 values of 12.67, 9.65, and 11.07 µg/mL, respectively. In addition, compound 16 showed

M AN U

cytotoxicity against KB, BC1 and NCI-H187 cell lines with IC50 values of 20.77, 25.89, and 20.89 µg/mL, respectively. The first stereo selective total synthesis of 30 and 22 was finished by Yadav et al in 2010 [71]. In

2013,

Yuan

et

al.

reported

the

isolation

and

structure

elucidation

of

(3S,

6R)-6-hydroxylasiodiplodin (31), (3R)-lasiodiplodin (16), and (3R, 5S)-5-hydroxylasiodiplodin (20) from an endophytic fungus S. kiliense grown in rice medium isolated from the gut of healthy Apriona germari (Hope) collected from the campus of Jiangsu Normal University, Jiangsu Province, China [72].

TE D

The structure of 31 was elucidated by a combination of spectroscopic data interpretation, single-crystal X-ray diffraction analysis, and modified Mosher’s method. The absolute configuration at C-6 was determined by a modified Mosher’s method, based on the analysis of differences in proton chemical shifts between the (R), (+) -and (S), (−)-α-methoxy-α-(trifluoromethyl) phenylacetyl (MTPA) esters of 31. The absolute configuration at C-3 was deduced from the relative configuration of C-3 with C-6 by

EP

the X-ray diffraction analysis. The result indicated S configuration for C-3 in lasiodiplodin analogues also exist in nature.

(3R)-Lasiodiplodin (16) and (R)-de-O-methyllasiodiplodin (17) are the most widely studied metabolites in the lasiodiplodin series. (3R)-Lasiodiplodin (16), isolated from plant C. greveana, was

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

found to be active with IC50=0.045 µg/mL against the A2783 ovarian cancer cell line [51]. In 2007, 16 was revealed to be phytotoxic by blocking the electron transport chain in thylakoids at multiple targets that were different from those of current herbicides [73]. De-O-methyllasiodiplodin (17) obtained in the roots of A. euchroma (a traditional Chinese medicine), was found to be responsible, at least in part, for the pharmacological properties of the plant extracts as the result of its efficient inhibition of prostaglandin biosynthesis [55, 63]. In 2010, 17 was reported to be a potent inhibitor of pancreatic lipase (PL) (IC50=4.73 µmol/L), an enzyme that plays a key role in the efficient digestion of triglycerides and that was a target for treating obesity [74]. In 2011, 17 was found to be a novel, natural, non-steroidal mineralocorticoid receptor (MR) antagonist (IC50= 8.93 µmol/L) and an efficient therapeutic target for the treatment of hypertension and other cardiovascular diseases [75]. In this paper, the authors assumed that the acetylation at phenolic hydroxyl groups in analogs of 17 can increase the antagonistic effect against MR and the ring size of the lactone was also very crucial for its activity [75]. 15

ACCEPTED MANUSCRIPT[Benzenediol lactones] It should be noted that this is the first macrolide compound found to behave as MR antagonist to date, different from the other nonsteroidal MR antagonists. Curvularin macrolides (DAL12) Fig. 6. The structures of curvularin macrolides (DAL12)

AC C

EP

TE D

M AN U

SC

RI PT

1 2 3 4

5 6 7

This group including 30 compounds, 32-60 (Table 1 and Fig. 6), is a subclass of DALs with a 12-membered ring system. These compounds almost apparently stem from varied oxidation levels at 16

ACCEPTED MANUSCRIPT[Benzenediol lactones] the C-11 and C-12 position. They are produced by some fungal species mainly from different ascomycete species including Alternaria, Aspergillus, Curvularia, and Penicillium spp. Curvularin (32), the best known member of the curvularin family, was originally isolated as an antifungal agent from the culture filtrate of a species of Curvularia by Musgrave in 1956 [76] and later reisolated from a number of phytopathogenic fungal genera such as Alternaria [77, 78], Ascochytula [79], Beauveria [80], Cochliobolus [81], Curvularia [82, 83], Chrysosporium [84], Drechslera [85], Eupenicillium [86], Helminthosporium [87], Penicillium [88-93], Ulocladium [94] etc and from the

RI PT

root extract of Patrinia scabra plant [95, 96]. The structure of 32 was established by spectroscopic, X-ray and synthetic studies [97-103] and it was usually synthesized based on the Alder–Rickert strategy [100, 102]. Another well known curvularin-type compound, α, β-dehydrocurvularin (trans-dehydrocurvularin) (33), which features a trans double bond in C-10 - C-11 usually co-occurs with curvularin in fungi [78, 83-85, 90, 104-112]. Besides, it was also independently found from Cercospora scirpicola [113], Nectria galligena [114], Stemphylium radicinum [115] and several 9F

SC

series marine fungi [116]. Early biosynthetic studies supported that α, β-dehydrocurvularin was a polyketide product, which was excreted from the cells and reduced to curvularin by extracelullar enzymes [107, 117]. Recently, Molnár’s group predicted the biosynthesis of 33 and revealed

M AN U

collaborating hrPKS-nrPKS pairs whose efficient heterologous expression in Saccharomyces cerevisiae provided a convenient access to the DAL12 scaffolds [13].

Both curvularin (32) and α, β-dehydrocurvularin (33), have shown non-specific phytotoxic activity e.g. against Zinnia elegans and Cirsium arvense [106-107, 118], selective antimicrobial activity e.g. against B. subtilis, S. aureus, and Escherichia coli [83, 111], inhibition of TGF-b signaling activities (IC50 = 34.2 and 1.7 µM, respectively) [118], cytotoxic activities against different human tumour cell lines like A549 (IC50 = 13.91±0.15 and 2.10±0.33 µM, respectively), HeLa (IC50 = 25.64

TE D

±0.37 and 21.01±0.19 µM, respectively), MDA-MB-231 (IC50 = 1.3±0.37 and 9.34±0.38 µM, respectively) and MCF-7 (IC50 = 21.89±0.19 and 11.19±0.31 µM, respectively) [84]. Additionally, 32 has been revealed to reduce the expression of the proinflammatory enzyme iNOS in a glucocorticoid-resistant model of rheumatoid arthritis by inhibiting the JAK/Stat signaling pathway [96, 119-120]. It also exhibited 80% acetylcholinesterase (AChE) inhibitory activity [84], and had the

EP

antioxidant activity of H. maydis FBF at a high concentration [87]. While compound 33 showed good superoxide anion scavenging activity with an EC50 value of 16.71 µg/mL [84] and it was also active against COLO 205 with an IC50 of 7.9 µM [84]. Taking together, the IC50 values of 33 were almost always one order of magnitude higher than those of 32 against the corresponding test cell lines except

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

for MDA-MB-231 cell line. However, the presence of a double bond is the only structural difference between two compounds (as one counterpart), which might be one of the important structural property essential for effective interaction in the bioactive profiles. Ten new members of the curvularin family, which include 11-α-hydroxycurvularin (34a),

11-β-hydroxycurvularin (34b), cis-dehydrocurvularin

(36),

11-α-methoxycurvularin 12-oxo-curvularin

(37),

(35a), 11-β-methoxycurvularin

(35b),

11-β-hydroxy-12-oxo-curvularin

(38),

11,12-dihydroxycurvularin (39), 12-hydroxy-10,11- trans- dehydrocurvularin (40) and citreofuran (41) together with two known 32 and 33, were isolated from the mycelium of the hybrid strain ME 005 derived from P. citreoviride 4692 and 6200, upon fermentation on rice [91, 108, 121]. Their structures were established by spectroscopic data. It is worth mentioning that citreofuran (41) is a structurally unique octaketide derivative belonging to the curvularin family containing a furan ring. Furthermore, these authors still pointed that the abosolute configuration of the biologically active compound 17

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1

previously isolated from A. tomato in 1976 was β-hydroxycurvularin (34b) [106]. Later,

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

11-hydroxycurvularin (34) and 11-methoxycurvularin (35) (the mixtures of 11α- and 11β-epimers) were also reported as metabolites of several fungi such as A. cervinus [111] and Penicillium sp. [122] (Trichonaceae) isolated from the rhizosphere of Anicasanthus thurberi and Fallugia paradoxa, respectively. The author assumed 34 and 35 was probably produced by a Michael-type addition of H2O and MeOH to the enone system of 33 during the processing of these microorganisms, respectively [111]. In 2007, the first total synthesis of 35a and 35b was reported by Liang et al., in which the

RI PT

spectral data of originally proposed structure for 11-methoxycurvularin were pointed to 35a [123]. The curvularin derivatives 32-41 were discovered to have considerable cytotoxicities toward a panel of human cancer cell lines (NCI-H460, MCF-7, SF-268, MIA. Pa Ca-2), with IC50 values ranging from 0.90 to 13.30 µM [111, 122]. They acted as inhibitors in blocking the cell division [81] by specifically disordering the microtubule centers [102] and inducing barrel-shaped spindles [90], and inhibited

SC

Hsp90 [16], a promising target for anticancer drug discovery (Janin 2005). The ability likely originated from the conformation of their macrocyclic ring, which mimiced the M-helicity of the colchicin skeleton, a stereochemical feature known to be important for tubulin binding of these agents [81, 101]. Compounds, 34, 37 and 38, had nematicidal activities against P. penetrans of 35%, 80% and 23% at a

M AN U

respective concentration of 300mg/L [110, 124]. Moreover, 11-α-methoxycurvularin (35a) was independently identified from the crude extract of C. oryzae MTCC 2605 by silica gel column chromatography in 2009 [125]. In this paper, it showed strong antibacterial activities against gram-positive bacteria e.g. S. aureus, B. sphericus with a mean MIC value of 100 µg/mL and gram-negative bacteria e.g. Pseudomonas aeruginosa, P. oleovorans with inhibition zones between 12 to 16 mm, as well as moderate antifungal activity. Against Spodoptera litura 4th instar larvae LD50 was determined to be 205.59 µg/mL [125].

TE D

Recently, the chemical investigation of the endophytic fungus Penicillium sp. obtained from stem tissues of Limonium tubiflorum (from L. tubiflorum growing in Egypt) resulted in the isolation of two new curvularins, named penilactone (42) and 10,11-epoxycurvularin (43), along with four known curvularin-type metabolites (33, 35a, 35b, 36) [126]. Penilactone (42) is characterized by a 4-chromanone ring system formed by connecting the oxygenated aromatic carbon C-7 and CH-11. It

EP

was noteworthy that 42 incorporated a 4-chromanone ring system was rather rare among macrolides. The configuration of 42 and 43 at C-15 was assumed by comparison the of measured [α]D values with those of similarly curvularin-type metabolites [126]. In this paper, compounds (35a and 35b) showed

AC C

pronounced antitrypanosomal activities with MIC values of 4.96 and 9.68 µM, respectively. Compounds (35a, 35b, 36) selectively inhibited the growth of human T cell leukemia cells (Jurkat) (IC50 = 3.9, 2.3, and 5.5 µM, respectively), and human histiocytic lymphoma cells (U937) (IC50 = 1.8, 4.6, 2.5, and 7.6 µM, respectively), and reduced TNFα-induced NF-κB activation as expressed by their IC50 values of 4.7, 10.1, 5.6, and 1.6 µM, respectively [126]. We presume that the presence of a free methoxy group at C-11 is essential for antitrypanosomal activity, whereas trans isomer is more active than cis one. Other well known curvularin analogues included β, γ-dehydrocurvularin (44) isolated from the culture filtrate and mycelial mats of Aspergillus sp. [110], E-6-chloro-10, 11-dehydrocurvularin (45) from C. spicifer [81], 11-O-acetyldehydrocurvularin (46) from C. scirpicola [113]. 44 showed nematicidal activities against P. penetrans of 35% and 87% at a respective concentration of 300 and 1000 mg/L [110]. In 2005, three new curvularin-derived products were obtained during the course of biotransformation of curvularin (32) with B. bassiana ATCC 7159. They were identified as 18

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35

curvularin-7-O-β-D-glucopyranoside (47),

curvularin-4'-O-methyl-7-O-β-D-glucopyranoside (48),

36 37 38 39 40 41 42 43 44

oxa-Michael addition of the phenolic hydroxyl group to the α, β-unsaturated ketone moiety. The

6-hydroxycurvularin-4'-O-methyl-7-O-β-D-glucopyranoside (49) which resulted from hydroxylation, glucosidation, and methylglucosidation of the substrate, respectively [80]. In 2008, a new curvularin analogue, oxacyclododecindione (50), was isolated from the fermentations of the fungal strain Exserohilum rostratum IBWF99121, as a potent inhibitor of IL-4 signaling by inhibiting the binding of the activated Stat6 transcription factors to the DNA binding site without affecting tyrosine phosphorylation [127]. The structure of 50 was determined by a combination of spectroscopic

RI PT

techniques. Unfortunately, neither the relative nor the absolute configuration could be determined from the existing data. Later, in a screening for fungal compounds with a cell-based reporter assay, 32, 33 and 50 were found to suppresse the TGF-b inducible activation of a Smad 2/3 dependent transcriptional reporter in HepG2 cells in a dose-dependent manner without displaying cytotoxic effects. Among them, 50 was the most highly potent inhibition with IC50 values of 190-217 nM [119]. We assume that free

SC

methyl groups may enhance the activity.

During the chemical investigation of the cytotoxic extract of the marine-derived fungus Curvularia sp. (strain no. 768), isolated from the red alga Acanthophora spicifera, six novel curvularin-type macrolides, (+)- (15R)- 10,11- E-dehydrocurvularin (51), (+)- (15R)- 12-oxocurvularin (+)- (15R)- 12- hydroxy- 10,11- E- dehydrocurvularin (53), (+)- (15R)- 13- hydroxy- 10,11- E-

M AN U

(52),

dehydrocurvularin (54), and C-11 stereocenter epimers of (+)- (15R)- 11- hydroxycurvularin (55 and 56), were obtained together with a 14-membered macrolide. The planar structures of 51-56 were confirmed by interpretation of extensive 1D and 2D NMR experiments (HMBC, HSQC and COSY). Based on their optical rotation and CD spectra, the configuration at C-15 of newly isolated metabolites 51-56 was assigned as R, which was different from the known curvularin-type compounds as described above. The newly identified metabolites 51 and 53-55 were found to exhibit structure-dependent

TE D

cytotoxic properties. Especially, 51 displayed strong cytotoxicity, about 24-fold more cytotoxic potency than its 13-hydroxy derivative 53, against nine tested human cancer cell lines (BXF 1218L (bladder cancer), BXF T24 (bladder cancer), CNXF SF268 (glioblastoma), LXFA 289L (lung adenocarcinoma), MAXF 401NL (mammary cancer), MEXF 462NL (melanoma), MEXF 514L (melanoma), OVXF 899L (ovarian cancer), and PRXF PC3M (prostate cancer)) in a

EP

concentration-dependent manner (IC50 = 1.25 µM), suggesting that the stereochemistry at C-15 chiral center of the curvularin-type macrolides is not crucial for their cytotoxic activity. In contrast, variations of the oxidation levels around the macrocycle seem to influence the cytotoxic activity of the

AC C

(+)-(15R)-curvularin-type metabolites [128]. Curvulone A (57), another novel (15R)-curvularin-type metabolites has been isolated from a Curvularia sp. associated with a marine red alga Gracilaria folifera, with two known 15R series compounds (51 and 54). Structurally, 57 features a unique benzofuranone ring linked to 12-membered macrolactone, which can be formed by an intramolecular relative configurations of two chiral centres were determined as 10S*, 15R* by X-ray diffraction analysis, and the absolute configuration of 57 was accomplished independently by the solid state TD-DFT ECD method and by measuring the anomalous dispersion effect. The metabolites (51, 54 and 57) had pronounced antibacterial activities against the B. megaterium, antifungal activities against Microbotryum violaceum and Septoria tritici, and were antialgal against Chlorella fusca. Particularly, 57 showed the strong activity against the Gram positive bacterium Bacillus strain 6540 inhibiting 92% of the bacterial growth [129]. More recently, in order to find new bioactive secondary metabolites from marine-derived fungi, 19

ACCEPTED MANUSCRIPT[Benzenediol lactones] three rare curvularin derivatives, sumalarins A–C (58-60), along with three known metabolites (32, 33 and 47), were identified in the cytotoxic extract of P. sumatrense MA-92, a fungus obtained from the rhizosphere of the mangrove Lumnitzera racemosa. 58-60 were the first examples of curvularin derivatives possessing sulfur substitution. Their structures were established by a detailed interpretation of NMR and MS data, including determining the crystal X-ray structure of 58. The authors assumed that 60 was likely produced via Michael addition of the cysteine metabolite 3-mercaptolactate to the double bond of dehydrocurvularin and thus was reminiscent of glutathione detoxification of Michael

RI PT

acceptors in mammalian metabolism. Compounds 58 and 59 can be derived from 60 by esterification or esterification and acylation, respectively. The bioassay results showed that the sulfur-containing curvularin derivatives 58-60 exhibited cytotoxic activities against seven tested tumor cell lines (Du145, HeLa, Huh 7, MCF-7, NCI-H460, SGC-7901, and SW1990) with IC50 values ranging from 3.8 to 10 µM, while 32 was inactive. The data implied that the structural feature of sulfur substitution at C-11 or Novel 12-membered resorcylides Fig. 7. The structures of novel 12-membered resorcylides

SC

a double bond at C-10 significantly increased the cytotoxic activities of the curvularin analogues [130].

To date, 7 12-membered benzenediol lactones including trans-resorcylide (61), cis-resorcylide (62), dihydroresorcylide (67), 7-hydorxyresorcylide enantiomers (63 and 64), and their methyl esters (65 and 66) (Table 1 and Fig. 7), have been included in this type. They are produced by fungi Penicillium spp. [131], Pyrenophora teres [132], D. phlei [133] and Acremonium (Sarocladium) zeae

EP

[134]. Structurally, they are related to both curvularinds and lasiodiplodins, but possess a modified carbon skeleton [131]. The skeleton of these compounds contains one more ketone carbonyl than lasiodiplodin’s. Meanwhile, the ketone carbonyl in these compounds is linked to the isolated benzylic CH2 corresponding to C-10 rather than directly to the aromatic ring in curvularins.

AC C

16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

TE D

M AN U

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Trans- (61) and cis-resorcylide (62) were first isolated as new plant growth inhibitors from an

unidentified species of Penicillium by Oyama et al. in 1978 [135]. Later, they were reisolated from another fungus Penicillium SC2193 along with 7-hydorxyresorcylide enantiomers (63 and 64), and their methyl esters (65 and 66) [131]. In 1996, 62 has been isolated from P. roseopurpureum [136]. Besides the genus Penicillium, trans- (61) and cis-resorcylide (62) also was found in P. teres, the cause of net-type net blotch of barley [132], and D. phlei, a pathogenic fungus on timothy grass, along with 63 and 64 [133]. The structures of 61-66 were determined using a combination of spectroscopic techniques with emphasis on NMR spectroscopy. The spectral data indicated that 61 and 62 possess a rather unique structural mark: the cis isomer (62) is characterized by a strong H-bond between the phenol hydroxyl and the lactone carbonyl, while trans isomer (61) lacks that feature [135]. The S-configuration at C-3 of resorcylides was originally assigned on the basis of the chemical degradation of cis-resorcylide [135], and this was later confirmed by total synthesis [137]. In 2008, a new 20

ACCEPTED MANUSCRIPT[Benzenediol lactones]

7-hydroresorcylide (64 and 64) [134]. The configuration of a methyl group in 67 had been originally determined to be the S configuration; however, the optical rotation sign of naturally isolated 67 {[α]25.0 D +15.0 (c 0.33, MeOH)} was opposite to that of the synthetic 67 {[α]24.0 –40.0 (c 0.8, MeOH)}, D indicating that the configuration of naturally isolated 67 most likely needs to further revised [138]. Resorcylides were phytotoxic as demonstrated in leaf puncture wound assays and inhibited

RI PT

seedling root elongation [132-135, 139]. 61 caused necrosis on corn and crabgrass (Digitaria sanguinalis) at 0.06 µg per leaf. At 2 µg/leaf, timothy (Phleum pratense), wild poinsettia, and sunflower (Helianthus annuus) were also very sensitive to 61, while 62 was inactive at 1 µg per leaf. Other saturated resorcylide (63-66) fell midway between the two and retained activity at 0.5 µg per leaf. Compound 61 was a plant growth inhibitor and was >10 times more effective than 62 at inhibiting

SC

seedling root elongation [135]. Additionally, trans-resorcylide (61) exhibited cytotoxicity against a panel of tumor cell lines, anti-microbial activity against Pyricularia oryzae, and was an inhibitor of 15-hydroxyprostaglandin dehydrogenase, whereas 62 does not inhibit that enzyme [140]. Cis-R-(−)-resorcylide (62) specifically inhibits blood coagulation factor XIIIa and may be

M AN U

advantageous to enhance fibrinolysis and resolve blood clots [136, 141]. Apparently, the configuration of double bond has a strong impact on bioactivities. 2.4 13-Membered benzenediol lactones

Fig. 8. The structures of 13-membered benzenediol lactones

TE D

21 22 23 24 25 26 27 28 29

of 154 isolates), a protective endophytic fungus of maize, along with the two diastereomers of

Natural occurring 13-membered 1, 3-dihydroxybenzene macrolides are very rare. Only two compounds without definite names were reported as the derivatives of 14-Membered RALs so far. The

EP

6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

metabolite, dihydroresorcylide (67), was also revealed to have wide spread occurrence in A. zeae (105

authors thinked that these two 13-membered macrolides (68, 69) (Table 1 and Fig. 8) isolated from the marine mangrove fungus Aigialus parvus BCC 5311 were possibly produced from air oxidation of aigialomycin A and hypothemycin, respectively. Possible mechanism for the production of 68 is

AC C

1 2 3 4 5

proposed in Scheme 1. Rearrangement from a 14-membered to 13- membered macrolides could be explained by carbon migration of hemiacetal [142]. Scheme 1. Possible mechanisms for the production of 13-Membered benzenediol lactones

21

2.5 14-Membered benzenediol lactones

14-Membered benzenediol lactones constituted the largest group of BDLs including more than 100 substances, the majority of which belonged to RALs family. These included aigialomycins, caryospomycins,

cochliomycins,

hamigeromycins,

hypothemycins,

monocillins,

monordens,

neocosmosins, paecilomycins, pochonins, zeaenols and zearalenones etc, which were mainly obtained from fungal species of genera Aigialus, Caryospora, Cochliobolus, Hamigera, Hypomyces,

TE D

Monocillium, Humicola, Neocosmospora, Paecilomyces, Pochonia, Drechslera and Fusarium etc. In addition, only three natural 14-membered benzenediol lactones resembled the DALs. Aigialomycins

EP

Fig. 9. The structures of aigialomycins

AC C

1 2 3 4 5 6 7 8 9 10 11 12

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

22

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

1

TE D

Compounds of this type include 9 RALs, 70-78 (Table 1 and Fig. 9). They were mainly reported as the secondary metabolites of the genus Aigialus.

In the course of screening for novel bioactive compounds from microbial sources, five resorcylic macrolides, aigialomycins A-E (70-74), were isolated from the mangrove fungus, A. parvus BCC 5311, along with previously reported hypothemycin (95) (see in Fig. 13) in 2002 [143]. The structures and

EP

absolute configuration of 70-74 were elucidated by chemical conversions, detailed analysis of the NMR spectroscopic and mass data in conjunction with X-ray data obtained on hypothemycin [143]. Several years later, another six new nonaketide metabolites including aigialomycin F and G (75 and 76), which can be looked as distinct aigialomycin derivatives, were isolated from the same marine

AC C

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

mangrove fungus by the same group [144]. The stereochemistry of 75 and 76 was addressed by conversion to acetonide derivatives. The result that the same compound could be prepared by hydrogenation of 75 and reduction of 76 with NaBH4 revealed that 75 and 76 possessed the same relative stereochemistry. On the basis of these experimental data and by biogenetic correlation to 71, the absolute configurations of 75 and 76 were deduced to be 1'R, 2'S, 4'S, 5'S, 6' S, 10'S and 1'R, 2'S, 4'S, 5'S, 10'S, respectively. Compounds (70-75) displayed weak activity against Plasmodium falciparum K1, cytotoxicities against two cancer cells (KB, BC-1) and Vero cells (African green monkey kidney fibroblast), of which aigialomycin D (73) was the most potent (IC50 were 6.6, 3.0, 18 and 1.8 µg/mL, respectively) [143]. Additionally, 73 also exhibited kinase inhibitory activities [145, 146] (IC50: 6 µM for CDK1/5, 21 µM for CDK2, and 14 µM for GSK) although it did’ t belong to those resorcylic macrolides containing a cis-enone moiety, which are well known as one of the more promising groups of natural products that have recently emerged as new lead structures for kinase 23

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3 4

inhibition. The result suggested that different RAL scaffolds might inhibit different kinases via various

5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21

The compounds of this class have been found from other fungi. 1', 2'-Epoxyaigialomycin D (77)

resulting in eight total syntheses [145, 147-153]. In addition, 73 has recently been shown to bind to Hsp90, but does not function as an indiscriminate ATP antagonist [154].

was first isolated from H. subiculosus in 2006 with other hypothemycin analogues [155]. The structure was established by analysis of NMR and MS data, and the relative configurations were assigned by

RI PT

NMR analyses (1H, 13C, HSQC, HMBC, and TOCSY) and performing X-ray crystallographic analysis. The analysis revealed that the structure of 77 was very similar to that of 73, which just bore an epoxide functionality between C-1' and C-2' instead of the trans-olefinic bond [155]. Aigialomycin D (73), isolated from the cultures of Fusarium sp. LN-10, an endophytic fungus originated from the leaves of Melia azedarach in 2011, was revealed to have significant toxicity toward brine shrimp larvae (76.7%

SC

at a concentration of 10 µg/mL) [156]. In 2012, compounds 71-73, 75 and 77, obtained from the mycelial solid culture of Paecilomyces sp. SC0924, were found to have weak activities against Peronophythora litchii in an assessment of antifungal activity by the well plate diffusion method [157].

M AN U

Recently, deoxy-aigialomycin C (78) was first reported as a natural product isolated from the sea anemone-derived fungus C. lunatus. Its relative and absolute configurations were confirmed using NOE experiments of its acetonide derivative and the CD exciton chirality method with its pdimethylaminobenzoyl derivative. It exhibited potent antifouling activity at nontoxic concentrations with the EC50 value of 22.5 µg/mL [158]. Previously, 78 was obtained as a new aigialomycin analogue along with epi-Aigialomycin D through the syntheses studies of aigialomycin D [154]. Caryospomycins

Fig. 10. The structures of caryospomycins

EP

TE D

22 23

mechanisms. Consequently, 73 has attracted considerable interest from the synthetic community,

24

Three compounds of this type, 79−81 (Table 1 and Fig. 10), possessed a rare structure in which a 1,2-dimethoxy-4-hydroxybenzene ring fused into the 14-membered macrolides. In 2007, our group

AC C

25 26 27 28 29 30 31 32 33 34 35 36 37

reported the isolation of caryospomycins A-C (79-81), as a result of the investigation of the metabolites with activities against plant parasite nematodes from the fresh-water fungus C. callicarpa YMF1.01026, which was initially isolated from a submerged woody substrate collected from a freshwater habitat in Yunnan Province, China. The chemical structures of 79 − 81 were determined through NMR spectroscopic analysis. The spectra data combined with 2D NMR experiments indicated 80 was a deacetonided derivative of 79, and 81 was an analogue of 80 but with a different oxidation state at C-6'. In this report, these compounds were demonstrated to be active in an in vitro microtiter plate assay and exhibited moderate killing activity against the nematode Bursaphelenchus xylophilus with the LC50 values of 103.1, 105.8, and 105.1 ppm, at 36 hr, respectively [159]. Cochliomycins Fig. 11. The structures of cochliomycins 24

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

1

The compounds of this class, 82−87 (Table 1 and Fig. 11), were only obtained from fungus C. lunatus.

Cochliomycins A-B (82-83), two with a rare natural acetonide group, and cochliomycin C (84), one with a 5-chloro-substituted lactone, were isolated from the culture broth of C. lunatus, a fungus

TE D

obtained from the gorgonian Dichotella gemmacea collected in the South China Sea [160]. Their structures were elucidated by extensive spectroscopic analysis including 1D NOE and 2D NOESY experiments and chemical conversions. The only difference between cochliomycin A and B was the position of acetonide group. The absolute configurations of 82 and 83 were further confirmed after treatment of zeaenol (155) with 2, 2-dimethoxypropane in the presence of p-toluenesulfonic acid

EP

(PTSA) resulted in a mixture of cochliomycin A and B. However, cochliomycin C initially published structure 84a, was later revised to 84, which was assigned as 4'S, 5'R, 6'S, 10'S, by the same authors [161]. Interestingly, a transetherification reaction that compound 83 in a solution of CDCl3 slowly can be rearranged to give 82 at room temperature indicated cochliomycin A can arise from cochliomycin B

AC C

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26

via a transetherification reaction in which the ether migrated from O-4' to O-6' [160]. Recently, the reinvestigation of the secondary metabolites from the sea anemone-derived fungus C. lunatus (collected from Weizhou coral reef in the South China Sea) led to the isolation of three new 14-membered resorcylic acid lactones, cochliomycins D−F (85-87), and eight known analogues. Detailed analysis of the 1D and 2D NMR spectra established the same planar structures for 85-87. Their absolute configurations were established by the CD exciton chirality method and TDDFT ECD calculations. The result indicated that these compounds were diastereomers differing from each other by the absolute configurations of the 4', 5'-diol chiral centers [158]. In antifouling assays, compound 82, 85 and 87 exhibited potent antifouling activities against the larval settlement of the barnacle B. amphitrite at nontoxic concentrations, with EC50 values of 1.2, 17.3 and 6.67 µg/mL, respectively. Structure−activity relationships suggested that the enone and acetonide functionalities and hydroxy configurations may have obvious influence on antifouling activity. Additionally, only 82 was found to 25

ACCEPTED MANUSCRIPT[Benzenediol lactones] have moderate antibacterial activity against S. aureus with an inhibition zone of 11 mm in diameter at a concentration of 50 µg/mL [158]. Very recently, cochliomycin A was firstly synthesized by Nanda’s group from L-(t)-tartaric acid in 6.5% overall yields employing Keck allylation [162] and then was stereoselectively synthesized based on chiron approach by Wang et al. [163]. Cochliomycin B (83) and zeaenol (155) have been synthesized from natural chiral template ι-arabinose in overall yield of 4.8%, by

Takai

olefination,

Suzuki

coupling,

alkoxide-mediated

transesterification,

and

RCM

Hamigeromycins Fig. 12. The structures of hamigeromycins

RI PT

macrocyclization as the crucial steps [164].

EP

This group of 7 compounds 88− − 94 (Table 1 and Fig. 12) features a 1, 2- dimethoxy- 4hydroxybenzene ring and a 14-member macrocyclic lactone ring, which includes a 1'-2' trans double bond, a ketone and one or two hydroxys. These compounds were all isolated from Hamigera avellanea.

AC C

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

TE D

M AN U

SC

1 2 3 4 5 6 7 8 9

Hamigeromycins A (88) and B (89) were first isolated from the soil fungus H. avellanea BCC

17816 together with radicicol (103) [142]. In a subsequent study, the same team got hamigeromycins C–G (90-94) by altering a liquid fermentation medium (PYGM). The structures and the stereochemistry of these compounds were deduced by analyses of the NMR spectroscopic and mass spectrometry data in combination with chemical means. It should be noted that 88, 90, 91 and 92 were stereoisomers differing from each other in the absolute configurations of the 4', 5'-diol moiety; while 93 and 94 were unusual 5'-keto-analogs, and they were 6'-epimers to each other. Besides, 89 were characterized by containing an oxygen bridge [165]. In a cytotoxicity assay, all compounds were inactive against cancer cell lines (KB, MCF-7, and NCI-H187) at a concentration of 50 mg/mL, whilst compounds 88 and 90 showed weak growth inhibition against Vero cells (African green monkey kidney fibroblasts) with IC50 values of 42 and 13 mg/mL, respectively. All compounds were inactive in an antimalarial activity assay against P. falciparum K1 at 10 mg/mL [165]. Recently, several synthestic 26

ACCEPTED MANUSCRIPT[Benzenediol lactones] pathways of 89 were accomplished [166, 167]. Hypothemycins Fig. 13. The structures of hypothemycins

Compounds in this type are some of the highly oxygenated analogues in the group of 14-membered resorcylic acid lactones. There are 8 compounds, 95−102, within this type (Table 1 and Fig. 13), mainly isolated from genus Hypomyces and Phoma.

As the best known compounds in this type, hypothemycin (95) was originally isolated from H. trichothecoides along with 7', 8'-dihydrohypothemycin (96), and later reisolated from the fungal

TE D

fermentations of A. parvus [143], Coriolus Versicolor [168], H. subiculosis [155] and Phoma sp. [169]. The absolute configuration of 95 had been tentatively determined many times by different methods, which included X-ray analysis using anomalous dispersion of oxygen atoms [168], X-ray analysis of an aigialomycin C4-bromobenzoyl derivative chemically correlated to hypothemycin [143], single crystal X-ray analysis in combination with the new solid-state CD/TDDFT methodology [169], and

EP

stereospecific total synthesis of 95 [170-172]. Both 95 and 96 were active against the protozoan, Tetrahymena furgasoni and the plant pathogenic fungi Ustilago maydis and Botrytis allii [173, 174]. In addition, hypothemycin 95 also exhibited strong antimalarial activity in vitro (IC50=2.2µg/mL) [143] and could identify therapeutic targets in Trypanosoma brucei [175]. Additionally, detailed enzymatic

AC C

4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30

M AN U

SC

RI PT

1 2 3

and biochemical studies revealed that 95 inhibited a variety of kinases including MEK1/2 (Kd=17 nM and 38 nM), β-type platelet-derived growth factor receptor (PDGFRβ ) (Kd=900 nM), FMS-like tyrosine kinase-3 (FLT-3) (Kd=90 nM), vascular endothelial growth factor receptor 1 (VEGFR1) (Kd=70

nM)

and

(VEGFR2)

(Kd=10

nM)

at

low

nanomolar

levels,

and

including

extracellular-signal-regulated kinase (ERK) (Kd=8.4 mM) at low micromolar levels [143, 168, 176-177]. It also exhibited potent antitumor efficacy [178-180]. In 2013, Xu et al. [181] made an investigation on the antifungal activity of hypothemycin against P. litchii in vitro and in vivo. They found that 95 inhibited spore germination of P. litchii with the inhibition rate of 90% at 0.39 µg/mL, of 100% at 0.78 µg/mL, caused the ultrastructural modifications of P. litchii, especially the plasma membrane, mitochondria, and vacuoles, reduced decay and suppressed peel browning of postharvest litchi fruit inoculated with P. litchii [181]. Several other hypothemycin analogues were found in fungi as well. 5'-O-Methylhypothemycin (97) 27

ACCEPTED MANUSCRIPT[Benzenediol lactones] was isolated from a Phoma sp. with 95. The structure of 97 was elucidated by means of spectroscopic data analysis [169]. In 1999, L-783,277 (98) and its trans-isomer, L-783,290 (99) were both obtained from a Phoma sp. (ATCC 74403) which was isolated from fruitbody of Helvella acetabulum [182]. 98 found by researchers at Merck was reported to be a potent and irreversible inhibitor of MEK1 with an IC50 at 4 nM, weakly inhibit Lck, and has no activity on Raf, PKA, and PKC to a limited degree. Detailed analysis revealed the inhibition was ATP-competitive associated with the formation of a covalent adduct between the enzyme and the inhibitor. The latter compound (99) showed dramatically

RI PT

reduced inhibitory effect on MEK compared to 98 (IC50=300 nM). Three hypothemycin analogues (100-102) were isolated from the fungal strains H. subiculosus DSM 11931 and DSM 11932. The structures of these compounds were elucidated by spectroscopic methods, in combined with chemical conversion of the C-4 hydroxyl group of 100 to an O-methyl group and X-ray crystallographic analysis on a single crystal of 101. One of the analogues, 4-O-demethylhypothemycin (100) exhibited potent and selective cytotoxic activities against cell lines (COL829, HT29 and SKOV3 with IC50 values of Monordens and monocillins Fig. 14. The structures of monordens and monocillins

SC

0.038, 0.10 and 1.8 µM, respectively) with a BRAF mutation [155].

AC C

EP

TE D

M AN U

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

17 18

This subgroup of 14-membered benzenediol lactones with 20 members, (103-119) (Table 1 and 28

ACCEPTED MANUSCRIPT[Benzenediol lactones] Fig. 14), were always isolated together from Humicola sp. Among them, monordens are structurally characterized by a chlorine substitution on the aromatic ring. Monocillins were first isolated from M. nordinii except compound monocillin II glycoside (115). The first 14-membered resorcylic lactone, monorden (103), (also known as monorden A or radicicol), was originally isolated as a potent tranquilizer from Monosporium bonorden in a soil sample collected in the Belgian Congo in 1953 [183]. After ten years, the same molecule was independently obtained from the culture filtrate of Neonectria radicicolas (anam. Cylindrocarpon destructans) (syn.

RI PT

C. radicicola) and was named radicicol [184]. With the establishment of its three chiral centers by X-ray crystal structure in 1987, it was figured out that the initial structure proposed for monorden was incorrect, thus leading to the common acceptance of radicicol as the name of this molecule [185, 186]. In 1992, Lett and Lampilas reported the first total synthesis of radicicol, which confirmed its absolute configurations in three stereocenters [187-189]. The biosyntheses of radicicol in Chaetomium chiversii

SC

and Pochonia chlamydosporia have been studied [190-192]. Later, radicicol has been found to widely distribute in diverse fungal species, including C. chiversii [193], Colletotrichum graminicola [194], Humicola sp. FO-4910 & FO-2942 [195, 196], Neophaeosphaeria quadriseptata [193, 197], Paecilomyces sp. SC0924 [157], P. chlamydosporia var. catenulatum (syn. Diheterospora

M AN U

chlamydosporias; Verticillium chlamydosporium) [198], P. luteo-aurantium [199] etc. It was also found to have antifungal activities against Neurospora sitophila, Giberella zeae [199] and A. flavus (MIC > 28 mg/mL) [200], antimalarial activity (with in vivo efficacy) [195], the inhibition of various signal transduction products such as p60v-src, p60c-src and p53/56lyn [201-203], inhibition of HIV-1 Tat transactivation (IC50 =0.027 µM) [204], inhibition of human breast cancer [205] and inhibition in vitro of heat shock protein 90 (Hsp90) (IC50=20 nM) [206-208]. Of them, the inhibition of Hsp90 attracted much attention of many researchers. Further study indicated 103 inhibited the ATPase activity

TE D

of Hsp90 via competitive binding to the N-terminal ADP/ATP binding pocket with nanomolar affinity [207]. However, due to the highly sensitive functionalities of 103, including a Michael acceptor and an epoxide, both of which were readily metabolized, it was inctive to Hsp90 in vivo [209, 210]. Monorden B (tetrahydromonorden) (104) was originally reported as a synthetic compound but later was shown to be secondary metabolites of the fungus Humicola sp. FO-2942, which was

EP

originally discovered as a producing fungus of amidepsines, inhibitors of diacylglycerol acyltransferase (DGAT), by UV spectrum-guided purification [196]. Along with 104, monoden and three new analogues, monordens C to E (105-107) were also isolated from the same fungus [196]. Based on spectroscopic evidence mainly by NMR analysis, the structures of compounds (105-107) were

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

elucidated. The result showed that 105 was 5', 6'-dihydromonorden A, that 106 and 104 lacked the epoxide moiety of 105 and 103, instead by transenone [196, 211]. All monordens (103-107) caused the cell cycle arrest at G1 and G2/M phases in Jurkat cells at a higher concentration of 30 µM. But among them, only monordens A and E (113, 117) showed antifungal activities against A. niger with IC50 values of 12 and 70 µM, respectively [196]. Furthermore, monordens D (106) was found to inhibit Herpes Simplex Virus (HSV) and be antiparasitic against parasitic protozoan Eimeria tenella [212]. Monocillins were closely related to monorden in the aspects of biogenetic derivation, structure and bioactivity. Monocillins I~V (108-112) were first isolated as co-metabolites of monorden (103) from the fungus M. nordinii, a destructive mycoparasite of pine stem rusts in North America in 1980. Among them, only monorden (103) and monocillin I (108) were detected to have significant antifungal activities (MICs = 10-25 mg/mL) against P. debaryanum and weak activities against Rhizoctonia solani (MICs = 100-200 mg/mL) [213]. Seven years later, these metabolites were re-isolated with two 29

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

new structurally related compounds, namely nordinone (113) and nordinonediol (114), from the same

43 44

moderate activities in the memory T cell model of HIV-1 latency (EC50 = 9.1 and 24.9 µM,

fungi [214]. The structures of 108-112 were all identified based on 1H and

13

C-NMR techniques and

mass spectra while that of 113 was confirmed by its direct conversion into radicicol [186, 214]. Structurally, monocillins V, III, II and IV (112, 110, 109 and 111) corresponded to 13-unchlorinated monordens B, C, D and E (104-107), respectively. In recent decades, the monocillins were usually encountered from diverse fungal species. Monocillin II and III (109, 110) and a monocillin II glycoside (115) were isolated from P. chlamydosporia var. catenulate [212]. Monocillins I (108) was separately

RI PT

isolated from N. quadriseptata (Basionym: Paraphaeosphaeria quadriseptata) occurring in the rhizosphere of the Christmas cactus Opuntia leptocaulis, and C. chiversii endophytic on Mormon tea Ephedra fasciculate [193, 197]. Monocillin IV (111) was reported from H. fuscoatra [200] and Penicillium sp. [204] together with 103. 111 displayed antimicrobial activity against the A. flavus (MIC = 56 mg/mL) and inhibitory activity on HIV-1 Tat transactivation (IC50 = 5 µM) [204]. Both monocillin

SC

I and II (108, 109) exhibited selective cytotoxicities against the human cancer cell lines MCF-7 and NCI-H460 (IC50=0.98 and 0.62 mM, respectively), SF-268, and MIA Pa Ca-2, MDAMB-231 [197, 205]. In 2009, 108-110 were reported as bioactive secondary metabolites from C. graminicola (Holomorph: Glomerella graminicola), which displayed significant antifungal activities against A.

M AN U

flavus and F. verticillioides in conventional paper disc assay [194]. In the same year, 108-111 were found from P. chlamydosporia TF-0480, and the former three (108-110) and 103 showed WNT-5A expression inhibitory activities with the IC50 values of 1.93, 7.36, 17.62 and 0.19, respectively [216]. It was also noteworthy that monocillin I was well-known for its potent inhibitory activity against Hsp90 as radicicol (103) did [193]. Moreover, monocillin I (108) displayed in vitro activity against the stalkand ear-rot pathogen S. maydis [194]. As expected, the biosyntheses of monocillin I (108) and monocillin II (109) have been proven to use a similar pathway although there are structural variations

TE D

that have been attributed to different oxidation and reduction patterns during the hrPKS step and putative post-PKS enzymatic modifications (e.g. epoxidation and halogenation) [13, 191]. Tichkowsky and Lett reported the total synthesis of 108 and 103 via Miyaura-Suzuki couplings in 2002 [215]. A new monorden analog, monorden analogue-1 (116), along with two known monorden (103) and monocillin IV (111), was isolated from H. fuscoatra NRRL 22980. In this paper, the authors only 13

C-NMR, 2D-NMR, and mass

EP

established the planar structure of 116 after analysis of 1H NMR,

spectral data [200]. In 2009, Shinonaga and co-works reported the reisolation of 116 from P. chlamydosporia var. chlamydosporia and its WNT-5A expression inhibitory activities. Here, the

AC C

relative stereochemistry of 116 was elucidated to be 3α-methyl and 5α-hydroxyl based on 1H-1H coupling constants and NOESY data [216]. More recently, three new radicicol analogues, radicicols B-D (117-119), with four known

compounds radicicol and pochonins B, C, and N, were obtained from a marine-derived fungus H. fuscoatra (UCSC strain no. 108111A) via a bioassay-guided isolation process [217]. The extract of this species was shown to reactivate latent HIV-1 expression in an in vitro model of central memory CD4+ T cells. In this report, the structures of 117-119 were determined on the basis of the molecular formulas and NMR properties of the known compounds. The C-2 stereochemistry of 117-119 was all proposed to be analogous to that of radicicol (103). However, the relative configurations of the secondary alcohols of 118 and 119 were not assigned by NOESY correlation spectroscopy, due to the flexible skeleton lacking α, β, γ, δ-unsaturated ketone functionalities. Compounds 103 and 117 exhibited respectively), while 118 and 119 were inactive, indicating that the epoxide functionality is not required 30

ACCEPTED MANUSCRIPT[Benzenediol lactones] Michael acceptor functionality, something the inactive compounds 118 and 119 do not possess. Neocosmosins

RI PT

Fig. 15. The structures of neocosmosins

In 2013, bioassay-guided fractionation of a fungus Neocosmospora sp. (UM-031509) led to the isolation of three new resorcylic acid lactones, named neocosmosins A (120), B (121), and C (122),

SC

(Fig. 15) along with three known RALs (103, 109 and 111). The structures of new compounds were established on the basis of extensive 1D and 2D-NMR spectroscopic analysis, and mass spectrometric (ESIMS) data in conjunction with X-ray data obtained on 111. In the binding assays in vitro, neocosmosins C showed good binding affinity for the human opioid receptors, which selectively

M AN U

inhibited 54.9% of the specific binding of [3H]-enkephalin to CHO-K1 cell membranes expressing human δ-opioid receptors at the concentration of 10 µM (IC50 value of 14.82 µM). Meanwhile, known compounds 103 and 109 also showed potent binding affinities at the µ-opioid receptor (63.5% and 84.9%, respectively). The authors declared that this was the first report on resorcylic acid lactone derivatives with strong affinities for human opioid receptors [218]. Paecilomycins

TE D

Fig. 16. The structures of paecilomycins

EP

5 6 7 8 9 10 11 12 13 14 15 16 17 18

for activation of HIV-1. Moreover, it was also noted that the active compounds (103, 117) all contain a

AC C

1 2 3 4

31

EP

This group of 13 RALs, 123-135 (Table 1 and Fig. 16), are marked as the secondary metabolites of Paecilomyces sp. They are usually highly oxygenated analogues possessing epoxy, hydroxyl, and other oxygen-containing groups on the macrolide ring.

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

During the investigation on the secondary metabolites of Paecilomyces sp. SC0924, Xu et al.

reported the isolation and identification of 13 new benzenediol lactones, named paecilomycins A-M (123-135), together with several known RALs (71-73, 77, 109, 159, 161, 181) from the mycelial solid culture of this fungus, successively [219-221]. Their structures (123-135) were elucidated by extensive NMR analysis and chemical correlations in conjunction with X-ray analysis of paecilomycins A and J. Furthermore, the absolute configurations of paecilomycins J-M (132-135) were assigned by Time-Dependent Density-Functional Theory (TDDFT) calculations of their electronic circular dichroism (CD) spectra after their isolation. Paecilomycins E (127) and F (128) were reversedly assigned in the original isolation report then revised [222] while 135, initially assigned structure 135a, was also revised to 135 [220, 223]. Among them, paecilomycin C and D (125 and 126) having a dihydroisobenzofuranone ring and a polyhydroxylated linear C-10 side chain, which were believed to be biogenetically derived from common macrocyclic RALs through intramolecular SN2-type 32

ACCEPTED MANUSCRIPT[Benzenediol lactones] nucleophilic substitution, were unusual for RALs [144]. It was also very rare in natural RALs that 124 contained an oxygen bridge between C-1' and C-5' to form a pyran ring and 132-135 each contained an ether linkage between C-2' and C-5' to form a tetrahydrofuran (THF) ring. In the assay against the P. falciparum, paecilomycin E (127) displayed antiplasmodial activity against P. falciparum line 3D7 with IC50 value of 20.0 nM, while paecilomycin A, B, F (123, 124, 128) showed moderate activities with IC50 values of 0.78, 3.8, 1.1 µM, respectively. Additionally, 127 and 128 showed moderate activities against P. falciparum line Dd2 with IC50 values of 8.8, 1.7 µM, respectively. Paecilomycin M (135)

RI PT

exhibited weak antifungal activity against P. litchii [157, 181, 221]. In 2012, the stereoselective total synthesis of 128 and the asymmetric total synthesis of 127 were achieved by Srihari [224] and Nanda group [162], respectively. More recently, paecilomycin F (128) was identified from a sea anemone-derived fungus C. lunatus [158]. However, no activity was reported in this paper. Pochonins Fig. 17. The structures of pochonins

14 15 16 17 18 19 20 21 22 23 24 25

AC C

EP

TE D

M AN U

SC

1 2 3 4 5 6 7 8 9 10 11 12 13

These RALs include 16 pochonins, 136-150 (Table 1 and Fig. 17) and 106 (Fig. 14) which were

obtained from P. chlamydosporia var. catenulata strains P0297 and TF-0480 by two different research groups. The entire family of pochonins shared a common structural motif of a 14-membered macrocyclic resorcylic acid lactone core, which was analogous to radicicol (103) except for pochonins F and J (140 and 144) chlorinated at C-5 of the aromatic ring [212, 216]. Differed from the other pochonins, 140 and 144 demonstrated more semblances to the aigialomycin and hypothemycin families of natural products, partly, due to the lack of C-13 chlorination [143, 173]. Interestingly, many pochonins were found to inhibit Hsp90, which may be due to the same pharmacophore as radicicol. Pochonins A-F (136-138, 106, 139-140) were isolated from the strain P. chlamydosporia var. catenulata P0297 [212] while pochonins E-P (139-150) were obtained from P. chlamydosporia var. chlamydosporia TF-0480 [216, 225]. Their structures were elucidated by means of a combination of 33

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1D and 2D spectroscopic techniques. The stereochemistry of the C-7' chlorine in pochonin C (138) had not been firmly established until confirmation by total syntheses [170, 226-227]. While the stereochemistry of the C-6' hydroxyl group in pochonin E and F (139, 140) was first assigned as S-stereochemistry by Shinonaga’s group [225], but later revised R by comparison between the reported proton NMR data of the natural products and the synthetic compounds [228]. An efficient synthesis of ent- pochonin J has been achieved by Martinez-Solorio et al., but does not match the spectroscopic data of pochonin J (144) as initially reported. Further attempts at the structural verification of pochonin J are

RI PT

ongoing [229]. Interestingly, we find pochonin D has the same chemical structure as monorden D (106). In the HSV1 (Herpes Simplex Virus 1) replication assay, all the pochonins with epoxide moieties or allyl alcohol moieties (e.g., 136, 137, 139, 140) exhibited bioactivities in the low µM range. Pochonin D (106) which contains a double bond instead of the epoxide ring, showed only cytostatic effects. The authors presumed that the chlorine substituent was probably not essential for the antiviral activity

SC

because monocillin (110) without halogenation also showed inhibitory activity against HSV1 in the nM range. The antiparasitic activities of 136-140 and 106 were tested against Neospora caninum and Eimeria tenella. The result indicated that all the test compounds was not susceptible to N. caninum, but the partly hydrogenated pochonins containing an epoxide function showed moderate activities against

M AN U

E. tenella. Among these, pochonin A (136) exhibited the best selectivity toward the parasite. In searching for an inhibitor of WNT-5A expression, radicicol (103) showed the strongest WNT-5A inhibitory activity of the tested samples, however, it also showed relatively high cytotoxicity. The inhibitory activities of 141, 145 and 149 were 10-fold weaker than that of 103 while pochonins 142-144, 147-148 and 150 were practically inactive. However, compared with 103, compound 141 and 149 showed no cytotoxicity at concentrations above 100 µM. The data implied that the 4,5-epoxide or 4,5-E-olefin moieties present in 145 and 149, or 139 and 140, respectively, may be necessary for

TE D

radicicol-type compounds designed to inhibit WNT-5A expression, but the chlorine atom at C-13 may decrease the toxicity against dermal papilla cells.

Besides the species P. chlamydosporia, recently, pochonin N (148) was found in the cultures of Fusarium sp. LN-10, an endophytic fungus originated from the leaves of Melia azedarach, and displayed significant growth inhibitory activity against the brine shrimp Artemia salina at a

EP

concentration of 10 µg/mL, with mortality rate of 82.8% [156]. Pochonins B-C and N (137-138, 148) were also identified from the extract of H. fuscoatra (UCSC strain no. 108111A) with 103. Among them, 137 and 138 showed moderate activities in the memory T cell model of HIV-1 latency [217]. Many pochonins have already been determined the inhibitory activities of Hsp90 by severeal

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43

authors because they are structurally related to radicicol [210, 228]. The result found only pochonin A, D, E and F (136, 106, 139, 140) were active to Hsp90. Of pochonins tested, the best strong compound pochonin D had an affinity for Hsp90 (IC50 = 80 nM), which was 4-fold less than radicicol (IC50 = 20 nM). The epoxide derivative, 136 was also found to be a good inhibitor of Hsp90 (IC50 = 90 nM) whereas the 7′, 8′-diol analog was inactive [19]. The data suggested that neither an epoxide ring nor the dienone was critical for Hsp90 binding. The study on action mechanism indicated these compounds had very different conformations in solution which could rationalize their different biological activities. Interestingly, Moulin E. and his team have exemplified how macrocycle functionalization could not only direct kinase selectivity, but also selectivity between target classes [145, 230]. Zeaenols Fig. 18. The structures of zeaenols

34

M AN U

Natural resorcylic lactones of fungal origin from the zeaneols group, 151-158, (Table 1 and Fig. 18) are mainly spread in D. portulacae, C. lunatus, and other unidentified fungus. They exhibit a wide range of biological properties which include oestrogenic, antifungal, phytotoxic and anti-inflammatory activity.

LL-Z1640-1 to 4 (151-154) were first isolated from an unidentified fungus (Lederle Culture 21640) in 1978 [231]. In this paper, the authors found these compounds no particularly interesting activities. They emphasized that these compounds were devoid of anabolic and estrogen-like activity even if they

TE D

were structurally related to zearalenone, a well-known oestrogen hormones. Here, the information about the double bond in 154 was not given. In the same year, Japanese scientists Yuki et al. reported the isolation and structure elucidation of a new metabolite zeaenol (155) from C. lunata grown on potato and (or) Czapek-Dox medium during the chemical studies on aversion-antagonism [232]. In this paper zeaenol was not only confirmed to have the same planar structure to 154 but also the authors

EP

further revealed that the stereochemistry of the double bond at C-7'-C-8' was trans-configuration while the C-6 hydroxyl had α-configuration orientation. Later, some researchers held the opinion that 154 and 155 were the same compound [233]. In 1992, a paper published the fact that LL-Z1640-1 (151) and zeanol (the same structure as that of 155) cooccured in Drechslera, inhibited plant growth [234].

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29

SC

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

Additionally, we found that compounds L-783,278 and L-783,279 reported from strains (MF6280, MF6293) during the course of screening for potent inhibitors of MEK kinase were 152 and 151, respectively, by comparing their physical and spectroscopic data [235]. Recently, total syntheses of 155 were accomplished by Jana and Nanda [162], Gao et al. [164] and Dakas et al. [236], respectively. In 1999, the capacity of several zearalenone derivatives (152, 155, 156 and 157) to inhibit

lipopolysaccharide

(LPS)-induced

cytokine

production

in

phorbol

12-myristate-13-acetate

(PMA)-treated cultured myelomonocytic cells (U937) was revealed. They exhibited IC50 values of 6 nM, 10 mM, 500 nM and 5 µM, respectively [233]. They were obtained from a 50 L fermentation of Xenova fungus 20416, C. lunatu and named 5Z-7-oxo-zeaenol (152), zeaenol (155), 7-oxo-zeaenol (156) as well as 5, 6-dihydo-5-methoxy-7-oxo-zeaenol (157), respectively. The first total synthesis of 152 was published by Tatsuta et al. [170] in 2001 followed by the Lett et al. [171, 172]. Structurally, we found that 5Z-7-oxo-zeaenol had the same structure as 152. Later, 152 was found to be an 35

ACCEPTED MANUSCRIPT[Benzenediol lactones] ATP-competitive potent and selective inhibitors of TAK1 (IC50=8.1 nmol/L) and ERK2 (IC50=8.0 nmol/L) [17, 237]. Since then, interest in these molecules was renewed. Many 152 analogs were surveyed their inhibitory activities against a variety of kinases and their SARs were presented [25]. With regard to the macrocyclic ring of these molecules, the cis-enone was essential for the activities and modification on the enone moiety or adjacent position (C4) may improve metabolic stability by summarizing the SAR of RALs’ inhibitory activity on kinases established so far. It was also noted that the diol moiety at C8-C9 contributed to the bioactivity, although the C9-OH was more important than

RI PT

the C8-OH. Moreover, in the case of the resorcylic ring: addition of N or O-functional group at C13 or C13-C14 fused bicyclic heterocycle can enhance both potency and PK properties [25].

Two new compounds, 15-O-desmethyl-(5Z)-7-oxozeaenol (158) and 7-epi-zeaenol, along with four known RALs (151, 152, 155, 156) were found in filamentous fungal extract of the Mycosynthetix library (MSX 63935; related to Phoma sp.) [15]. Noticeably, the author described 7-epi-zeaenol not a

SC

natural product which was produced by reduction of (5E)-7-oxozeaenol (156) with sodium borohydride. In a series of assays, 152 was the most potent representative of this class of compounds in both cytotoxicity assays against cancer cell lines (MCF-7, H460, SF268, HT-29, MDA-MB-435) and the NF-κB assay [15].

M AN U

In the recent studies, LL-Z1640-1 (151), LL-Z1640-2 (152), zeaenol (155), (5E)-7-oxo-Zeaenol (156), have also been found in the culture broth of C. lunatus (obtained from the gorgonian D. gemmacea) [158]. All the four compounds showed potent antifouling activity against the larval settlement of barnacle B. amphitrite (EC50 values of 5.3, 1.82, 1.2, 18.1 µg/mL, respectively). 151 showed moderate cytotoxicities against A549 and HepG2 tumor cell lines with IC50 values of 44.5 and 98.6 µM, respectively, while LL-Z1640-2 showed cytotoxicity against HCT-116 with an IC50 value of 3.24 µM. Among these compounds, only 152 exhibited promising antifungal activity against Pestalotia

TE D

calabae (MIC = 0.39 µM) and excellent agricultural fungicidal activities against Plasmopara viticola and P. infestans on the whole-plant assay [158, 160, 161]. Additionally, it was also noted that 151 found in Paecilomyces sp. showed weak activity against P. litchii [157]. Zearalenone and its structurally related conpounds

EP

Fig. 19. The structures of zearalenone and its structurally related conpounds

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

36

EP

The compounds (159-181) (Table 1 and Fig. 19) of this group were usually reported as the metabolites of zearalenone in fungi, plants and mammalian systems, as well as the pharmacokinetics. These compounds shared the same carbon backbone close spatial similarity to 17-estradiol [238]. A few review papers have been given on the occurrence, structures and pharmacokinetics of a part of this

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

group [239-241] before. Here we summarize and update the information to date. Compound 159 described as 6-(10-hydroxy-6-oxotrans-1-undecenyl)-ß- resorcyclic acid lactone,

was first isolated from the mycelium of the fungus Gibberella zeae (Syn. F. graminearum) growing as a mould on corn by Stob et al. in 1962 [242], and its chemical structure disclosed by Urry et al, who also coined the name zearalenone (ZEA, ZEN, 159), by combining elements of its predominant occurrence (‘zea’ for maize) with characteristic structural features (‘RAL’ for the resorcylic acid lactone structure, ‘en’ for the olefinic double bond, and ‘one’ for the keto group) [243]. Zearalenone has attracted much attention since it was found. So far, it has been shown to have estrogenicity [244], teratogenicity [245], mutagenicity [246], genotoxicity [247-249], cytotoxicity [248], immunotoxicity [250-253] and hepatonephrotoxicity [254, 255], and to be an enhancer of lipid peroxidation [256] and a cattle-growth stimulant [257, 258] etc. Additionally, it was noted that zearalenone is also one of the most worldwide distributed mycotoxins [259]. The major mycotoxicity of 159 was attributed to its 37

ACCEPTED MANUSCRIPT[Benzenediol lactones] estrogenic effects on genital organs [260] and on reproductive performance [261, 262]. The toxicity to the reproductive system resulted in uterine enlargement, alterations to the reproductive tract, reduced litter size, increased embryolethal resorption, decreased fertility and changes in progesterone and estradiol serum levels in animals [263-267]. Zearalenone (159) is the only member of this family of which the biosynthesis has been genetically characterized, with two reports detailing the identification of the biosynthetic gene cluster from G. zeae [268, 269]. For zearalenone, more than ten synthetic strategies have been designed after its first total synthesis was published in 1967 [270].

RI PT

Following the isolation of zearalenone, Bolliger and Tamm reported the identification of four zearalenone-related fungal metabolites in F. graminearum cultures, i.e. 13-formyl-zearalenone (160), 5, 6-dehydro-zearalenone (161), and the two stereoisomers of 5-hydroxy-zearalenone (162a, 162b) in 1972 [271]. The formation of 5-hydroxy-zearalenone (162) was later confirmed by several groups [272-274], and further metabolites were disclosed, i.e. 5-hydroxy-zearalenol (163) [275], the epimers

SC

of 10-hydroxy-zearalenone (164a, 164b) [276].

In 1985, several closely related compounds were detected in fungal incubations, i.e. zearalenone (159), α-zearalenol (165a), β-zearalenol (165b), zearalanone (166), α-zearalanol (also named zeranol) (167a), ß-zearalanol (also named taleranol) (167b), cis-zearalenone (168), as well as the epimers of

M AN U

cis-zearalanol (169a, 169b) [277]. Early studies on the metabolism of zearalenone have also disclosed the formation of reductive metabolites, in particular stereoisomers of zearalenol (165a, 165b) [249, 278-281]. Among them, α-Zearalenol (167a) exhibited an even higher estrogenic activity than β-zearalanol (taleranol) (167b) and zearalenone (159) suggesting that the orientation of the hydroxyl group at the aliphatic ring may have a pronounced effect on the oestrogenic activity [274, 282].

α-Zearalanol (167a), also named zeranol, was used as a growth promoter in livestock under the product name Ralgro® [283] because of its high oestrogenic activity. In 2003, a study has demonstrated that the

TE D

catechol metabolites of the 167a had a DNA-damaging potential [247]. Recent investigation indicated that 159, 167a and 167b can interfere with various enzymes involved in steroid metabolism [284, 285]. A series of important papers on zearalenone (159) transformation by fungi and actinomycetes have been published [286, 291, 292,]. Two new transformation products, 2, 4-dimethoxyzearalenone (168) and 2-methoxyzearalenone (171), were obtained from 159 by the cultures of the fungus

EP

Cunninghamella bainieri. They were found not bind to rat oestrogen receptor, indicating a loss of oestrogenicity [286]. A considerable amount of 159 was found to be present as zearalenone14-β-D-glucopyranoside (172) [287] or zearalenone -14- O-sulfate (173) in Fusarium cultures [288, 289]. In 1988, Engelhardt et al. found that 159 can be transformed to 172 by maize cell suspension

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

cultures [287]. 172 can also be produced from 159 by several fungi and actinomycetes such as Rhizopus sp. [291], Mucor bainieri, and Thamnidium elegans etc. [292]. Taking together, there were two sites for glycosylation present in 159, but conjugation in position 14 was strongly favored compared to position 16 [293]. Zearalenone-14-sulfate (formerly known as zearalenone-4-sulfate) (173) was first isolated by

Plasencia and Mirocha from rice inoculated with F. graminearum [288]. El-Sharkawy et al. reported the conversion of zearalenone into 173 by many microorganisms [294] and Berthiller et al. identified zearalenone-14-sulfate as phase II metabolite of zearalenone in the model plant Arabidopsis thaliana [295]. The conjugate most likely retained biological properties of the mycotoxin, since the sulfate moiety was easily cleaved under acidic conditions and in rats [294]. Therefore, 173 was included in analytical methods for the determination of free and masked mycotoxins during the last years [296-298]. 38

ACCEPTED MANUSCRIPT[Benzenediol lactones] Recent decades, some new natural zearalenone derivatives were identificated. In 2008, 5'-hydroxyzearalenol (174) was isolated from the culture broth of a marine-derived fungus Fusarium sp. 05ABR26 with three known compounds (159, 162a and 165). The structure and relative stereochemistry of 174 were elucidated on the basis of spectroscopic data and single-crystal X-ray diffraction data. Of the antifungal and antimitotic activities tested, compound 159 displayed potent inhibitory activity against P. oryzae with a MIC value of 6.25 mg/mL, while compound 162a was much less active; however, 174 and 165 showed no obvious activity [299]. In the same year, two new

RI PT

derivatives, 8'-hydroxyzearalanone (175) and 2'-hydroxyzearalanol (176), were isolated from the marine-derived fungus Penicillium sp. along with four known zearalanone derivatives. The structures were elucidated by spectroscopic methods [300]. In 2010, a culture of F. graminearum has been reinvestigated using LC-DAD-MS and, in part, NMR spectroscopy. In addition to confirming most of the previously reported zearalenone metabolites (162a, 162b, 164a, 164b), several novel compounds

SC

were identified. Three of the metabolites were shown to have the structures of zearalenone -11, 12-oxide (177), zearalenone-11, 12-dihydrodiol (178) and 10-keto-zearalenone (179) [301]. As a novel compound, 5'-hydroxyzearalenone (180) and six known ß-resorcylic macrolides (15, 159, 161, 162, 165b and 174) were isolated from the seagrass-derived fungus Fusarium sp. PSU-ES73 in 2011. Their

M AN U

structures were established by analysis of spectral data. Among them only the known compound zearalenone (159) displayed weak antibacterial against S. aureus (MIC = 400 µM) and antifungal activities against C. neoformans (MIC = 50.26 µM) [40]. In 2010, trans-7', 8'-dehydrozearalenol (181), along with zearalenones (159 161), were isolated from the mycelial solid culture of Paecilomyces sp. SC0924. Their structures were established by spectroscopic methods and chemical means [220]. Structure-activity relationships for human estrogenic activity in zearalenone mycotoxins have been reported. The 6' functional group had the largest effect on estrogenicity. In both the zearalenone

TE D

series (trans double bond at 1', 2') and the zearalanone series (saturated double bond at 1' , 2') the order of estrogenicity for 6' substituents was α-OH

NH2≥ =O ≈ β-OH

β-OAc. Meanwhile, increased

flexibility in the macrocyclic lactone ring would be expected to favor tighter binding of a zearalenone analog to the estrogen receptor. Because saturation of the 1', 2' double bond increased flexibility of the macrolide ring, the order of estrogenicity for 1', 2' double bond was dihydro

trans

cis. This

EP

order of activity was observed for 6'-ketone derivatives, however, not observed in the present study for derivatives with other 6'-functional groups (α-OH, β-OH, β-OAc), in which case saturating the 1', 2' trans double bond resulted in a 5- to 8-fold reduction in estrogenicity. In addition, resorcinol OH groups contributed little to the binding of zearalenones to the human estrogen receptor. Placing a

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

hydroxyl group near the macrolide ring methyl group enhanced receptor binding, even though the hydroxyl was one carbon further away than in 17β-estradiol [238]. Other 14-membered resorcinylic acid lactones Fig. 20. The structures of other 14-membered resorcinylic acid lactones

39

RI PT

ACCEPTED MANUSCRIPT[Benzenediol lactones]

1

Apart from the types as described above, there are some special 14-membered RALs, 182-185,

SC

(Table 1 and Fig. 20).

Radicicol A (87-250904-F1) (182) belongs to the subset of RALs bearing a cis-enone. 182 was first reported as a kinase inhibitor by researchers from Sandoz who identified this fungal metabolite during a screen for IL1ß inhibition [302]. The authors found that 182 accelerated the specific mRNA

M AN U

sequences including IL1ß’ s and inhibited tyrosine [233, 302-304]. As a natural product, radicicol A was also isolated from C. lunatus [233] and H. avellanea [142]. In 2005, Marzinzik and co-workers accomplished the first synthesis of 182 [305]. Later, concise syntheses and total syntheses of 182 were also reported by Dakas et al. [304] and Hofmann et al. [24], respectively. In 1998, using high throughput screening of microbial broths (Curvularia sp.), Williams et al identified Ro 09-2210 (183) as a selective inhibitor of MAP kinase kinase 1 (MEK1) with an IC50 at 59 nM. Interestingly, Ro 09-2210 was a highly selective MEK1 inhibitor although it only showed marginal inhibition on other

TE D

kinases such as PKC, PhK, PKA, ZAP-70, and Lck [306].

Queenslandon (184) characterized by a dihydroxyacetone subunit and a highly oxidized benzoic acid, was isolated from the strain C. queenslandicum IFM51121, which was isolated from a soil sample collected in Egypt, and deposited in the collection of Research Center for Pathogenic Fungi and Microbial Toxicoses, Chiba University, Japan. The structure, initially elucidated 184a on the basis of

EP

optical spectroscopy(IR), electrospray (ESI-MS) and high-resolution electron impact mass spectrometry (HREI-MS), one and two dimensional NMR spectroscopy (1H, 13C, DEPT, COSY, HMQC, HMBC, NOESY and TOCSY) [307], was later revised to 184 on the basis of the synthetic studies [308]. It showed distinct activity against fungi including A. nidulans IFM 5369, A. alternata

AC C

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34

IFM 41348, P. variotii IFM 40913, P. chrysogenum IFM 40614, A. flavus IFM 41934, A. fumigatus IFM 41088, A. terreus IFM 40851, and A. niger IFM 5368, but not against bacteria such as Micrococcus luteus IFM2066 and B. subtilis PCI 219 [308]. In 2012, in the course of screening for environmentally benign secondary metabolites regulating

the motility and viability of zoospores of P. viticola, cryptosporiopsin A (185) together with hydroxypropan-2', 3'-diol orsellinate, pentapeptide were isolated from the culture of Cryptosporiopsis sp., an endophytic fungus from leaves and branches of Zanthoxylum leprieurii (Rutaceae) [309]. The structures of 185 were elucidated on the basis of spectroscopic analysis and comparison of its spectrometric data with that of ponchonin D. It was noteworthy that resorcylic acid monomethyl ethers with a non-chelated OH group as in 185 were very rare. Cryptosporiopsin A exhibited motility inhibitory and lytic activities against zoospores of the grapevine downy mildew pathogen P. viticola (10µg/mL), inhibitory activity against mycelial growth of two other peronosporomycete 40

ACCEPTED MANUSCRIPT[Benzenediol lactones] Rhizoctonia solani, and displayed cytotoxic activity to brine shrimp larvae [309]. 14-membered dihydroxyphenylacetic acid lactones

RI PT

Fig. 21. The structures of 14-membered dihydroxyphenylacetic acid lactones

So far only three compounds, 186-188 (Table 1 and Fig. 21), belong to 14-membered DALs. Two naturally derived 14-membered macrolactones, y5-02-B and y5-02-C (186, 187), have appeared in

SC

a Japanese patent as lead structure concerning the inhibition of neuropeptide Y receptor for anti-obesity programs [310], however the more detailed information could not be obtained. In 2008, the chemical investigation of the cytotoxic extract of the marine-derived fungus Curvularia sp. (strain no. 768), isolated from the red alga A. spicifera, yielded the novel macrolide apralactone A (188). The structure was elucidated by interpretation of its spectroscopic data (1D and 2D NMR, CD, MS, UV and IR). The compound was evaluated for the cytotoxic activity towards a panel of up to 36 human tumor cells and showed concentration-dependent cytotoxicities with a mean IC50 value of 9.87 µM [128]. 2.5 Novel benzenediol lactones

Fig. 22. The structures of novel benzenediol lactones

17 18 19 20 21 22 23 24 25 26 27 28

AC C

EP

TE D

5 6 7 8 9 10 11 12 13 14 15 16

phytopathogens, Pythium ultimum, Aphanomyces cochlioides and a basidiomycetous fungus

M AN U

1 2 3 4

Menisporopsin A (189) (Table 1 and Fig. 22), a novel BDL, was isolated from a cell extract of

the seed fungus Menisporopsis theobromae. The structure of 189 was elucidated on the basis of spectroscopic analysis and chemical transformations, with the absolute configuration established by application of the modified Mosher’s method and by using chiral HPLC. Menisporopsin A possessed an unprecedented residue, 2, 4-dihydroxy-6-(2, 4-dihydroxy-n-pentyl) benzoic acid. This compound exhibited antimalarial activity with an IC50 value of 4.0 µg/mL, and antimycobacterial activity with a MIC value of 50 µg/mL [311]. 3. Conclusions The review summarized in the above sections illustrates the isolation, structural determination, biogenetic studies, biological evaluation, and synthesis of eight-, ten-, twelve-, thirteen-, fourteen-membered BDL compounds. These natural products have attracted much attention from 41

ACCEPTED MANUSCRIPT[Benzenediol lactones] biologists and chemists because they were isolated from a variety of fungi and possessed fascinating molecular architectures and attractive biological activities. To date, about 190 BDLs with broad-spectrum bioactivities have been identified from the fungi. The members of BDL family are still expanding and the wide-range bioactivities of BDLs are still explored extensively. For this kind of compounds, general biosynthetic pathways have been defined; however, significant pathway branching for individals is still needed to explore. From a chemical synthesis perspective, while several elegant approaches to important BDLs have been reported, synthesizing such compounds through concise and

RI PT

modular routes which can be used to extend the diversity of this family remains challenging. From a therapeutic perspective, these fungal lactones may serve as promising lead structures for the development of new therapeutics for the treatment of cancer progression and metastasis as well as chronic fibrotic diseases. However, the preclinical development of some promising results from in vitro studies was hindered due to poor stability in blood plasma and liver microsomes, which need an

SC

extensive medicinal chemistry effort to improve its pharmacokinetic properties. Meanwhile, demand for BDLs in medicinal lead identification and in understanding mechanism of activities will lead to continuing the research in this field. Abbreviations Benzenediol lactones

RALs

Resorcylic acid lactones

M AN U

BDLs:

Dihydroxyphenylacetic acid lactones

IUPAC

International union of pure and applied chemistry

iPKSs

Iterative polyketide synthases

nrPKS

Nonreducing iterative polyketide synthases

Hsp90

Heat shock protein 90

NMR

Nuclear magnetic resonace

cAMP-PDE EGF PDGFR ATP HPLC ESIMS CD

Cyclic adenosine 3', 5'-monophosphate phosphodiesterase

50% inhibitory concentration

Extracellular-signal-regulated kinase Beta-type platelet-derived growth factor receptor

Adenosine Triphosphate

EP

IC50

TE D

DALs

AC C

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

High performance liquid chromatography

Electrospray ionization mass spectrometry

Circular dichroism

TDDFT

Time-dependent density functional theory

NOESY

Nuclear overhauser effect spectroscopy

MDR

Multidrug-resistant

NCI

Nonsmall cell lung

MTPA

(−)-α-methoxy-α-(trifluoromethyl) phenylacetyl

PL

Pancreatic lipase

MR

Mineralocorticoid receptor

TGF-b

Transforming growth factor-β

A549

Human lung adenocarcinoma epithelial cell line

HeLa

Human Henrietta Lacks cervical cancer cell line

Ehrlich

Mice Ehrlich ascites carcinoma cell line;

MCF-7

Human breast adenocarcinoma cell line 42

ACCEPTED MANUSCRIPT[Benzenediol lactones] Inducible nitric-oxide synthase

AChE

Acetylcholinesterase

FBF

fermented B. formosana

ED50

50% effective dose, ED50

DNA

Deoxyribonucleic acid

COSY

Correlation spectroscopy

HMBC

Heteronuclear multiple bond coherence

HMQC

Heteronuclear multiple quantum correlation Human prostate cancer

Huh

Human hepatocarcinoma

SGC

Human gastric carcinoma

SW1990

Human pancreatic cancer

PYGM

A Liquid Fermentation Medium

PDGFRβ

β-type platelet-derived growth factor receptor

MEK

Methyl ethyl ketone

FLT-3

FMS-like tyrosine kinase-3

VEGFR

Vascular endothelial growth factor receptor

ERK

Extracellular-signal-regulated kinase

PKA

Proteinkinase A

PKC

Proteinkinase C

MIC

Minimum inhibitory concentration

HIV

Human immunodeficiency virus

DGAT

Diacylglycerol acyltransferase

UV

Ultraviolet spectroscopy

M AN U

TE D

Herpes simplex virus

THF

tetrahydrofuran

SAR

Structure activity relationship

TNF-α

Tumour necrosis factor-alpha

ZAL

AC C

IL

Zearalenone

EP

ZEA ZEN

RNA

10 11

SC

Du1

HSV

1 2 3 4 5 6 7 8 9

RI PT

iNOS

Zearalenone Zearalanol

Interleukin Ribonucleic Acid

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China

(20762015, 31270091) and the Scientific Research Foundation of Southwest University to Dr. Jinyan Dong (swu109046). References

[1] R.H. Cichewicz, Epigenome manipulation as a pathway to new natural product scaffolds and their congeners, Nat. Prod. Rep. 27 (2001) 11-22.

[2] J.M. Gao, New biologically active metabolites from Chinese higher fungi, Curr. Org. Chem. 10 (2006) 849-871. 43

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3

[3] A.A.L. Gunatilaka, Natural products from plant-associated microorganisms: distribution,

4 5

[4] H.W. Zhang, Y.C. Song, R.X. Tan, Biology and chemistry of endophytes, Nat. Prod. Rep. 23

6 7

[5] D. Hoffmeister, N.P. Keller, Natural products of filamentous fungi: enzymes, genes, and their

8 9

[6] B.S. Kim, B.K. Hwang, Microbial fungicides in the control of plant diseases, J. Phytopathol. 155

10 11

[7] A.L. Demain, S. Sanchez, Microbial drug discovery: 80 years of progress, J. Antibiot. (Tokyo) 62

12 13

[8] Q. Zhang, H.Q. Li, S.C. Zong, J.M. Gao, A.L. Zhang, Chemical and bioactive diversities of the

14 15

[9] N. Winssinger, S. Barluenga, Chemistry and biology of resorcylic acid lactones, Chem. Commun.

16 17

[10] R. Thomas, A biosynthetic classification of fungal and streptomycete fused-ring aromatic

18 19 20

[11] Y. Xu, T. Zhou, Z. Zhou, S. Su, S.A. Roberts, W.R. Montfort, J. Zeng, M. Chen, W. Zhang, J.

21 22

[12] M. Metzler, Proposal for a uniform designation of zearalenone and its metabolites, Mycotoxin

23 24 25

[13] Y. Xu, T. Zhou, S. Zhang, L.J. Xuan, J. Zhan, I. Molnár, Thioesterase domains of fungal

26 27

[14] N. Winssinger, J.G. Fontaine, S. Barluenga, Hsp90 inhibition with resorcyclic acid lactones

28 29 30 31

[15] S. Ayers, T.N. Graf, A.F. Adcock, D.J. Kroll, S. Matthew, E.J. Carcache de Blanco, Q. Shen,

32 33 34 35 36

[16] N. Matsushita, S. Akinaga, T. Agatsuma, International Patent (2004) No.WO 2004/024141. [17] J. Ninomiya-Tsuji, T. Kajino, K. Ono, T. Ohtomo, M. Matsumoto, M. Shiina, M. Mihara, M.

37 38 39

[18] J. Xu, C. Jiang, Z. Zhang, W. Ma, Y. Guo, Recent progress regarding the bioactivities,

40 41

[19] T. Taldone, W. Sun, G. Chiosis, Discovery and development of heat shock protein 90 inhibitors,

42 43

[20] J. Liu, Y. Hu, D.L. Waller, J. Wanga, Q. Liu, Natural products as kinase inhibitors, Nat. Prod.

structural, diversity, bioactivity and implications of their occurrence, J. Nat. Prod. 69 (2006) 509-526. (2006) 828-829.

RI PT

regulation, Nat. Prod. Rep. 24 (2007) 393-416. (2007) 641-653. (2009) 5-16.

(2007) 22–36.

M AN U

polyketides, ChemBioChem 2 (2001) 612−627.

SC

genus chaetomium secondary metabolites, Mini-Rev. Med. Chem. 12 (2012) 127-148.

Zhan, I. Molnár, Rational reprogramming of fungal polyketide first-ring cyclization, Proc. Natl. Acad. Sci. U. S. A. 110 (2013) 5398−5403. Res. 27 (2011) 1–3.

TE D

nonreducing polyketide synthasesact as decision gates during combinatorial biosynthesis, J. Am. Chem. Soc. 135 (2013) 10783−10791.

(RALs), Curr. Top. Med. Chem. 9 (2009) 1419-1435.

EP

S.M. Swanson, M.C. Wani, C.J. Pearce, N.H. Oberlies, Resorcylic acid lactones with cytotoxic and NF-kappaB inhibitory activities and their structure-activity relationships, J. Nat. Prod. 74

AC C

(2011) 1126-1131.

Tsuchiya, M. Kunihiro, A resorcylic acid lactone, 5Z-7-oxozeaenol, prevents inflammation by inhibiting the catalytic activity of TAK1 MAPK kinase kinase. J. Biol. Chem. 278 (2003) 18485–18490.

biosynthesis and synthesis of naturally occurring resorcinolic macrolides, Acta. Pharmacol. Sin. 35 (2014) 316–330. Bioorg. Med. Chem. 17 (2009) 2225–2235. Rep. 29 (2012) 392.

44

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[21] J. Mallinson, I. Collins, Macrocycles in new drug discovery, Future Med. Chem. 4 (2012)

3 4

[22] P.W. Piper, S.H. Millson, Spotlight on the microbes that produce heat shock protein 90- targeting

5 6

[23] R.R.A. Kitson, C.J. Moody, Learning from nature: advances in geldanamycin- and radicicol-

7 8

[24] T. Hofmann, K.H. Altmann, Resorcylic acid lactones as new lead structures for kinase inhibition,

1409–1438. antibiotics, Open Biol. 2 (2012) 120-138.

C. R. Chimie 11 (2008) 1318-1335.

RI PT

based inhibitors of Hsp90, J. Org. Chem. 78 (2013) 5117-5141.

9 10

[25] L. Wei, J. Wu , G. Li, N. Shi, Cis-enone resorcylic acid lactones (RALs) as irreversible protein

11 12

[26] J. Patocka, O. Soukup, K. Kuca, Resorcylic acid lactones as protein kinase inhibitors and

13 14

[27] W. Ebrahim, A.H. Aly, V. Wray, P. Proksch, A. Debbab, Unusual octalactones from Corynespora

15 16 17

[28] K. Kinoshita, T. Sasaki, M. Awata, M. Takada, S. Yaginuma, Structure of sporostatin (M5032),

18 19

[29] J.S. Yadav, N.T.K. Uma Gayathri, B.V. Subba Reddy, A.R. Prasad, First stereoselective total

20 21

[30] Y. Murakami, A. Ishii, S. Mizuno, S. Yaginuma, U. Yoshimasa, Sporostatin, a novel and specific

22 23 24 25

[31] R.A. Edrada, M. Heubes, G. Brauers, V. Wray, A. Berg, U. Grafe, M. Wohlfarth, J. Muhlbacher,

26 27

[32] G. Bringmann, G. Lang, M. Michel, M. Heubes, Stereochemical assignment of the fungal

28 29 30

[33] T. Yoshino, F. Ng, S.J. Danishefsky, A total synthesis of xestodecalactone A and proof of its

31 32 33

[34] Q. Liang, J. Zhang, W. Quan, Y. Sun, X. She, X. Pan, The first asymmetric total synthesis and

34 35 36

[35] G. Bringmann, R.A. Edrada, E. Gunther, M. Heubes, P. Proksch, Novel decalactones from

37 38 39 40

[36] W. Ebrahim, A.H. Aly, A. Mándi, F. Totzke, M.H.G. Kubbutat, V. Wray, W.H. Lin, H. Dai, P.

41 42 43

[37] K.K. Li, Y.J. Lu, X.H. Song, Z.G. She, X.W. Wu, L.K. An, C.X. Ye, Y.C. Lin, The metabolites of

kinase inhibitors, Curr. Pharm. Des. 18 (2012) 1186-1198.

SC

naturally occurring toxins, Mini-Rev. Med. Chem. 13 (2013) 000-000

cassiicola, an endophyte of Laguncularia racemosa, Tetrahedron Lett. 54 (2013) 6611–6614. an inhibitor of cyclic adenosine 3', 5'-monophosphate phosphodiesterase, J. Antibiot. 50 (1997)

M AN U

961-964.

synthesis of sporostatin and determination of absolute configuration, Synlett 5 (2009) 790–792. inhibitor of EGF receptor kinase. Anticancer Res. 19 (1999) 4145-4150. K. Sudarsono Schaumann, G. Bringmann, P. Proksch, Online analysis of xestodecalactones

TE D

A-C, novel bioactive metabolites from the fungus Penicillium cf. montanense and their subsequent isolation from the sponge Xestospongia exigua, J. Nat. Prod. 65 (2002) 1598-1604. metabolite xestodecalactone A by total syntheses, Tetrahedron Lett. 45 (2004) 2829-2831.

EP

absolute stereochemistry: interesting observations on dienophilic control with 1, 3-disubstituted nonequivalent allenes, J. Am. Chem. Soc. 128 (2006) 14185-14191.

AC C

determination of absolute configuations of xestodecalactones B and C, J. Org. Chem. 72 (2007) 2694-2697.

associated fungi of marine sponges and their synthetic derivatives as medicaments, New Dev. Mar. Biotechnol. 10 (2004) EP 1474413-A2. Proksch, T. Kurtán, A. Debbab, Decalactone derivatives from Corynespora cassiicola, an endophytic fungus of the mangrove plant Laguncularia racemosa, Eur. J. Org. Chem. (2012) 3476–3484. mangrove endophytic fungus Zh6-B1 from the South China Sea, Bioorg. Med. Chem. Lett. 20 (2010) 3326–3328.

45

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[38] J.S. Yadav, K.S. Shankar, A.S. Reddy, B.V.S. Reddy, The stereoselective total synthesis of (3R,

3 4

[39] S.S.

5 6 7

[40] J. Arunpanichlert, V. Rukachaisirikul, Y. Sukpondma1, S. Phongpaichit, O. Supaphon, J.

8 9

[41] K.H. Leet, N. Hayashi, M. Okano, I.H. Hall, R.Y. Wu, A.T. Mcphailti, Lasiodiplodin, a potent

10 11 12

[42] X.S. Yao, Y. Ebizuka, H. Noguchi, F. Kiuchi, Y. Iitaka, U. Sankawa, H. Seto, Structure of

13 14

[43] R.C. Cambie, A.R. Lai, P.S. Rutledge, P.D. Woodgate, Ent-14[S], 16β,17-trihydroxyatisan-3-one

15 16

[44] P.R.D. Dos Santos, A.A. Morais, R. Braz-Filho, Alkaloids from Annona dioica. J. Braz. Chem.

17 18 19

[45] A.L.B.S. Barreiros, M.L. Barreiros, J.M. David, J.P. David, L.P. de Queiroz, Atividade

20 21

[46] Rudiyansyah, M.J. Garson, Secondary metabolites from the wood bark of Durio zibethinus and

22 23

[47] X. Guo, N. Wang, X. Yao, Chemical constituents of Pholidota yunnanensis, Shenyang Yaoke

24 25

[48] Q. Wu, X. Yang, S. Yang, L. Z, J. Yan, Chemical constituents of Cibotium barometz. Tianran

26 27

[49] H. Wang, Y. Wang, Y. Shi, X. Li, C. L, Chemical constituents in roots of Osbeckia opipara,

28 29

[50] L-W. Chen, M-J. Cheng, C-F. Pengc, I-S. Chen, Secondary metabolites and antimycobacterial

30 31 32

[51] S. Cao, Y. Hou, P. Brodiea, J.S. Millerb, R. Randrianaivoc, E. Rakotobed, V.E. Rasamisond,

33 34

[52] J. Mi, C. Wu, L. Sun, C. Li, F. Xi, W. Chen, Chemical constituents from Ampelopsis japonica,

35 36 37 38

[53] L.M. De Sousa, R.W.D. Gois, T.L.G. Lemos, A.M.C. Arriaga, M. Andrade-Neto, G.M.P.

39 40

[54] H. Liu, G. Ma, J. Yuan, X. Tan, Q. Zheng, Z. Sun, J. Yang, X. Xu, A new cassane-type diterpene

41 42 43 44

[55] X.S. Yao, Y. Ebizuka, H. Noguchi, F. Kiuchi, M. Shibuya, Y. Iitaka, H. Seto, U. Sankawa,

5R)-sonnerlactone and (3R, 5S)-sonnerlactone, Tetrahedron Lett. 53 (2012) 6380–6382. Kornblum,

S.B

Stoopak,

A

new

tablet

disintegrating

agent:

cross-linked

polyvinylpyrrolidone, J. Pharm. Sci. 62 (1973) 43-49. Sakayaroj, A β-resorcylic macrolide from the seagrass-derived fungus Fusarium sp. PSU-ES73,

RI PT

Arch. Pharmacal Res. 34 (2011) 1633-1637. antileukemic macrolide from Euphorbia splendens, Phytochemistry 21 (1982) 1119-1121.

arnebinol, a new ANSA-type monoterpenylbenzenoid with inhibitory effect to prostaglandin

SC

biosynthesis, Tetrahedron Lett. 24 (1983) 2407-2410.

and further constituents from Euphorbia fidjiana, Phytochemistry 30 (1991) 287-292.

M AN U

Soc. 14 (2003) 396-400.

antioxidante de substâncias presentes em Dioclea violacea e Erythroxylum nummularia, Rev. bras. Farmacogn. [online] 13 (2003) 8-11.

Durio kutejensis, J. Nat. Prod. 69 (2006) 1218-1221.

TE D

Daxue Xuebao 23 (2006) 205-208.

Chanwu Yanjiu Yu Kaifa 19 (2007) 240-243, 302. China J. Chin. Mater. Med. 34 (2009) 414-418.

EP

activities from the roots of Ficus nervosa, Chem. Biodiversity 7 (2010) 1814-1821. D.G.I. kingston antiproliferative compounds of Cyphostemma greveana from a Madagascar Dry

AC C

Forest, Chem. Biodiversity 8 (2011) 643-650. Zhongguo Shiyan Fangjixue Zazhi 19 (2013) 86-89. Santiago, R. Braz, J.G.M. Da Costa, F.F.G. Rodrigues, Chemical constituents and evaluation of antibacterial activity of Macroptilium Lathyroides (L.) Urb. (Fabaceae), Quimica. Nova 36 (2013) 1370-1374. and other constituents from Caesalpinia minax, Bull. Korean Chem. Soc. 34 (2013) 1541-1542. Biologically active constituents of Arnebia euchroma: structure of arnebinol, an ansa-type monoterpenylbenzenoid with inhibitory activity on prostaglandin biosynthesis, Chem. Pharm. Bull. 39 (1991) 2956−2961. 46

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[56] D.C. Aldridge, S. Galt, D. Giles, W.B. Turner, Metabolites of Lasiodiplodia theobromae, J.

3 4 5

[57] H. Matsuura, K. Nakamori, E.A. Omer, C. Hatakeyama, T. Yoshihara, A. Ichihara, Three

6 7

[58] Q. Yang, M. Asai, H. Matsuura, T. Yoshihara, Potato micro-tuber inducing hydroxylasiodiplodins

8 9 10

[59] P. Li, K. Takahashi, H. Matsuura, T. Yoshihara, Novel potato micro-tuber-inducing compound,

11 12

[60] G.B. Jones, R.S. Huber, The first catalytic enantioselective synthesis of (R)-lasiodiplodin, Synlett

13 14

[61] F. Bracher, B. Schulte, Enantiodivergent synthesis of both enantiomers of the macrocyclic lactone

15 16

[62] A. Furstner, N. Kindler, Macrocycle formation by ring-closing-metathesis: an efficient synthesis

17 18 19

[63] A. Furstner, G. Seidel, N. Kindler, Macrocycles by ring-closing-metathesis, XI: Syntheses of

20 21

[64] T. Kashima, K. Takahashi, H. Matsuura, K. Nabeta, Biosynthesis of resorcylic acid lactone

22 23 24

[65] T. Kashima, K. Takahashi, H. Matsuura, K. Nabeta, Biosynthesis of resorcylic acid lactone

25 26 27

[66] Y. Xu, T. Zhou, P. Espinosa-Artiles, Y. Tang, J. Zhan, I. Molnár, Insights into the biosynthesis of

28 29 30

[67] R.Y. Yang, C.Y. Li, Y.C. Lin, G.T. Peng, Z.G. She, S.N. Zhou, Lactones from a brown alga

31 32 33

[68] V. Rukachaisirikul, J. Arunpanichlert, Y. Sukpondma, S. Phongpaichit, J. Sakayaroj, Metabolites

34 35 36 37

[69] S. Sultan, L. Sun, J.W. Blunt, A.L.J. Cole, M.H.G. Munro, K. Ramasamy, J-F.F. Weber, Evolving

38 39 40

[70] M. Buayairaksa, S. Kanokmedhakul, K. Kanokmedhakul, P. Moosophon, C. Hahnvajanawong, K.

41 42 43

[71] J.S. Yadav, N. Thrimurtulu, A. Rahman Md, J.S. Reddy, A.R. Prasad, B.V.S. Reddy

Chem. Soc. (C) (1971) 1623-1627. lasiodiplodins from Lasiodiplodia theobromae IFO 31059 in honour of professor G.H., Phytochemistry 49 (1998) 579-584.

RI PT

from Lasiodiplodia theobromae, Phytochemistry 54 (2000) 489-494. (3R, 6S)-6-hydroxylasiodiplodin, from a strain of Lasiodiplodia theobromae, Biosci. Biotechnol. Biochem. 69 (2005) 1610-1612.

SC

05 (1993) 367-368. lasiodiplodin, J. Chem. Soc., Perkin Trans. 1 (1996) 2619-2622.

M AN U

of enantiomerically pure (R)-(+)-lasiodiplodin, Tetrahedron Lett. 37 (1996) 7005-7008. (R)-(+)-lasiodiplodin, zeranol and truncated salicylihalamides, Tetrahedron 55 (1999) 8215-8230.

lasiodiplodin in Lasiodiplodia theobromae, Biosci Biotechnol Biochem 73 (2009) 1118−1122. (5S)-5-hydroxylasiodiplodin in Lasiodiplodia theobromae, Biosci Biotechnol Biochem 73 2522−2524.

TE D

(2009)

12-Membered resorcylic acid lactones from heterologous production in Saccharomyces cerevisiae, ACS Chem. Biol. (2014) dx.doi.org/10.1021/cb500043g.

EP

endophytic fungus (No. ZZF36) from the South China Sea and their antimicrobial activities, Bioorg. Med. Chem. Lett. 16 (2006) 4205-4208

AC C

from the endophytic fungi Botryosphaeria rhodina PSU-M35 and PSU-M114, Tetrahedron 65 (2009) 10590-10595. trends in the dereplication of natural product extracts. 3: further lasiodiplodins from Lasiodiplodia theobromae, an endophyte from Mapania kurzii, Tetrahedron Lett. 55 (2014) 453–455.

Soytong, Cytotoxic lasiodiplodin derivatives from the fungus Syncephalastrum racemosum, Arch. Pharmacal Res. 34 (2011) 2037−2041. Stereoselective, Total synthesis of (3R,5R)-5-hydroxy-de-O-methyllasiodiplodin via the Prins Cyclization, Synthesis-Stuttgart 21 (2010) 3657-3662.

47

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3

[72] W. Yuan, N. Jiang, C. Dong, Z. Wei, H. Wu, C. Chen, Y. Zhao, S. Zhou, M. Zhang, W. Zheng,

4 5 6 7

[73] T.A.M. Veiga, S.C. Silva, A-C. Francisco, E.R. Filho, P.C. Vieira, J.B. Fernandes, M.F.G.F.

8 9 10 11

[74] H-Y. Wang, Y-W. Guo, X-D. Zhang, L. Li, (2010) CN Patent 101676281(A). [75] C-S. Jiang, R. Zhou, J-X. Gong, L-L. Chen, T. Kurtan, X. Shen, Y-W. Guo, Synthesis,

12 13 14

[76] O.C. Musgrave, Curvularin. Part I. isolation and partial characterization of a metabolic product

15 16

[77] D.J. Robeson, G.A. Strobei, α, β-Dehydrocurvularin and curvularin from Alternaria cinerariae, Z.

17 18

[78] D.J. Robeson, G.A. Strobel, R.N. Strange, The identification of a major phytotoxic component

19 20 21

[79] J. Malmstrom, C. Christophersen, J.C. Frisvad, Secondary metabolites characteristic of

22 23

[80] J.X. Zhan, A.A. Leslie Gunatilaka, Microbial transformation of curvularin, J. Nat. Prod. 68 (2005)

24 25

[81] E.L. Ghisalberti, C.Y. Rowland, 6-Chlorodehydrocurvularin, a new metabolite from Cochliobolus

26 27

[82] R.G. Coombe, J.J. Jacobs, T.R. Watson, Constituents of some curvularia species, Aust. J. Chem.

28 29 30

[83] B. Bicalho, R.A.C. Goncalves, A.P.M. Zibordi, G.P. Manfio, A.J. Marsaili, Antimicrobial

31 32 33 34

[84] C.G. Kumar, P. Mongolla, P. Sujitha, J. Joseph, K.S. Babu, G. Suresh, K.V.S. Ramakrishna, U.

35 36 37 38

[85] W.B. Turner, D.C. Aldridge, Fungal merabolites II; Academic: New York. (1983) 167-516. [86] L.W. Xie, Y.C. Ouyang, K. Zou, G.H. Wang, M.J. Chen, H.M. Sun, S.K. Dai, X. Li, Isolation and

39 40 41

[87] J. Dong, L. Zhao, L. Cai, H. Fang, X. Chen, Z. Ding, Antioxidant activities and phenolics of

42 43

[88] H. Raistrick, F.A.H. Rice, 2, 3-Dihydro-3, 6-dihydroxy-2-methyl-4-pyrone and curvularin from

Lasiodiplodin analogues from the endophytic fungus Sarocladium kiliense, Chem. Pharm. Bull. 61 (3) (2013) 363-365. Silva, M.W. Mueller, B. Lotina-Hennsen, Inhibition of photophosphorylation and electron transport chain in thylakoids by lasiodiplodin, a natural product from Botryosphaeria rhodina, J.

RI PT

Agric. Food Chem. 55 (2007) 4217−4221.

modification, and evaluation of (R)-de-O-methyllasiodiplodin and analogs as nonsteroidal antagonists of mineralocorticoid receptor. Bioorg. Med. Chem. a

new

species

of

Curvularia,

J.

Chem.

Soc.

(1956)

4301-4305

DOI:

SC

from

Lett 21 (2011) 1171−1175.

10.1039/JR9560004301.

M AN U

Naturforsch. 36 (1981) 1081-1083.

from Alternaria macrospora as α, β-dehydrocurvularin, J. Nat. Prod. 48 (1985) 139-141. Penicillium citrinum, Penicillium steckii and related species, Phytochemistry 54 (2000) 301-309.

TE D

1271-1273.

spicifer, J. Nat. Prod. 56 (1993) 2175-2177. 21 (1968) 783-788

EP

compounds of fungi vectored by Clusia spp. (Clusiaceae) pollinating bees, Z. Naturforsch. 58 (2003) 746-751.

AC C

Purushotham, G.N. Sastry, A. Kamal, Metabolite profiling and biological activities of bioactive compounds produced by Chrysosporium lobatum strain BK-3 isolated from Kaziranga National Park, Assam, India. SpringerPlus 2 (2013) 122.

difference in anti-Staphylococcus aureus bioactivity of curvularin derivates from fungus Eupenicillium sp. Appl. Biochem. Biotechnol. 159 (2009) 284–293. fermented Bletilla formosana with eight plant pathogen fungi, J. Biosci. Bioeng. (2014) http://dx.doi.org/10.1016/j.jbiosc.2014.03.003 Penicillium gilmanii, J. Chem. Soc. (C) 18 (1971) 3069-3070.

48

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[89] R.F. Vesonder, A. Ciegler, D. Fennell, L.W. Tjarks, Curvularin from Penicillium baradicum

3 4 5

[90] A. Kobayashi, T. Hino, S. Yata, T.J. Itoh, H. Sato, K. Kawazu, Unique spindle poisons.

6 7 8

[91] S. Lai, Y. Shizuri, S. Yamamura, K. Kawai, Y. Terada, H. Furukawa, Novel curvularin-type

9 10

[92] Z. Gu, G. Yang, Y. Cong, Y. Xu, H. Chen, W. Zhang, Studies on chemical constituents from root

11 12

[93] D. Li, T. Zhu, Q. Gu, D. Li, Secondary metabliltes and theirbioactivities of Hibiscus tiliaceus L.

13 14

[94] B. Andersen, M. Hollensted, Metabolite production by different Ulocladium species. Inter J Food

15 16

[95] Z.B. Gu, G.J. Yang, H.Y. Cong, Y.X. Xu, H.S. Chen, W.D. Zhang, Studies on chemical

17 18 19

[96] Y. Yao, M. Hausding, G. Erkel, T. Anke, U. Förstermann, H. Kleinert, Sporogen, S14-95, and

20 21

[97] H. Gerlach, 2-(Trimethylsilyl)αthylester als Carboxylschutzgruppe; Anwendung bei der Synthese

22 23 24

[98] T. Takahashi, H. Ikeda, J. Tsuji, Synthesis of (±)-dimethyl curvularin based on the

25 26

[99] A. Kobayashi, S. Yata, T. Hino, K. Kawazu, A new fungal naphthoquinone which stimulates the

27 28 29

[100] A.J. Birch, N.S. Mani, G.S.R. Subba Rao, Strategies of synthesis based on cyclohexadienes: part

30 31

[101] F. Almassi, E.L. Ghisalberti, B.W. Skelton, A.H. White, Structural study of dehydrocurvularin,

32 33

[102] G.S.R. Subba-Rao, Birch reduction and its application in the total synthesis of natural products,

34 35

[103] J. Krauss, D. Unterreltmeler, C. Neudert, F. Bracher, Short and efficient approach towards

36 37

[104] H.D. Munro, O.C. Musgrave, R. Templeton, Curvularin. Part V. the compound C16H18O5, α,

38 39

[105] A.N. Starratt, G.A. White, Identification of some metabolites of Alternaria cucumerina (E. &

40 41 42

[106] S.B. Hyeon, A. Ozaki, A. Suzuki, S. Tamura, Isolation of α, β-dehydrocurvularin and

43 44

[107] K. Arai, B.J. Rawlings, Y. Yoshizawa, J.C. Vederas, Biosyntheses of antibiotic A26771B by

Baghdadi NRRL 3754, and biological effects, J. Environ Sci. Health B 11 (1976) 271-289. Curvularin and its derivatives, isolated from Penicillium species, Agric. Biol. Chem. 52 (1988) 3119-3123.

4692, Tetrahedron Lett. 30 (1989) 2241-2244. of Patrinia scabra, Chin. Tradit. Herb. drugs 33 (2002) 781-785.

SC

endoohytic fungi, Chin. J. Mar. Grugs 31 (2012) 17-22.

RI PT

metabolites of a hybrid strain ME 0005 derived from Penicillium citreo-viride B.IFO 6200 and

Microbiol 126 (2008) 172–179.

M AN U

constituents from root of Patrinia scabra, Chin. Tradit. Herb. Drugs (Chin.) 33 (2002) 781-782. S-curvularin, three inhibitors of human inducible nitric-oxide synthase expression isolated from fungi, Mol. Pharmacol. 63 (2003) 383-391.

des (−)-(S)-Curvularins Helvetica. Helv. Chim. Acta. 60 (1977) 3039–3044. palladium-catalyzed carbonylation of 3,5-dimethoxybenzyl chloride using a butadiene telomer

TE D

as a building block, Tetrahedron Lett. 21 (1980) 3885-3888.

production of antifungal compounds in alfalfa callus, Agric. Biol. Chem. 51 (1987) 2857-2860. 3, A novel route to macrolide aromatic polyketides, J. Chem. Soc., Perkin Trans. 1 (1990)

EP

1423-1427.

AC C

an inhibitor of microtubule assembly, Aust. J. Chem. 47 (1994) 1193-1197. Pure Appl. Chem. 75 (2003) 1443-1451. macrocyclic lactones based on a sonogashira reaction, Arch. Pharm. Chem. 338 (2005) 605-608. β-dehydrocurvularin, J. Chem. Soc. (1967) 947-948. E.) Ell, Phytochemistry 1 (1968) 1883-1884. β-hydroxycurvularin from Alternaria tomato as sporulation-suppressing factors, Agric. Biol. Chem. 40 (1976) 1663-1664. Penicillium turbatum and dehydrocurvularin by Alternaria cinerariae: comparison of 49

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

stereochemistry of polyketide and fatty acid enoyl thiol ester reductases, J. Am. Chem. Soc. 111

3 4 5

[108] S. Lai, Y. Shizuri, S. Yamamura, K. Kawai, Y. Terada, H. Furukawa, New curvularin-type

6 7 8

[109] M. Vurro, A. Evidente, A. Andolfi, M.C. Zonno, F. Giordano, A. Motta, Brefeldin A and

9 10 11

[110] M. Kusano, K. Nakagami, S. Fujioka, T. Kawano, A. Shimada, Y. Kimura, β,

12 13 14

[111] J. He, E.M.K. Wijeratne, B.P. Bashyal, J. Zhan, C.J. Seliga, M.X. Liu, E.E. Pierson, L.S.

15 16

[112] D. Li, T. Zhu, Q. Gu, D. Li, Secondary metabolites and their bioactivities of Hibiscus Tiliaceus

17 18

[113] G. Assante, R. Locci, L. Camarda, L. Merlini, G. Nasini, Screening of genus Cercospora for

19 20 21

[114] M. Gutierrez, C. Theoduloz, J. Rodriguez, M. Lolas, G. Schmeda-Hirschmann, Bioactive

22 23

[115] J.F. Grove, Metabolic products of Stemphylium radicinum. Part IV. Minor products, J. Chem.

24 25

[116] L.L. Zhao, H.P. Zhang, Isolation of secondary metabolites of 9F series marine fungi and their

26 27 28

[117] Y. Liu, Z. Li, J.C. Vederas, Biosynthetic incorporation of advanced precursors into

29 30 31

[118] K. Rudolph, A. Serwe, G. Erkel, Inhibition of TGF-b signaling by the fungal lactones

32 33 34

[119] S. Elzner, D. Schmidt, D. Schollmeyer, G. Erkel, T. Anke, H. Kleinert, Inhibitors of inducible

35 36 37

[120] N. Schmidt, A. Pautz, J. Art, P. Rauschkolb, M. Jung, G. Erkel, Transcriptional and

38 39 40

[121] S. Lai, Y. Shizuri, S. Yamamura, K. Kawai, Y. Terada, H. Furukawa, New metabolites of two

41 42 43 44

[122] J. Zhan, E.M.K. Wijeratne, C.J. Seliga, J. Zhang, E.E. Pierson, L.S. Pierson, H.D. VanEtten,

(1989) 3391-3399. metabolites from the hybrid strain ME 0005 derived from Penicillium citreo-viride B.IFO 6200 and 4692, Bull. Chem. Soc. 64 (1991) 1043-1050. a,b-dehydrocurvularin, two phytotoxins from Alternaria zinniae, a biocontrol agent of Xanthium

RI PT

occidentale, Plant Science 138 (1998) 67–79.

γ-Dehydrocurvularin and related compounds as nematicides of Pratylenchus penetrans from the fungus Aspergillus sp, Biosci. Biotechnol. Biochem. 67 (2003) 1413-1416.

SC

Pierson, H.D. VanEtten, A.A.L. Gunatilaka, Cytotoxic and other metabolites of Aspergilllus inha biting the rhizosphere of sonoran desert plants, J. Nat. Prod. 67 (2004) 1985-1991.

M AN U

L. endophytic fungi, Chin. J. Mar. Drugs 31 (2012) 17-22.

secondary metabolites, Phytochemistry 16 (1977) 243-247.

metabolites from the fungus Nectria galligena, the main apple canker agent in Chile, J. Agric. Food Chem. 53 (2005) 7701-7708.

TE D

Soc. 12 (1971) 2261-2263

bioactivities against Pyricularia oryzae, Nat. Prod. Res. Dev. 17 (2005) 677-681. dehydrocurvularin, a polyketide phytotoxin from Alternaria cinerariae, Tetrahedron 54 (1998)

EP

15937-15958.

(S)-curvularin, dehydrocurvularin, oxacyclododecindione and galiellalactone, Cytokine 61

AC C

(2013) 285–296.

NO synthase expression: Total synthesis of (S)-Curvularin and its ring homologues, Chem. Med. Chem. 3 (2008) 924–939 post-transcriptional regulation of iNOS expression in human chondrocytes, Biochem. Pharmacol. 79 (2010) 722–732. hybrid strains ME 0004 and 0005 derived from Penicillium citreo-viride B.IFO 6200 and 4692, Chem. Lett. (1990) 589-592. A.A.L. Gunatilaka, A new anthraquinon and cytotoxic curvularins of a Penicillium sp. (Trichonaceae) from the Rhizosphere of Fallugia paradoxa of the Sonoran Desert, J. Antibiot. 57 (2004) 341-344. 50

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[123] Q. Liang, Y. Sun, B. Yu, X. She, X. Pan, First total syntheses and spectral data corrections of

3 4

[124] Y. Kimura, S. Nakahara, S. Fujioka, Aspyrone, a nematicidal compound isolated from the

5 6

[125] S. Busi, P. Peddikotla, S.M. Upadyayula, V. Yenamandra, Secondary metabolites of Curvularia

11-α-Methoxycurvularin and 11-β-Methoxycurvularin, J. Org. Chem. 72 (2007) 9846-9849. fungus, Aspergillus melleus, Biosci. Biotechnol. Biochem. 60 (1996) 1375-1376. oryzae MTCC 2605. Rec. Nat. Prod. 3 (2009) 204-208.

7 8 9 10

[126] A.H. Aly, A. Debbab, C. Clements, Ru.A. Edrada-Ebel, B. Orlikova, M. Diederich, V. Wray,

11 12 13

[127] G. Erkel, H. Belahmer, A. Serwe, T. Anke, H. Kunz, H. Kolshorn, J. Liermann, T. Opatz,

14 15 16

[128] H. Greve, P.J. Schupp, E. Eguereva, S. Kehraus, G. Kelter, A. Maier, H-H. Fiebig, G.M. König,

17 18 19

[129] J. Dai, K. Krohn, U. Flörke, G. Pescitelli, K. Gábor, T. Papp, K.E. Kövér, A.C. Bényei, S.

20 21 22

[130] L-H. Meng, X-M. Li, C-T. Lv, C-S. Li, G-M. Xu, C-G. Huang, B-G. Wang, Sulfur-Containing

23 24

[131] C.J. Barrow, New macrocyclic lactones from a Penicillium species, J. Nat. Prod. 60 (1997)

25 26 27

[132] M. Nukina, T. Sassa, H. Oyama, M. Ikeda, Structures and biological activities of fungal

28 29

[133] Y. Fu, Isolation and characterization of novelphytotoxins from pathogenic fungi, Ph.D. thesis,

30 31 32

[134] S.M. Poling, D.T. Wicklow, K.D. Rogers, J.B. Gloer, Acremonium zeae, a protective endophyte

33 34

[135] H. Oyama, T. Sassa, M. Ikeda, Structures of new plant growth inhibitors, trans- and

35 36 37

[136] R.R. West, T. Martinez, H.R. Franklin, P.D. Bishop, B.R. Rassing, Cis-resorcylide,

38 39 40

[137] E.A. Couladouros, A.P. Mihou, E.A. Bouzas, First total synthesis of trans- and cis-resorcylide:

41 42 43

[138] L. Zhang, W.Q. Ma, L.L. Xu, F. Deng, Y.W. Guo, Efficient total synthesis of

RI PT

W.H. Lin, P. Proksch, NF kappa B inhibitors and antitrypanosomal metabolites from endophytic fungus Penicillium sp., isolated from Limonium tubiflorum. Bioorg. Med. Chem. 19 (2011) 414-421.

Oxacyclododecindione, a novel inhibitor of IL-4 signaling from Exserohilum rostratum, J.

SC

Antibiot. 61(5) (2008) 285–290.

Apralactone A and a new stereochemical class of curvularins from the marine-derived fungus

M AN U

Curvularia sp., Eur. J. Org. Chem. 2008 (30) (2008) doi:10.1002/ejoc.200800522 Draeger, B. Schulz, T. Kurtán, Curvularin-type metabolites from the fungus Curvularia sp. isolated from a marine alga, Eur J. Org. Chem. (2010) 6928–6937.

Cytotoxic Curvularin Macrolides from Penicillium sumatrense MA-92, a Fungus Obtained from the Rhizosphere of the Mangrove Lumnitzera racemosa, J. Nat. Prod. 76 (2013) 2145−2149.

TE D

1023-1025.

macrolides, pyrenolide and resorcylide. Koen Yoshishu - Tennen Yuki Kagobutsu Toronkai, 22 (1979) 362−369.

EP

Cornell University, Ithaca, NY (1989)

of maize, produces dihydroresorcylide and 7-hydroxydihydroresorcylides, J. Agric. Food Chem.

AC C

56 (2008) 3006−3009.

m-resorcylide, Agric. Biol. Chem. 42 (1978) 2407-2409. pharmaceutical composition containing it, and use thereof in the treatment of thrombosis and related disorders, (1996) (Zymogenetics, Inc., USA; Novo Nordisk A/S). Patent WO 9640671. remarkable hydrogen-bondcontrolled, stereospecific ring-closing metathesis, Org. Lett. 6 (2004) 977–980. (S)-dihydroresorcylide, a bioactive twelve-membered macrolide, Chin. J. Chem. 31 (2013) 339-343.

51

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[139] G. Strobel, D. Kenfield, G. Bunkers, F. Sugawara, J. Clardy Phytotoxins as potential herbicides,

3 4

[140] T. Sassa, M. Nukina, M. Ikeda, Electrophilic reactivities and biological activities of trans- and

5 6

[141] R.R. West, T. Martinez, H.R. Franklin, P.D. Bishop, Factor XIIIA inhibitor. US Patent 5 710

7 8 9

[142] M. Isaka, P. Chinthanom, S. Veeranondha, S. Supothina, J.J. Luangsa-ard, Novel cyclopropyl

Experientia 47 (1991) 819-826. cis-resorcylides, Nippon Kagaku Kaishi (1981) 883−885. 174. 1998.

RI PT

diketones and 14-membered macrolides from the soil fungus Hamigera avellanea BCC 17816, Tetrahedron 64 (2008) 11028–11033.

10 11 12

[143] M. Isaka, C. Suyarnsestakorn, M. Tanticharoen, P. Kongsaeree, Y. Thebtaranonth,

13 14 15

[144] M. Isaka, A. Yangchum, S. Intamas, K. Kocharin, E.B.G. Jones, P. Kongsaeree, S. Prabpai,

16 17

[145] E. Moulin, S. Barluenga, F. Totzke, N. Winssinger, Diversity-oriented synthesis of pochonins

18 19 20

[146] J. Xu, A. Chen, M-L. Go, K. Nacro, B. Liu, C.L.L. Chai, Exploring aigialomycin D and its

21 22 23 24

[147] Z.Q. Yang, X. Geng, D. Solit, C.A. Pratilas, N. Rosen, S.J. Danishefsky, New efficient synthesis

25 26 27

[148] X. D. Geng, S.J. Danishefsky, Total Synthesis of Aigialomycin D, Org. Lett. 6 (2004) 413. [149] J. Lu, J. Ma, X. Xie, B. Chen, X. She, X. Pan, Enantioselective total synthesis of aigialomycin

28 29

[150] N.Q. Vu, C.L.L. Chai, K. P. Lim, S. C. Chia, C. Anqi, An efficient and practical total synthesis

30 31 32

[151] C.C. Chrovian, B. Knapp-Reed, J. Montgomery, Total synthesis of aigialomycin D: surprising

33

[152] L.J. Baird, M.S.M. Timmer, P.H. Teesdale-Spittle, J.E. Harvey, Total synthesis of aigialomycin

SC

parvus, Org. Chem. 67 (2002) 1561–1566

Aigialomycins and related polyketide metabolites from the mangrove fungus Aigialus parvus BCC 5311, Tetrahedron 65 (2009) 4396–4403.

M AN U

and biological evaluation against a panel of kinases. Chemistry 12 (2006) 8819-8834. analogues as protein kinase inhibitors for cancer targets, ACS Med. Chem. Lett. 2 (2011) 662–666.

of resorcinylic macrolides via Ynolides:  establishment of cycloproparadicicol as synthetically feasible preclinical anticancer agent based on Hsp90 as the target. J. Am. Chem. Soc. 126

TE D

(2004) 7881–7889.

D, Tetrahedron: Asymmetry 17 (2006) 1066-1073.

EP

of aigialomycin D, Tetrahedron 63 (2007) 7053. chemoselectivity dependence on alkyne structure in nickel-catalyzed cyclizations, Org. Lett. 10(5) (2008) 811-814. doi: 10.1021/ol702961v.

AC C

34

Aigialomycins A−E, new resorcylic macrolides from the marine mangrove fungus Aigialus

D using a Ramberg-Bäcklund/RCM strategy, J. Org. Chem. 74 (2009) 2271-2277.

35

[153] F. Calo, J. Richardson, A.G. M. Barrett, Synthesis of aigialomycin D, Org. Lett. 11 (2009) 4910.

36 37 38

[154] N. Bajwa, M.P. Jennings, Syntheses of epi-aigialomycin D and deoxy-aigialomycin C via a

39 40

[155] J.L. Wee, K. Sundermann, P. Licari, J. Galazzo, Cytotoxic hypothemycin analogues from

diastereoselective ring closing metathesis macrocyclization protocol, Tetrahedron Lett. 49 (2008) 390-393. Hypomyces subiculosus, J. Nat. Prod. 69 (2006) 1456–1459.

52

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3

[156] S-X. Yang, J-M. Gao, Q. Zhang, H. Laatsch, Toxic polyketides produced by Fusarium sp., an

4 5

[157] L. Xu, J. Xue, Y. Zou, S. He, X. Wei, Three new β-resorcylic acid lactones from Paecilomyces

6 7 8

[158] Q.A. Liu, C-L. Shao, Y.C. Gu, M. Blum, L-S. Gan, K. Wang, M. Chen, C-Y. Wang,

9 10 11

[159] J. Dong, Y. Zhu, H. Song, R. Li, H. He, H. Liu, R. Huang, Y. Zhou, L. Wang, Y. Cao, K.

12 13 14

[160] C.L. Shao, H.X. Wu, Ch.Y. Wang, Q.A. Liu, Y. Xu, M.Y. Wei, P.Y. Qian, Y.C. Gu, C.J. Zheng,

15 16 17

[161] C.L. Shao, H.X. Wu, Ch.Y. Wang, Q.A. Liu, Y. Xu, M.Y. Wei, P.Y. Qian, Y.C. Gu, C.J. Zheng,

18 19

[162] N. Jana, S. Nanda, Asymmetric Total syntheses of cochliomycin A and zeaenol, Eur J. Org.

20 21

[163] L. Wang, Y. Gao, J. Liu, C. Cai, Y. Du, Stereoselective total synthesis of cochliomycin A,

22 23

[164] Y. Gao, J. Liu, L. Wang, M. Xiao, Y. Du, Total syntheses of cochliomycin B and zeaenol. Eur J.

24 25 26

[165] M. Isaka, P. Chinthanom, S. Kongthong, S. Supothina, Pataranun Ittiworapong,

27 28 29

[166] M.S. Probasco, J.W. Kirby, J.D. Carrick, Progress towards the enantioselective total synthesis of

30 31 32

[167] A.H. Cleveland, M.S. Probasco, J.W. Kirby, J.D. Carrick, Efforts toward the asymmetric total

33 34

[168] T. Agatsuma, A. Takahashi, C. Kabuto, S. Nozoe, Revised structure and stereochemistry of

35 36 37 38

[169] H. Hussain, K. Krohn, U. Flörke, B. Schulz, S. Draeger, G. Pescitelli, P. Salvadori, S. Antus, T.

39 40

[170] K. Tatsuta, S. Takano, T. Sato, S. Nakano, The first total synthesis of a macrocyclic

41 42

[171] P. Sellès, R. Lett, Convergent stereospecific synthesis of C292 (or LL-Z1640-2), and

43 44

[172] P. Sellès, R. Lett, Convergent stereospecific synthesis of LL-Z1640-2 (or C292), hypothemycin

endophytic fungus isolated from Melia azedarach, Bioorg. Med. Chem. Lett. 21 (2011) 1887-1889. sp SC0924, Chin. J. Chem. 30 (2012) 1273—1277.

Cochliobolus lunatus, J. Agric. Food Chem. 62 (2014) 3183-3191.

RI PT

Antifouling and fungicidal resorcylic acid lactones from the sea anemone-derived fungus

Zhang, Nematicidal resorcylides from the aquatic fungus Caryospora callicarpa YMF1.01026. J. Chem. Ecol. 33 (2007) 1115–1126.

SC

Z.G. She, Y.C. Lin, Correction to potent antifouling resorcylic acid lactones from the gorgonian-derived fungus Cochliobolus lunatus, J. Nat. Prod. 76 (2013) 302.

Z.G. She, Y.C. Lin, Potent Antifouling resorcylic acid lactones from the gorgonian-derived

M AN U

fungus Cochliobolus lunatus, J. Nat. Prod. 74 (2011) 629–633. Chem. (2012) 4313–4320 DOI: 10.1002/ejoc.201200241. Tetrahedron 70 (2014) 2616-2620.

TE D

Org. Chem. (2014) 2092–2098 DOI: 10.1002/ejoc.201301613.

Hamigeromycins C–G, 14-membered macrolides from the fungus Hamigera avellanea BCC 17816, Tetrahedron 66 (2010) 955–961.

hamigeromycin, 245th ACS National Meeting & Exposition, New Orleans, LA, United States,

EP

April 7-11, 2013 (2013), ORGN-191.

synthesis of hamigeromycin b. Abstracts of Papers, 247th ACS National Meeting & Exposition,

AC C

Dallas, TX, United States, March 16-20, 2014 (2014), ORGN-599. hypothemycin, Chem. Pharm. Bull. 41 (1993) 373–375. Kurtán, Absolute configuration of hypothemycin and 5'-O-methylhypothemycin from Phoma sp.—a test case for solid state CD/TDDFT approach, Tetrahedron: Asymmetry 18 (2007) 925–930. anti-protozoan, LL-Z1640-2, Chem. Lett. 30 (2001) 172-173. hypothemycin, Part 1, Tetrahedron Lett. 43 (2002) 4621-4625. and related macrolides, Part 2 Tetrahedron Lett. 43 (2002) 4627-4631. 53

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[173] M.S.R. Nair, S.T. Carey, Metabolites of pyrenomycetes XIII: Structure of (+) hypothemycin, an

3 4 5

[174] M.S.R. Nair, S.T. Carey, J.C. James, Metabolites of pyrenomycetes. XIV. Structure and partial

6 7 8 9

[175] M. Nishino, J.W. Choy, N.N. Gushwa, J.A. Oses-Prieto, K. Koupparis, A.L. Burlingame, A.R.

10 11 12

[176] A. Schirmer, J. Kennedy, S. Murli, R. Reid, D.V. Santi, Targeted covalent inactivation of

13 14 15 16

[177] H. Fukazawa, Y. Ikeda, M. Fukuyama, T. Suzuki, H. Hori, T. Okuda, Y. Uehara, The resorcylic

17 18 19

[178] R. Camacho, M.J. Staruch, C. DaSilva, S. Koprak, T. Sewell, F.J. Dumont, Hypothemycin

20 21

[179] H. Tanaka, K. Nishida, K. Sugita, T. Yoshioka, Antitumor efficacy of hypothemycin, A new

22 23 24

[180] H. Sonoda, K. Omi, K. Hojo, K. Nishida, K. Sugita, Suppression of oncogenic transformation

25 26 27

[181] L. Xu, J. Xue, P. Wu, D. Wang, L. Lin, Y. Jiang, X. Duan, X. Wei, Antifungal activity of

28 29

[182] A. Zhao, S.H. Lee, M. Mojena, Resorcylic acid lactones: naturally occurring potent and

30 31

[183] P. Delmotte, J. Delmotte-Plaque´e, A new antifungal substance of fungal origin, Nature 171

32 33

[184] R.N. Mirrington, E. Ritchie, C.W. Shoppee, W.C. Taylor, S. Sternhell, The constitution of

34 35 36

[185] R.N. Mirrington, E. Ritchie, C.W. Shoppee, S. Sternhell, W.C. Taylor, Some metabolites of

37 38 39

[186] H.G. Cutler, R.F. Arrendale, J.P. Springer, P.D. Cole, R.G. Roberts, R.T. Hanlin, Monorden

40 41

[187] M. Lampilas, R. Lett, Convergent stereospecific total synthesis of monochiral monocillin I

42 43

[188] M. Lampilas, R. Lett, Convergent stereospecific total synthesis of monocillin I and monorden

antibiotic macrolide from Hypomyces trichothecoides, Tetrahedron Lett. 21 (1980) 2011–2012. stereochemistry of the antibiotic macrolides hypothemycin and dihydrohypothemycin, Tetrahedron 37 (1981) 2445-2449. Renslo, J.H. McKerrow, J. Taunton, Hypothemycin, a fungal natural product, identifies

RI PT

therapeutic targets in Trypanosoma brucei. Biochemistry | Cell biology eLife 2 (2013) e00712. DOI:10.7554/eLife.00712.

protein kinases by resorcylic acid lactone polyketides, Proc. Natl. Acad. Sci. 103 (2006)

SC

4234–4239.

acid lactone hypothemycin selectively inhibits the mitogen-activated protein kinase kinase-extracellular signal-regulated kinase pathway in cells, Biol. Pharm. Bull. 33(2) (2010)

M AN U

168—173.

inhibits the proliferative response and modulates the production of cytokines during T cell activation, Immunopharmacology 44 (1999) 255-265.

ras-signaling inhibitor, Cancer Sci. 90 (1999) 1139–1145. by

hypothemycin

associated

with

accelerated

cyclin

D1

degradation

through

TE D

ubiquitin-proteasome pathway, Life Sci. 65 (1999) 381-394.

hypothemycin against Peronophythora Litchii in vitro and in vivo, J. Agric. Food Chem. 61 (2013) 10091−10095.

EP

selective inhibitors of MEK, J. Antibiot. 52 (1999) 1086-1094.

AC C

(1953) 344-347.

radicicol, Tetrahedron Lett. 7 (1964) 365-370. Nectria radicicola Gerlach & Nilsson (syn. Cylindrocarpon radicicola WR.): the structure of radicicol (monorden), Aust. J. Chem. 19 (1966) 1265-1284. from a novel source, Neocosmospora tenuicristata, stereochemistry and plant growth regulatory properties, Agric. Biol. Chem. 51 (1987) 3331–3338. related macrolides, Tetrhedron Lett. 33 (1992) 773-776. (or radicicol), Tetrhedron Lett. 33 (1992) 777-780.

54

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[189] R.M. Garbaccio, S.J. Danishefsky, Efficient asymmetric synthesis of radicicol dimethyl ether: a

3 4

[190] H. Zhou, K. Qiao, Z. Gao, J.C. Vederas, Y. Tang, Insights into radicicol biosynthesis via

5 6 7

[191] S. Wang, Y. Xu, E.A. Maine, E. Wijeratne, P. Espinosa-Artiles, A. Gunatilaka, I. Molnár,

8 9

[192] J. Zeng, J. Zhan, A novel fungal flavin-dependent halogenase for natural product biosynthesis,

10 11 12 13

[193] T.J. Turbyville, E.M.K. Wijeratne, M.X. Liu, A.M. Burns, C.J. Seliga, L.A. Leuvano, C.L.

14 15 16

[194] D.T. Wicklow, A.M. Jordan, J.B. Gloer, Antifungal metabolites (monorden, monocillins I, II,

17 18 19

[195] Y. Tanaka, K. Shiomi, K. Kamei, M. SugohHagino, Y. Enomoto, F. Fang, Y. Yamaguchi, R.

20 21

[196] M. Arai, K. Yamamoto, I. Namatame, H. Tomoda, S. Omura, New monordens produced by

22 23 24 25

[197] E.M.K. Wijeratne, C.A. Carbonezi, J.A. Takahasi, C.J., Seliga T.J. Turbyville, E.E. Pierson,

26 27

[198] M. Stadler, H-V. Tichy, E. Katsiou, V. Hellwig, Chemotaxonomy of pochonia and other

28 29

[199] K. Nozawa, S. Nakajima, Isolation of radicicol from Penicillium luteo-aurantium, and

30

[200] D.T. Wicklow, B.K. Joshi, W.R. Gamble, J.B. Gloer, P.F. Dowd, Antifungal metabolites

31

(monorden, monocillin iv, and cerebrosides) from Humicola fuscoatra Traaen NRRL 22980, a

heterologous synthesis of intermediates and analogs, J. Biol. Chem. 285 (2010) 41412–41421. Functional characterization of the biosynthesis of radicicol, an Hsp90 inhibitor resorcylic acid

RI PT

lactone from Chaetomium chiversii, Chem. Biol. 15 (2008) 1328–1338. Chem. Biol. Chem. 11 (2010) 2119–2123.

David, S.H. Faeth, L. Whitesell, A.A.L. Gunatilaka, Search for Hsp90 inhibitors with potent anticancer activity: isolation and SAR studies of radicicol and monocillin I from two plant

SC

–associated fungi of the sonoran desert, J. Nat. Prod. 69 (2006) 178-184.

III) from Colletotrichum graminicola, a systemic vascular pathogen of maize, Mycol. Res. 12

M AN U

(2009) 1433-1442.

Masuma, C.G. Zhang, X.W. Zhang, S. Omura, Antimalarial activity of radicicol, heptelidic acid and other fungal metabolites, J. Antibiot. 51 (1998) 153-160.

amidepsine-producing fungus Humicola sp FO-2942. J. Antibiot. 6 (2003) 526-533. L.S. Pierson, H.D. VanEtten, L. Whitesell, V.D.S. Bolzani, A.A.L. Gunatilaka, Isolation,

TE D

optimization of production and structure-activity relationship studies of monocillin I, the cytotoxic constituent of Paraphaeosphaeria quaddriseptata, J. Antibiot. 57 (2004) 541-546. conidial fungi with verticillium-like anamorphs, Mycological Pr200ogress 2 (2003) 95-122.

EP

meleagrin, a new metabolite, from Penicillium meleagrinum, J. Nat. Prod. 4 (1979) 374-377.

AC C

32

novel application of ring-forming olefin metathesis, Org. Lett. 2 (2000) 3127-3129.

mycoparasite of Aspergillus flavus sclerotia, Appl. Environ. Microbiol. 11 (1998) 4482–4484.

33 34

[201] H.J. Kwon, M. Yoshida, Y. Fukui, S. Horinouchi, T. Beppu, Potent and specific inhibition of

35 36 37 38 39

[202] P. Chanmugam, L. Feng, S. Liou, B.C. Jang, M. Boudreau, G. Yu, J.H. Lee, H.J. Kwon, T.

40 41 42

[203] S. Soga, T. Kozawa, H. Narumi, S. Akinaga, K. Irie, K. Matsumoto, S.V. Sharma, H. Nakano,

p60v-src protein kinase both invivo and in vitro by radicicol, Cancer Res. 52 (1992) 6926–6930. Beppu, M. Yoshida, Y. Xia, C.B. Wilson, D. Hwang, Radicicol, a potent protein tyrosine kinase inhibitor, suppresses the expression of mitogen-induced cyclooxigenase in macrophages stimulated with lipopolysaccharide and in experimental glomerulonephritis, J. Biol. Chem. 270 (1995) 5418–5426. T. Mizukami, M. Hara, Radicicol leads to selective depletion of Raf kinase and disrupts KRasactivated aberrant signaling pathway, J. Biol. Chem. 273 (1998) 822–828. 55

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2 3 4

[204] H. Jayasuriya, D.L. Zink, J.D. Polishook, G.F. Bills, A.W. Dombrowski, O. Genilloud, F.F.

5 6 7

[205] H. Wei, L. Xu, M. Yu, L. Zhang, H. Wang, X. Wei, Y. Ruan, Monocillin II Inhibits human

8 9

[206] S.V. Sharma, T. Agatsuma, H. Nakano, Targeting of the protein chaperone, Hsp90, by the

10 11 12

[207] S.M. Roe, C. Prodromou, R. O’Brien, J.E. Ladbury, P.W. Piper, L.H. Pearl, Structural basis for

13 14 15

[208] T.W. Schulte, S. Akinaga, S. Soga, W. Sullivan, B. Stensgard, D. Toft, L.M. Neckers,

16 17 18

[209] T. Agatsuma, H. Ogawa, K. Akasaka, A. Asai, Y. Yamashita, T. Mizukami, S.A.Y. Saitoh,

19 20

[210] N. Winssinger, J.G. Fontaine, S. Barluenga, Hsp90 inhibition with resorcyclic acid lactones

21 22 23

[211] K. Yamamotoa, H. Hatanoa, M. Araia, K. Shiomia, H. Tomodaa, S. Omuraa, Structure

24 25 26

[212] V. Hellwig, A. Mayer-Bartschmid, H. Muller, G. Greif, G. Kleymann, W. Zitzmann, H.V.

27 28 29

[213] W.A. Ayer, S.P. Lee, A. Tsuneda, Y. Hiratsuka, The isolation, identification, and bioassay of

30 31

[214] W.A. Ayer, L. Pena-Rodriguez, Minor metabolites of Monocillium nordinii, Phytochemistry 26

32 33

[215] I. Tichkowsky, R. Lett, Improvements of the total synthesis of monocillin I and radicicol via

34 35 36

[216] H. Shinonaga, Y. Kawamura, A. Ikeda, M. Aoki, N. Sakai, N. Fujimoto, A. Kawashima, The

37 38 39 40 41

[217] E.J. Mejia, S.T. Loveridge, G. Stepan, A. Tsai, G.S. Jones, T. Barnes, K.N. White, M.D. ković,

42 43

[218] J. Gao, M.M. Radwan, F. León, O.R. Dale, A.S. Husni, Y. Wu, S. Lupien, X. Wang, S.P.

Pelaez, L. Herranz, D. Quamina, R.B. Lingham, R. Danzeizen, P.L. Graham, J.E. Tomassini, S.B. Singh, Identification of diverse microbial metabolites as potent inhibitors of HIV-1 tat transfiguration, Chem Biodiversity 2 (2005) 112-122. breast cancer

growthpartially by inhibiting MAPK

pathways and

CDK2

Thr160

RI PT

phosphorylation, ChemBioChem 13 (2012) 465–475. transformation suppressing agent, radicicol, Oncogene 16 (1998) 2639–2645.

inhibition of the Hsp90 molecular chaperone by the antitumor antibiotics radicicol and

SC

geldanamycin, J. Med. Chem. 42 (1999) 260-266.

Antibiotic radicicol binds to the N-terminal domain of Hsp90, and shares important biologic activities with geldanamycin, Cell Stress Chaperones 3 (1998) 100–108.

Chem. 10 (2002) 3445–3454.

M AN U

Halohydrin and oxime derivatives of radicicol: synthesis and antitumor activities, Bioorg. Med.

(RALs), Curr. Top. Med. Chem. 9 (2009) 1419-1435.

elucidation of new monordens produced by Humicola sp. FO-2942, J. Antibiot. 56 (2003)

TE D

533-538.

Tichy, M. Stadler, Pochonins A-F, new antiviral and antiparasitic resorcylic acid lactones from Pochonia chlamydosporia var. catenulate, J. Nat. Prod. 66 (2003) 829 -837. the antifungal metabolites produced by Monocillium nordinii, Can. J. Microbiol. 26 (1980)

EP

766-773.

AC C

(1987) 1353-1355.

Miyaura-Suzuki couplings, Tetrahedron Lett. 43 (2002) 4003–4007. search for a hair-growth stimulant: new radicicol analogues as WNT-5A expression inhibitors from Pochonia chlamydosporia var. chlamydosporia, Tetrahedron Lett. 50 (2009) 108-110. K. Tenney, M. Tsiang, R. Geleziunas, T. Cihlar, N. Pagratis, Y. Tian, H. Yu, C. Phillip, Study of marine natural products including resorcyclic acid lactones from Humicola fuscoatra that reactivate latent HIV‑1 expression in an in vitro model of central memory CD4+ T Cells, J. Nat. Prod. 77 (2014) 618−624. Manly, R.A. Hill,

F.M. Dugan, H.G. Cutler, S.J. Cutler, Neocosmospora sp.-derived resorcylic

56

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

acid lactones with in vitro binding affinity for human opioid and cannabinoid receptors, J. Nat.

3 4

[219] L. Xu, Z. He, J. Xue, X. Chen, X. Wei, β-resorcylic acid lactones from a Paecilomyces fungus,

5 6

[220] L. Xu, Z. He, J. Xue, X. Chen, X. Wei, Paecilomycin F was also isolated from culture broth of

7 8

[221] L. Xu, P. Wu, H. Wei, J. Xue, X. Hu, X. Wei, Paecilomycins J–M, four new b-resorcylic acid

Prod. 76 (2013) 824−828. J. Nat. Prod. 73 (2010) 885–889. Cochlioolus lunatus, J. Nat. Prod. 73 (2012) 885–889.

RI PT

lactones from Paecilomyces sp. SC0924, Tetrahedron Lett. 54 (2013) 2648–2650.

9 10

[222] L. Xu, Z. He, J. Xue, X. Chen, X. Wei, Correction to β-resorcylic acid lactones from a

11 12

[223] L. Xu, J. Xue, P. Wu, X. Wei, Absolute configurations of four resorcylic acid lactones,

13 14

[224] P. Srihari, B. Mahankali, K. Rajendraprasad, Stereoselective total synthesis of paecilomycin E,

15 16 17

[225] H. Shinonaga, Y. Kawamura, A. Ikeda, M. Aoki, N. Sakai, N. Fujimoto, A. Kawashima,

18 19

[226] S. Barluenga, P. Lopez, E. Moulin, N. Winssinger, Modular asymmetric synthesis of pochonin

20 21

[227] S. Barluenga, P. Lopez, E. Moulin, N. Winssinger, Modular asymmetric synthesis of pochonin

22 23 24

[228] G. Karthikeyan, C. Zambaldo, S. Barluenga, V. Zoete, M. Karplus, N. Winssinger, Asymmetric

25 26 27

[229] D. Martinez-Solorio, K.A. Belmore, M.P. Jennings, Synthesis of the purported ent-pochonin j

28 29 30

[230] E. Moulin, V. Zoete, S. Barluenga, M. Karplus, N. Winssinger, Design, synthesis, and

31 32 33

[231] G.A. Ellestad, F.M. Lovell, N.A. Perkinson, R.T. Hargreaves, W.J. McGahren, New

34 35 36

[232] T. Yuki, K. Toronkai, K. Yoshishu, Chemical studies on aversion-antagonism among different

37 38 39

[233] P. Rawlinsa, T. Mandera, R. Sadeghia, S. Hilla, G. Gammona, B. Foxwellf, S. Wrigleya, M.

40 41 42

[234] F. Sugawara, K.W. Kim, K. Kobayashi, J. Uzawa, S. Yoshida, N. Murofushi, N. Takahashi, G.A.

43 44

[235] A. Dombrowski, R. Jenkins, S. Raghoobar, G. Bills, J. Polishook, F. PelAez, B. Burgess, A.

paecilomyces fungus. J. Nat. Prod. 75 (2012) 1006−1006.

SC

Paecilomycins J_M, by CD/TDDFT Calculations, Chirality 26 (2014) 44–50. Tetrahedron Lett. 53 (2012) 56–58.

Pochonins K–P: new radicicol analogues from Pochonia chlamydosporia var chlamydosporia

M AN U

and their WNT-5A expression inhibitory activities. Tetrahedron 65 (2009) 3446–3453. C, Angew Chem 116 (2004) 3549–3552. C. Angew Chem 43 (2004) 3467-3470.

synthesis of pochonin E and F, revision of their proposed structure, and their conversion to

TE D

potent hsp90 inhibitors. Chem. Eur. J. 18 (2012) 8978-8986.

structure featuring a stereoselective oxocarbenium allylation, J. Org. Chem. 76 (2011) 3898–3908.

EP

biological evaluation of Hsp90 inhibitors based on conformational analysis of radicicol and its analogues, J. Am. Chem. Soc. 127 (2005) 6999-7004.

AC C

zearalenone related macrolidesand isocoumarins from an unidentired fungus, J. Org. Chem. 43 (1978) 2339-2343.

strains of the same fungal species. Aversion factor and new metabolites of Cochliobolus lunata, Hokkaido Daigaku Nogakubu 21 (1978) 144-151. Moore, Inhibition of endotoxin-induced TNF-a production in macrophages by 5Z-7-oxo-zeaenol and other fungal resorcylic acid lactones, Int. J. Immunopharmacol. 21 (1999) 799-814. Strobel, Zearalenone derivatives produced by the fungus Drechslera portulacae, Phytochemistry 31 (1992) 1987-1990. Zhao, L. Huang, Y. Zhang, M. Goetz, Production of a family of kinase-inhibiting lactones from 57

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1

fungal fermentations, J. Antibiot. 52 (1999) 1077-1085.

2 3 4

[236] P.Y. Dakas, R. Jogireddy, G. Valot, S. Barluenga, N. Winssinger, divergent syntheses of

5 6 7

[237] M. Ohori, T. Kinoshita, S. Yoshimura, M. Warizaya, H. Nakajima, H. Miyake, Role of a cysteine

8 9

[238] W.T. Shier, A.C. Shier, W. Xie, C.J. Mirocha, Structure-activity relationships for human

10 11

[239] M. Metzler, E. Pfeiffer, Genotoxic potential of xenobiotic growth promoters and their

12 13 14

[240] A. Zinedine, J.M. Soriano, J.C. Molto´, J. Man˜ es, Review on the toxicity, occurrence,

15 16

[241] M. Metzler, E. Pfeiffer, A.A. Hildebrand, Zearalenone and its metabolites as endocrine

17 18

[242] J. Stob, R.S. Baldwin, J. Tuite, F.N. Andrews, K.G. Gillette, Isolation of an anabolic,

19 20

[243] W.H. Urry, H.L. Wehrmeister, E.B. Hodge, P.H. Hidy, The structure of zearalenone,

21 22 23

[244] F. Bravin, R.C. Duca, P. Balaguer, M. Delaforge, In vitro cytochrome P450 formation of a

24 25

[245] J.A. Ruddick, P.M. Scott, J. Harwig, Teratological evaluation of zearalenone administered

26 27

[246] L. Ghedira-Chekir, K. Maaroufi, E.E. Creppy, H. Bacha, Cytotoxic and genotoxic effects of

28 29 30

[247] S. Abid-Essefi, I. Baudrimont, W. Hassen, DNA fragmentation, apoptosis and cell cycle arrest

31 32 33

[248] S. Abid-Essefi, Z. Ouanes, W. Hassen, Cytotoxicity, inhibition of DNA and protein syntheses

34 35

[249] M.B. Lioi, A. Santoro, R. Barbieri, Ochratoxin and zearalenone: a comparative study on

36 37 38

[250] H.A. Atkinson, K. Miller, Inhibitory effect of deoxynivalenol, 3-acetyldeoxynivalenol and

39 40 41

[251] J.H. Pestka, M.F. Tai, D.E. Witt, J.H. Dixon, Forsell suppression of immune response in the

42 43 44

[252] D. Luongo, R. De Luna, R. Russo, Effects of four Fusarium toxins (fumonisin B(1),

resorcylic acid lactones: L-783277, LL-Z1640-2, and hypothemycin, Chemistry 15 (2009) 11490–11497. residue in the active site of ERK and the MAPKK family, Biochem Biophys Res Commun 353

RI PT

(2007) 633–637. estrogenic activity in zearalenone mycotoxins, Toxicon 39 (2001) 1435-1438. metabolites, APMIS 109 (2001) 89–95.

Food Chem. Toxicol. 45 (2007) 1-18.

SC

metabolism, detoxification, regulations and intake of zearalenone: an oestrogenic mycotoxin,

M AN U

disrupting chemicals. World Mycotoxin J. 3 (4) (2010) 385-401.

utertrophic compound from corn infected with Gibberella zeae. Nature 196 (1962) 1318. Tetrahedron Lett. 7 (1966) 3109-3114.

mono-hydroxylated metabolite of zearalenone exhibiting estrogenic activities possible

TE D

occurrence of this metabolite in vivo, Int. J. Mol. Sci. 10 (2009) 1824-1837. orally to the rat, Bull. Environ. Contam. Toxicol. 15 (1976) 678–681. zearalenone: prevention by vitamin E, J. Toxicol. Toxin. Rev. 18 (1999) 355–368.

EP

induced by zearalenone in cultured DOK, Vero and Caco-2 cells: prevention by vitamin E., Toxicology 192 (2003) 237–248.

AC C

and oxidative damage in cultured cells exposed to zearalenone, Toxicol. In Vitro 18 (2004) 467–474.

genotoxic effects and cell death induced in bovine lymphocytes, Mutat. Res. 557 (2004) 19–24. zearalenone on induction of rat and human lymphocyte proliferation, Toxicol. Lett. 23 (1984) 215–221. B6C3F1 mouse after dietary exposure to the Fusarium mycotoxins deoxynivalenol (vomitoxin) and zearalenone, Food Chem. Toxicol. 25 (1987) 297–304. alpha-zearalenol, nivalenol and deoxynivalenol) on porcine whole-blood cellular proliferation, Toxicon 52 (2008) 156–162. 58

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1 2

[253] Y.C. Wang, J.L. Deng, S.W. Xu, Effects of zearalenone on IL-2, IL-6, and IFN-g mRNA levels

3 4 5

[254] Z. Ouanes, S. Abid, I. Ayed, Induction of micronuclei by zearalenone in Vero monkey kidney

6 7 8

[255] C. Bouaziz, C. Martel, O. Sharaf el dein, Fusarial toxin-induced toxicity in cultured cells and in

9 10 11

[256] L. Zourgui, E.E. Golli, C. Bouaziz, Cactus (Opuntia ficus-indica) cladodes prevent oxidative

12 13 14

[257] K.P. Lone, Natural sex steroids and their xenobiotic analogs in animal production: growth,

15 16

[258] W.H. Utian, Comparative trial of P1496, a new non-steroidal oestrogen analogue. Br. Med. J. 1

17 18

[259] K.A. Scudamore, S. Patel, Survey for aflatoxins, ochratoxin A, zearalenone and fumonisins in

19 20 21 22

[260] S.Z. Jiang, Z.B. Yang, W.R. Yang, B.Q. Yao, H. Zhao, F.X. Liu, C.C. Chen, F. Chi, Effects of

23 24

[261] L.G. Young, H. Ping, G.J. King, Effects of feeding zearalenone to sows on rebreeding and

25 26

[262] M. Etienne, J.Y. Dourmand, Effects of zearalenone or glucosinolates in the diet on reproduction

27 28

[263] T. Kuiper-Goodman, P.M. Scott, H. Watanabe, Risk assessment of the mycotoxin zearalenone,

29 30 31

[264] H. Bacha, L. Chekir, F. Ellouz, R. Hadidane, E.E. Creppy, Effect of zearalenone on fertilisation

32 33 34

[265] J.P.F. D’Mello, C.M. Placinta, M.C. Macdonald, Fusarium mycotoxins, a review of global

35 36

[266] T.F.X. Collins, R.L. Sprando, T.N. Black, Effects of aminopentol on in utero development in

37 38 39

[267] F. Koraichi, B. Videmann, M. Mazallon, M. Benahmed, C. Prouillaca, S. Lecoeurc, Zearalenone

40 41

[268] Y.T. Kim, Two different polyketide synthase genes are required for synthesis of zearalenone in

42 43

[269] I. Gaffoor, F. Trail, Characterization of two polyketide synthase genes involved in zearalenone

44

[270] D. Taub, N.N. Girotra, R.D. Hoffsommer, C.H. Kuo, H.L. Slates, S. Weber, Total synthesis of

in the splenic lymphocytes of chickens, Sci. World J. (2012) (article ID 567327). cells and in bone marrow cells of mice: protective effect of Vitamin E., Mutat. Res. 538 (2003) 63–70. isolated mitochondria involves PTPC-dependent activation of the mitochondrial pathway of

RI PT

apoptosis, Toxicol. Sci. 110 (2009) 363–375.

damage induced by the mycotoxin zearalenone in Balb/C mice, Food Chem. Toxicol. 46 (2008) 1817–1824.

SC

carcass quality, pharmacokinetics, metabolism, mode of action, residues, methods, and epidemiology, CRC Crit. Rev. Food Sci. Nutr. 37 (1997) 93–209.

M AN U

(1973) 579-581.

maize imported into the United Kingdom, Food Addit. Contam. 15 (2000) 30–55. feeding purified zearalenone contaminated diets with or without clay enterosorbent on growth, nutrient availability, and genital organs in post-weaning female pigs, Asian-Australas J. Anim. Sci. 23 (2010) 74–81.

TE D

pregnancy, J. Anim. Sci. 68 (1990) 15–20.

in sows: a review, Livest. Prod. Sci. 40 (1994) 99–113

EP

Regul. Toxicol. Pharmacol. 7 (1987) 253–306. and gestation in rats K.A. Scudamore (Ed.), Proceeding of UK Workshop on Occurrence in

AC C

Significance of Mycotoxins, Central Science Laboratory, London (1993), pp. 258–262. implications for animal health, welfare and productivity. Anim. Feed Sci. Technol. 80 (1999) 183–205.

rats. Food Chem. Toxicol. 44 (2006) 161–169. exposure modulates the expression of ABC transporters and nuclear receptors in pregnant rats and fetal liver, Toxicology Lett. 211 (3) (2012) 246–256. Gibberella zeae. Mol Microbiol 58 (2005) 1102–1113. biosynthesis in Gibberella zeae. Appl Environ Microbiol 72 (2006) 1793–1799.

59

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1

the macrolide, zearalenone, Chem. Commun. (London) 1967 (1976) 225–226.

2 3 4

[271] G. Bolliger, C. Tamm, Vier neue Metabolite von Giberella zeae: 5-Formyl-zearalenon,

5 6 7

[272] R.A. Jackson, S.W. Fenton, C.J. Mirocha, G. Davis, Characterization of two isomers of

8 9

[273] R.D. Stipanovic, H.W. Schroeder, Zearalenol and 8’-hydroxyzearalenone from Fusarium

10 11 12

[274] W.M. Hagler, C.J. Mirocha, S.V. Pathre, J.C. Behrens, Identification of the naturally occurring

13 14

[275] J.A. Steele, C.J. Mirocha, S.V. Pathre, Metabolism of zearalenone by Fusarium roseum

15 16

[276] S.V. Pathre, S.W. Fenton, C.J. Mirocha, 3’-Hydroxyzearalenones, two new metabolites

17 18 19

[277] K.E. Richardson, W.M. Hagler, C.J. Mirocha, Production of zearalenone, α- and β-zearalenol,

20 21

[278] Y. Ueno, S. Ayaki, N. Sato, T. Ito, Fate and mode of action of zearalenone. Annales de la

22 23

[279] C.J. Mirocha, S.V. Pathre, T.S. Robinson, Comparative metabolism of zearalenone and

24 25 26

[280] M. Olsen, H. Pettersson, K.H. Kiessling, Reduction of zearalenone to zearalenol in female rat

27 28 29

[281] M. Olsen, K.H. Kiessling, Species differences in zearalenone-reducing activity in subcellular

30 31

[282] Y. Ueno, F. Tashiro, α-Zearalenol, a major hepatic metabolite in rats of zearalenone, an

32 33

[283] V.G. Smith, R.G. Brown, R.R. Hacker, Effects of ralgro on stress response in steers. Journal Of

34 35

[284] H. Malekinejad, R.F. Maas-Bakker, J. Fink-Gremmels, Bioactivation of zearalenone by porcine

36 37

[285] H. Malekinejad, R. Maas-Bakker, J. Fink-Gremmels, Species differences in the hepatic

38 39

[286] S.H. El-Sharkawy, Y.J. Abul-Hajj, Microbial transformation of zearalenone. 2. reduction,

40 41

[287] G. Engelhardt, G. Zill, B. Wohner, P.R. Wallnöfer, Transformation of the Fusarium mycotoxin

42 43

[288] J. Plasencia, C.J. Mirocha, Isolation and characterization of zearalenone sulfate produced by

44

[289] F. Berthiller, R. Schuhmacher, G. Adam, R. Krska, Formation, determination and significance

7’-Dehydrozearalenon, 8’-Hydroxy- und 8’-epi-Hydroxy-zearalenon, Helv. Chim. Acta 55 (1972) 3030-3048. 8’-hydroxyzearalenone and other derivatives of zearalenone, J. Agric. Food Chem. 22 (1974)

RI PT

1015-1019. roseum. Mycopathologia 57 (1975) 77-78.

isomer of zearalenol produced by Fusarium roseum ’Gibbosum’ in rice culture. Applied and

graminearum, J. Agric. Food Chem. 24 (1976) 89-97.

SC

Environmental Microbiology 37 (1979) 849-853.

M AN U

produced by Fusarium roseum, J. Agric. Food Chem. 28 (1980) 421-424.

and α- and β-zearalanol by Fusarium spp. in rice culture, J. Agric. Food Chem. 33 (1985) 862-866.

Nutrition de l’Alimentation 31 (1977) 935-948.

liver

by

TE D

transmission into bovine milk, Food and Cosmet. Toxicol. 19 (1981) 25-30. 3α-hydroxysteroid

dehydrogenase,

Acta

Pharmacologica

and

Toxicologica

(Copenhagen) 48 (1981) 157-161.

fractions of liver from female domestic animals. Acta Pharmacologica and Toxicologica

EP

(Copenhagen) 52 (1983) 287-291.

AC C

estrogenic mycotoxin of Fusarium species, J. Biochem. 89 (1981) 563-571. Animal Science 43 (1976) 305. hepatic biotransformation, Vet. Res. 36 (2005) 799-810. biotransformation of zearalenone, Vet. J. 172 (2006) 96-102. hydroxylation, and methylation products, J. Org. Chem. 53 (1988) 515-519. zearalenone in maize cell suspension cultures. Naturwissenschaften 75 (1988) 309-310. Fusarium spp., Appl. Environ. Microbiol. 57 (1991) 146-150.

60

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1

of masked and other conjugated mycotoxins, Anal. Bioanal. Chem. 395 (2009) 1243-1252.

2 3

[290] I.

4 5

[291] H. Kamimura, Conversion of zearalenone to zearalenone glycoside by Rhizopus sp., Appl.

6 7

[292] S. El-Sharkawy, Y. Abul-Hajj, Microbial transformation of zearalenone. 1. Formation of

8 9 10

[293] H. Mikula, C. Hametner, F. Berthiller, B. Warth, R. Krska, G. Adam, J. Froehlich, Fast and

11 12

[294] S.H. El-Sharkawy, M.I. Selim, M.S. Afifi, F.T. Halaweish, Microbial transformation of

13 14 15 16

[295] F. Berthiller, U. Werner, M. Sulyok, R. Krska, M.T. Hauser, R. Schuhmacher, Liquid

17 18 19

[296] O. Vendl, F. Berthiller, C. Crews, R. Krska, Determination of zearalenone and its metabolites in

20 21 22

[297] V. Vishwanath, M. Sulyok, R. Labuda, W. Bicker, R. Krska, Simultaneous determination of 186

23 24 25

[298] M. Cirlini, C. Dall’Asta, G. Galaverna, Hyphenated chromatographic techniques for structural

26 27 28

[299] L.L. Zhao, Y. Gai, H. Kobayashi, C.Q. Hu, H.P. Zhang, 5'-Hydroxyzearalenol, a new

29 30 31

[300] X. Yang, T.T. Khong, L. Chen, H.D. Choi, J.S. Kang, B.W. Son, 8'-Hydroxyzearalanone and

32 33 34 35

[301] E. Pfeiffer, A. Hildebrand, C. Becker, C. Schnattinger, S. Baumann, A. Rapp, H. Goesmann, C.

36 37 38

[302] T. Kastelic, J. Schnyder, A. Leutwiler, R. Traber, B. Streit, H. Niggli, A. MacKenzie, D.

39 40 41 42

[303] D. Cheneval, P. Ramage, T. Kastelic, T. Szelestenyi, H. Niggli, René Hemmig, M. Bachmann, A.

43 44

[304] P.Y. Dakas, S. Barluenga, F. Totzke, U. Zirrgiebel, N. Winssinger, Modular synthesis of radicicol

Schneweis,

K.

Meyer,

G.

Engelhardt,

J.

Bauer,

Occurrence

of

zearalenone-4-β-D-glucopyranoside in wheat. J. Agric. Food Chem. 50 (2002) 1736-1738. Environ. Microbiol. 52 (1986) 515-519.

RI PT

zearalenone 4-O-b-glucoside, J. Nat. Prod. 50 (1987) 520-521. reproducible chemical synthesis of zearalenone-14-β,D-glucuronide, World Mycotoxin. J. 5 (2012) 289–296.

SC

zearalenone to a zearalenone sulfate, Appl. Environ. Microbiol. (1991) 549–552.

chromatography coupled to tandem mass spectrometry (LC-MS/MS) determination of phase II metabolites of the mycotoxin zearalenone in the model plant Arabidopsis thaliana, Food Addit.

M AN U

Contam. 23 (2006) 1194–1200.

endometrial cancer by coupled separation techniques, Anal Bioanal Chem 395 (2009) 1347–1354.

fungal and bacterial metabolites in indoor matrices by liquid chromatography/tandem mass spectrometry, Anal. Bioanal. Chem. 395 (2009) 1355–1372.

145–152.

TE D

characterization and determination of masked mycotoxins, J. Chromatogr. A 1255 (2012)

β-resorcylic macrolide from Fusarium sp. 05ABR26., Chinese Chem. Lett. 19 (2008)

EP

1089–1092.

2'-hydroxyzearalanol: resorcyclic acid lactone derivatives from the marine-derived fungus

AC C

Penicillium sp., Chem. Pharm. Bull. 56(9) (2008) 1355-1356. Syldatk, M. Metzler, Identification of an aliphatic epoxide and the corresponding dihydrodiol as novel congeners of zearalenone in cultures of Fusarium graminearum, J. Agric. Food Chem. in press, (2010) DOI: 10.1021/jf1022498. Cheneval, Induction of rapid IL-1b mRNA degradation in THP-1 cells mediated through the AU-rich region in the 3'UTR by a radicicol analogue, Cytokine 8 (1996) 751-761. MacKenzieJ, Increased Mature Interleukin-1β (IL-1β) secretion from THP-1 Cells Induced by nigericin is a result of activation of p45 IL-1β-converting enzyme processing, Biol. Chem. 273 (1998) 17846-17851. A and related resorcylic acid lactones, potent kinase inhibitors, Angew. Chem. Int. Ed. 46 (2007) 61

ACCEPTED MANUSCRIPT[Benzenediol lactones] 1

6899-6902.

2 3

[305] G. Philipp, A.K. Gabriel, L.M. Andreas, Synthesis of Macrocyclic Scaffolds from Natural

4 5 6

[306] D. H. Williams, S. E. Wilkinson, T. Purton, A. Lamont, H. Flotow, E. J. Murray, Ro 09-2210

7 8 9

[307] Y. Hoshino, V.B. Ivanova, K. Yazawa, A. Ando, M. Yuzuru, Queenslandon, a new antifungal

10 11

[308] V. Navickas, M.E. Maier, Synthesis of the proposed structure of queenslandon, Tetrahedron 66

12 13 14

[309] F.M. Talontsi, P. Facey, M.D.K. Tatong, M.T. Islam, H. Frauendorf, S. Draeger, A. Tiedemann,

15 16 17 18

[310] K. Hanasaki, T. Kamigaito, JP 2000287697 A2 20001017 (Japanese). [311] M. Chinworrungsee, P. Kittakoop, M. Isaka, P. Maithip, S. Supothina, Y. Thebtaranonth,

Products and their Utilization for Solid-Phase, Chemistry Synthesis (2005) 2015-2021. exhibits potent anti-proliferative effects on activated T cells by selectively blocking MKK activity, Biochemistry 1998, 37, 9579-9585.

RI PT

compound produced by Chrysosporium queenslandicum: production, isolation and structure elucidation, J. Antibiot. 55 (2002) 516-519. (2010) 94–101.

SC

H. Laatsch, Zoosporicidal metabolites from an endophytic fungus Cryptosporiopsis sp. of Zanthoxylum leprieurii, Phytochemistry 83 (2012) 87–94.

M AN U

Isolation and structure elucidation of a novel antimalarial macrocyclic polylactone,

AC C

EP

TE D

menisporopsin A, from the fungus Menisporopsis theobromae, J. Nat. Prod. 67 (2004) 689-692.

62

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

· · · ·

Highlights The numerous biological activities of benzenediol lactones have been discussed. The chemical scaffolds of benzenediol lactones have been discussed. The natural sources of benzenediol lactones have been discussed. The structure activity relationships (SAR) of benzenediol lactones for different activities have been discussed in this review.