Stereochemical assignment of four diastereoisomers of a maculalactone derivative by computational NMR calculations

Stereochemical assignment of four diastereoisomers of a maculalactone derivative by computational NMR calculations

Accepted Manuscript Stereochemical assignment of four diastereoisomers of a maculalactone derivative by computational NMR calculations Daniel Previdi,...

2MB Sizes 0 Downloads 18 Views

Accepted Manuscript Stereochemical assignment of four diastereoisomers of a maculalactone derivative by computational NMR calculations Daniel Previdi, Viviani Nardini, Mayla Eduarda Rosa, Vinicius Palaretti, Gil Valdo José da Silva, Paulo Marcos Donate PII:

S0022-2860(18)31259-6

DOI:

https://doi.org/10.1016/j.molstruc.2018.10.064

Reference:

MOLSTR 25793

To appear in:

Journal of Molecular Structure

Received Date: 3 August 2018 Revised Date:

12 October 2018

Accepted Date: 19 October 2018

Please cite this article as: D. Previdi, V. Nardini, M.E. Rosa, V. Palaretti, G.V. José da Silva, P.M. Donate, Stereochemical assignment of four diastereoisomers of a maculalactone derivative by computational NMR calculations, Journal of Molecular Structure (2018), doi: https://doi.org/10.1016/ j.molstruc.2018.10.064. 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 1

Stereochemical assignment of four diastereoisomers of a maculalactone derivative by

2

computational NMR calculations

3 4

Authors: Daniel Previdi,* Viviani Nardini, Mayla Eduarda Rosa, Vinicius Palaretti, Gil

5

Valdo José da Silva and Paulo Marcos Donate

RI PT

6 7

Departamento de Química, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto,

8

Universidade de São Paulo, Avenida Bandeirantes, 3900, CEP 14040-901, Ribeirão Preto,

9

SP, Brazil.

*Corresponding author. E-mail address: [email protected] (D. Previdi).

SC

10 11

Abstract:

M AN U

12

Naturally occurring γ-butyrolactones and their synthetic analogues display a wide

14

range of bioactivities. Here, the multicomponent reaction of dimethyl 2-benzyl-3-

15

methylenesuccinate with bromobenzene and benzaldehyde catalyzed by cobalt(II) bromide

16

afforded a maculalactone derivative with three stereogenic centers. This reaction presented

17

moderate diastereoselectivity and yielded different proportions of all the four possible

18

diastereoisomers. The anti:anti (majority), anti:syn, syn:anti, and syn:syn diastereoisomers

19

were isolated and characterized by 1D and 2D NMR experiments. Because the stereochemical

20

assignment of all the diastereoisomers by Nuclear Overhauser Effect Difference (NOEDiff)

21

experiments was not definitive, the 1H and

22

theoretical calculations with the density functional theory at the B3LYP/6-31G(d) level. The

23

relative configurations of all the four diastereoisomers were assigned by using the CP3

24

parameter to compare the experimental and the calculated data and by determining the CP3

25

probability, which provided high level of confidence.

27

C NMR chemical shifts were predicted by

EP

13

AC C

26

TE D

13

Keywords: γ-butyrolactone, maculalactone, assignment, NMR, CP3.

28

1

ACCEPTED MANUSCRIPT 29

1. Introduction

30

1.1. γ-Butyrolactones The γ-butyrolactone ring exists in many natural products and synthetic analogues that

32

display diverse biological activities, such as antitumor,[1,2] antibacterial,[3] antiviral,[4] and

33

antiprotozoal[5] action, among other effects.[6–12] This skeleton occurs in numerous natural

34

products like mono-, di-, or tri-substituted monocyclic lactones, or it can be part of more

35

complex frameworks (Figure 1).[13] Maculalactones, which bear a benzylated γ-

36

butyrolactone core, are natural products that have been isolated from the marine

37

cyanobacterium Kyrtuthrix maculans.[14–20]

TE D

M AN U

SC

38

RI PT

31

39

Figure 1. Examples of natural γ-butyrolactones and some biological activities.

EP

40 41

Recently, Le Floch and co-workers developed a multicomponent reaction involving

AC C

42 43

dimethyl itaconate, an aryl halide, and a carbonyl compound catalyzed by cobalt(II) bromide

44

to synthesize paraconic acid analogues.[21,22] Our research group has employed microwave

45

irradiation to optimize this reaction [23,24] and used it to obtain maculalactone

46

derivatives.[25]

47

The single-step synthesis of maculalactone derivatives through this multicomponent

48

reaction affords γ-butyrolactones with three aromatic substituents and three stereogenic

49

centers (Scheme 1). All the reactions are moderately diastereoselective, and their products

50

consist of mixtures of all the four possible diastereoisomers as racemic mixtures. In almost all

51

cases, the major diastereoisomer is purified by crystallization and has the anti:anti relative 2

ACCEPTED MANUSCRIPT 52

configuration, as determined by NOEDiff NMR experiments.[25] To increase our level of

53

confidence in the stereochemical assignment, we isolated the four diastereoisomers, each one

54

as a racemic mixture, by preparative HPLC and analyzed them by NMR. Because the full

55

characterization of these compounds was not so clear-cut, we decided to calculate the 13C and

56

1

57

diastereoisomers.

RI PT

H NMR chemical shifts directly in order to assign the relative configurations of all the four

59 60

M AN U

SC

58

Scheme 1. Syntheses of maculalactone derivatives.

61 62 63

1.2. NMR chemical shift calculations and comparison parameters Aiming at the stereochemical assignment of molecules, computational calculations have been increasingly applied to predict NMR properties, including 1H and

65

chemical shifts and coupling constants.[26–30] This procedure has been employed to

66

characterize several natural products and synthetic compounds.[31–38]

TE D

64

13

C NMR

In 2002, Barone and co-workers used this technique in their benchmark studies based

68

on easy-to-calculate statistical correlation parameters, such as R2, the mean absolute error

69

(MAE), and the corrected mean absolute error (CMAE), to estimate confidence levels for the

70

assignments.[39,40] Recently, more refined procedures have been developed to compare

71

experimental and calculated NMR chemical shifts and to achieve higher levels of confidence

72

than those expected from simple statistical parameters.[41] Three main methods have been

73

designed: CP3,[42] DP4,[43,44] and ANN-PRA.[45,46]

AC C

EP

67

74

In 2009, Smith and Goodman developed three different comparison parameters,

75

designated CP1, CP2, and CP3. They validated these parameters with a set of 28 pairs of

76

diastereoisomers, and CP3 gave the best results.[42] CP3, calculated according to Equation 1,

77

provided the best match between two experimental spectra and the calculated chemical shifts

78

of two diastereoisomers. This parameter is based on the observation that, for similar atoms,

79

systematic errors in NMR chemical shift calculations can be removed by comparing the 3

ACCEPTED MANUSCRIPT 80

differences between two calculated chemical shifts and two experimental data. Cancelation of

81

the systematic errors indicates that these differences are calculated more accurately than the

82

chemical shifts themselves. CP3 has been applied in the stereochemical assignments of

83

various types of natural and synthetic compounds.[47–56]

CP3 =

∑ (∆ , ∆  )  ∑ ∆ where

RI PT

84

Equation 1

 ⁄ ∆ , ∆   = ∆ ∆  if ∆  ⁄∆ > 1

SC

∆ , ∆   = ∆ ∆  otherwise 85

CP3 was developed to solve the problem of attributing two sets of experimental data

87

(spectra X and Y) to two possible diastereomeric structures (structures x and y). As a

88

consequence, two assignments could be made: X = x and Y = y or X = y and Y = x.

89

The differences between the NMR chemical shifts calculated for similar atoms from

90

diastereomeric structures (∆δcalc, x – y) are compared to the differences between the

91

experimental NMR chemical shifts (∆δexp, X – Y), and CP3 is obtained according to Equation

92

1. A positive CP3 value indicates that the assignment is probably right (i.e. X = x and Y = y);

93

CP3 equal to one means perfect agreement between the experimental and calculated NMR

94

chemical shifts; and a negative CP3 value indicates that the assignment is probably wrong.

95

CP3 can be calculated for 1H and 13C NMR separately. The CP3 arithmetic mean for 1H and

96

13

EP

TE D

M AN U

86

C NMR, called “All Data”, usually affords better results for the assignments. Smith and Goodman (2009) also developed Equation 2 to calculate the probability that

98

the assignment is correct. Although CP3 is easy to compute “by hand”, its probability requires

99

knowledge of normal distribution descriptors (expectation value and standard deviation) and

100

statistical computer programs. The authors provided an applet to compute both CP3 and its

101

probability.[62]

AC C

97

102 (A |R and R  ) =

(A )(R |A )(R  |A ) (A )(R |A )(R  |A ) + (A )(R |A )(R  |A )

Equation 2

103 104

To obtain Equation 2, the authors used conditional probability elements and the Bayes

105

theorem. They assumed that the CP3 values for correct and wrong assignments are normally 4

ACCEPTED MANUSCRIPT 106

distributed, independent variables, and that expectation values and standard deviations can be

107

approximated on the basis of the values obtained for the original reference.[42] In Equation 2, P(A1|R1 and R2) is the probability that a given assignment is correct.

109

Figure 2 illustrates the probability elements in Equation 2. Two assignments (A1 and A2) are

110

possible. In the absence of any prediction about which assignment is right, P(A1) = P(A2) =

111

0.5. Therefore, two possibilities (R1 and R2) exist: that A1 is right, and consequently A2 is

112

wrong (R1), or that A2 is right, so A1 is wrong (R2). To calculate these probabilities, statistics

113

computer programs can be employed to compute the area above normal distribution curves.

114

The programs use the expectation values and standard deviations for CP3 for right and wrong

115

assignments.

SC

RI PT

108

TE D

M AN U

116

117 118

EP

Figure 2. Interpretation of probability elements in Equation 2.

119

In this study, we describe the assignment of the relative configurations of all the four

121

possible diastereoisomers of maculalactone derivative 4 (Scheme 2) by using a combination

122

of NMR experiments, DFT/GIAO chemical shift calculations, and CP3 computation.

123

AC C

120

124

2. Experimental section

125

2.1. General

126

γ-Butyrolactone 4 was prepared as described in our previous work [25]. A CEM

127

Discover® microwave reactor operating at maximum potency of 150 W was used. The

128

melting points of the synthesized compounds were obtained on a Bristolscope–871035

129

melting point apparatus and are uncorrected. IR spectra were acquired on a Shimadzu IR 5

ACCEPTED MANUSCRIPT 130

Prestige-21 spectrometer equipped with a germanium crystal; the thin solid film method and

131

the ATR (Attenuated Total Reflection) technique were employed. NMR experiments (1H, 13C,

132

NOEDiff, COSY, HMQC, and HMBC) were run on a Bruker DPX-300 spectrometer with

133

CDCl3 as solvent. The Chemical shifts are reported relative to TMS (0.00 ppm for 1H) and

134

CDCl3 (77.0 ppm for

135

conducted on a HPLC Shimadzu-LC-6AD with a preparative column Phenomenex® – Luna

136

C18 (250 x 21.2 mm, particle size: 5 µm, pore size: 100 Å, UV-Visible detector: 210 and 240

137

nm, loop: 1000 µL) and isocratic elution with acetonitrile (HPLC grade) and water (Milli-Q)

138

at a flow rate of 9.0 mL/min; the sample was diluted in acetonitrile at a concentration around

139

100 mg/mL for injection. Gas chromatography analyses were carried out on a GC-FID –

140

Shimadzu GC-2010 Plus with a Restek Rtx-5® column (film thickness: 0.25 µm, length: 30

141

m, and internal diameter: 0.25 mm) and nitrogen as carrier gas (linear velocity: 36.8 cm/s).

142

The heat program was 80 oC for 2 min, which was followed by a rise to 320 oC at 30 oC/min,

143

and 320 oC for 10 min. Electron impact (70 eV) mass spectra were registered on a Shimadzu

144

GC-MS-QP2010Plus under the same conditions describe above, except that the carrier gas

145

was helium.

C). High Performance Liquid Chromatography separations were

M AN U

SC

RI PT

13

146 2.2. Synthetic procedure

TE D

147

Compound 4 was synthesized as described previously (see Scheme 2).[25] A 25-mL

149

round-bottom flask under argon atmosphere was loaded with recently purified acetonitrile (5

150

mL),[63] dimethyl 2-benzyl-3-methylenesuccinate (3.1 g, 12.5 mmol), benzaldehyde (0.25

151

mL, 2.5 mmol), bromobenzene (0.42 mL, 4 mmol), and zinc dust (0.8 g, 12 mmol). This

152

mixture was briefly stirred at room temperature. Next, cobalt(II) bromide (0.13 g, 0.6 mmol),

153

trifluoroacetic acid (30 µL), and 1,2-dibromoethane (50 µL) were successively added to the

154

previous mixture, which was then irradiated in a CEM Discovery® focused microwave oven

155

at 60 °C and 150 W for 20 min. After that, the reaction mixture was filtered through Celite®,

156

which was washed several times with ethyl acetate. The organic fractions were combined and

157

concentrated under reduced pressure. The crude reaction product was purified by flash

158

column chromatography through silica gel (gradient elution with n-hexane/ethyl acetate from

159

9:1 to 6:4 (v/v)), to afford 0.8890 g of a diastereomeric mixture of maculalactone derivative 4

160

in 89% yield at a diastereomeric ratio (d.r., determined by GC-FID) of 73:14:8:5. The

161

diastereomeric mixture was separated by preparative HPLC. The relative configurations of all

AC C

EP

148

6

ACCEPTED MANUSCRIPT 162

the four diastereoisomers were determined by using a combination of NMR experiments,

163

DFT/GIAO chemical shift calculations, and CP3 computation.

164 Methyl 3,4-dibenzyl-5-oxo-2-phenyltetrahydrofuran-3-carboxylate (4a, anti:anti):

166

White solid, mp: 137–138 oC. 1H NMR (300 MHz, CDCl3), δ (ppm): 7.50 to 6.30 (m, 15H),

167

5.37 (s, 1H), 3.50 (s, 3H), 3.35 (dd, 1H, J = 14.9 Hz, J = 6.7 Hz), 3.23 (d, 1H, J = 15.4 Hz),

168

3.17 (d, 1H, J = 15.4 Hz), 3.03 (dd, 1H, J = 6.7 Hz, J = 5.0 Hz), 2.53 (dd, 1H, J = 14.9 Hz, J =

169

5.0 Hz).

170

134.8 (C), 131.3 (2 CH), 129.5 (2 CH), 129.1 (CH), 129.0 (2 CH), 128.8 (2 CH), 128.6 (2

171

CH), 127.7 (CH), 126.9 (CH), 126.0 (2 CH), 81.2 (CH), 61.7 (C), 52.0 (CH3), 47.7 (CH), 36.3

172

(CH2), 33.0 (CH2). IR (thin solid film, cm-1): 3063 (Csp2–H), 2950 (Csp3–H), 1771 and 1723

173

(C═O), 1218 and 1175 (C–O), 753 and 700 (═C─H). EI-MS (70 eV), m/z (relative

174

intensity, %): [M•+] 400 (<5), [M•+ – C7H6O] 294 (<5), [M•+ – C7H6O – C7H7•] 203 (100),

175

[M•+ – C7H6O – C7H7• – C4H4O2 – CO] 91 (72), [M•+ – C9H8O2] 252 (50), [M•+ – C9H8O2 –

176

MeOH] 220 (33).

C NMR (75 MHz, CDCl3), δ (ppm): 176.1 (C), 170.9 (C), 139.0 (C), 135.1 (C),

M AN U

SC

13

RI PT

165

177

Methyl 3,4-dibenzyl-5-oxo-2-phenyltetrahydrofuran-3-carboxylate (4b, anti:syn):

179

White solid, mp: 172–173 oC. 1H NMR (300 MHz, CDCl3), δ (ppm): 7.35 to 6.88 (m, 15H),

180

5.77 (s, 1H), 3.73 (s, 3H), 3.20 (dd, 1H, J = 7.5 Hz, J = 4.7 Hz), 3.10 (dd, 1H, J = 14.4 Hz, J =

181

7.5 Hz), 2.83 (d, 1H, J = 14.5 Hz), 2.57 (dd, 1H, J = 14.4 Hz, J = 4.7 Hz), 2.48 (d, 1H,

182

J = 14.5 Hz). 13C NMR (75 MHz, CDCl3), δ (ppm): 176.4 (C), 172.6 (C), 138.2 (C), 135.9

183

(C), 134.9 (C), 130.4 (2 CH), 129.2 (2 CH), 128.9 (CH), 128.6 (2 CH), 128.5 (2 CH), 128.4

184

(2 CH), 127.1 (CH), 127.0 (2 CH), 126.8 (CH), 83.9 (CH), 58.3 (C), 52.4 (CH3), 48.5 (CH),

185

39.5 (CH2), 33.5 (CH2). IR (thin solid film, cm-1): 3059 (Csp2–H), 2985 (Csp3–H), 1780 and

186

1734 (C═O), 1216 and 1124 (C–O), 756 and 699 (═C─H). EI-MS (70 eV), m/z (relative

187

intensity, %): [M•+] 400 (<5), [M•+ – C7H6O] 294 (<5), [M•+ – C7H6O – C7H7•] 203 (100),

188

[M•+ – C7H6O – C7H7• – C4H4O2 – CO] 91 (95), [M•+ – C9H8O2] 252 (92), [M•+ – C9H8O2 –

189

MeOH] 220 (63).

AC C

EP

TE D

178

190 191

Methyl 3,4-dibenzyl-5-oxo-2-phenyltetrahydrofuran-3-carboxylate (4c, syn:anti):

192

White solid, mp: 86–89 oC. 1H NMR (300 MHz, CDCl3), δ (ppm): 7.37 to 7.08 (m, 15H),

193

5.43 (s, 1H), 3.66 (dd, 1H, J = 7.4 Hz, J = 6.2 Hz), 3.43 (d, 1H, J = 15.2 Hz), 3.35 (d, 1H, J =

194

15.2 Hz), 3.19 (s, 3H), 2.99 (dd, 1H, J = 14.0 Hz, J = 7.4 Hz), 2.94 (dd, 1H, J = 14. Hz, J = 7

ACCEPTED MANUSCRIPT 13

195

6.2 Hz).

196

135.1 (C), 129.8 (2 CH), 129.3 (2 CH), 129.0 (CH), 128.8 (2 CH), 128.6 (2 CH), 128.5 (2

197

CH), 127.3 (CH), 126.9 (CH), 125.9 (2 CH), 83.6 (CH), 59.5 (C), 52.0 (CH3), 47. (CH), 37.2

198

(CH2), 31.9 (CH2). IR (thin solid film, cm-1): 3060 (Csp2–H), 2951 (Csp3–H), 1780 and 1706

199

(C═O), 1219 and 1158 (C–O), 745 and 695 (═C─H). EI-MS (70 eV), m/z (relative

200

intensity, %): [M•+] 400 (<5), [M•+ – C7H6O] 294 (<5), [M•+ – C7H6O – C7H7•] 203 (100),

201

[M•+ – C7H6O – C7H7• – C4H4O2 – CO] 91 (79), [M•+ – C9H8O2] 252 (68), [M•+ – C9H8O2 –

202

MeOH] 220 (50).

RI PT

C NMR (75 MHz, CDCl3), δ (ppm): 176.2 (C), 171.5 (C), 138.3 (C), 136.1 (C),

203

Methyl 3,4-dibenzyl-5-oxo-2-phenyltetrahydrofuran-3-carboxylate (4d, syn:syn):

205

White solid, mp: 98–101oC. 1H NMR (300 MHz, CDCl3), δ (ppm): 7.40 to 6.80 (m, 15H),

206

5.71 (s, 1H), 3.71 (s, 3H), 3.62 (dd, 1H, J = 8.2 Hz, J = 4.4 Hz), 3.01 (d, 1H, J = 15.0 Hz),

207

2.97 (dd, 1H, J = 14.2 Hz, J = 8.2 Hz), 2.92 (d, 1H, J = 15.0 Hz), 2.74 (dd, 1H, J = 14.2 Hz,

208

J = 4.4 Hz).

209

(C), 134.1 (C), 130.0 (2 CH), 129.1 (2 CH), 128.8 (CH), 128.6 (2 CH), 128.4 (2 CH), 128.3 (2

210

CH), 127.1 (2 CH), 126.8 (CH), 126.6 (CH), 83.7 (CH), 59.1 (C), 52.6 (CH3), 52.5 (CH), 34.4

211

(CH2), 32.2 (CH2). IR (thin solid film, cm-1): 3061 (Csp2–H), 2943 (Csp3–H), 1778 and 1730

212

(C═O), 1224 and 1169 (C–O), 744 and 696 (═C─H). EI-MS (70 eV), m/z (relative

213

intensity, %): [M•+] 400 (<5), [M•+ – C7H6O] 294 (<5), [M•+ – C7H6O – C7H7•] 203 (100),

214

[M•+ – C7H6O – C7H7• – C4H4O2 – CO] 91 (63), [M•+ – C9H8O2] 252 (52), [M•+ – C9H8O2 –

215

MeOH] 220 (33).

TE D

C NMR (75 MHz, CDCl3), δ (ppm): 175.3 (C), 172.7 (C), 139.6 (C), 136.1

AC C

EP

13

M AN U

SC

204

8

ACCEPTED MANUSCRIPT 216

2.3. Computational studies

217

Conformational searches and geometry optimizations followed practically the same

218

conditions and parameters as the conditions and parameters used in our previous

219

works.[64,65] Conformational searches were performed with the PCModel program version

220

7.0 [66] by using the GMMX routine with the parameters depicted in Table 1.

222

RI PT

221

Table 1. Parameters used during conformational searches with the PCModel program. Parameter

Value

Force field

MMX

SC

First cycle: 1.5 kcal/mol

Energy window

Second cycle: 1.0 kcal/mol

M AN U

Minimum energy found x times

5

Maximum conformation minimized

100,000

Boltzmann temperature

300 K

223

The geometry of each conformer in the energy window of the conformational search

225

was optimized with ORCA version 3.0.1 [67] in vacuum, at the B3LYP-D3(BJ)/def2-TZVP(-

226

f) level [68–73], in addition to the grid4 keyword and the RIJCOSX approximation.[74]

227

Dispersion interaction (D.I.) between atoms, which represents the energy of the van der Waals

228

forces, was taken into account by following the approach described by Grimme and

229

coworkers.[75–77]

EP

TE D

224

After the geometry was optimized, conformers amounting to 90% of the Boltzmann

231

distribution were selected. The number of conformers obtained for the four diastereoisomers

232

are shown in Table 2.

233

AC C

230

Isotropic magnetic shielding (IMS) was calculated with the GIAO method at the

234

B3LYP/6-31G(d) level by using Gaussian03.[78] The calculated chemical shifts were

235

obtained as the difference between the IMS of a particular 1H or 13C in a conformer and the

236

IMS of 1H or 13C in TMS, calculated by the same method.

237

Finally, CP3 was determined according to Equation 1 for all the possible assignment

238

combinations. Probabilities were obtained by using Equation 2 as well as the expectation

239

values and standard deviations from the original reference listed in Table 3.[42]

240

9

ACCEPTED MANUSCRIPT 241

Table 2. Number of conformers obtained during the conformational search with molecular

242

mechanics and after optimization geometry by DFT. Calculated diastereoisomer

Conformational search

4a anti:anti

34

4b anti:syn

41

4c syn:anti

50

4d syn:syn

16

Selected conformers after optimization*

RI PT

12 16 7

7

* Number of conformers amounting to 90% of the population according to Boltzmann

244

distribution.

SC

243

245

Table 3. Expectation values and standard deviations for CP3. 1

Assignment

13

C

All data

Right

0.478 ± 0.305

0.547 ± 0.253

0.512 ± 0.209

Wrong

-0.786 ± 0.835

-0.487 ± 0.533

-0.637 ± 0.499

247 3. Results and discussion

TE D

248

H

M AN U

246

We conducted the cobalt(II) bromide-catalyzed multicomponent reaction of dimethyl

250

2-benzyl-3-methylenesuccinate (1) with bromobenzene (2) and benzaldehyde under

251

microwave irradiation, to obtain maculalactone derivative 4 as a mixture of four

252

diastereoisomers as racemic mixtures at a diastereomeric ratio of 73:14:8:5, determined by

253

GC-FID analysis, and relative configurations anti:anti, anti:syn, syn:anti, and syn:syn.[25]

254

We purified the diastereoisomers by preparative HPLC. The NMR spectral data (1H,

255

DEPT-135, COSY, HMQC, HMBC, and NOEDiff) obtained for the four fractions did not

256

allow a clear-cut assignment of each spectrum corresponding to structures 4a-4d. Therefore,

257

we had to work out which spectrum referred to a given diastereoisomer (see Figures 3-4 and

258

supporting information for all the spectral data).

13

C,

AC C

EP

249

259

10

ACCEPTED MANUSCRIPT

O MeO

+

O

1

ACN Zn, CoBr2

O

Br

OMe

+

H

2

O

TFA, DBE MW, 150W 20 min

O

MeO2C

3

RI PT

4 89% (73:14:8:5)

11 12

10

13

9

24 23

22

O

2 1 O 3 4 14 15 20 16 19 17 18

MeO2C

4a anti:anti

O

O

SC

5 6 MeO2C 25 26 21

O

MeO2C

M AN U

8 7

4b anti:syn

O

4c syn:anti

MeO2C

O O

4d syn:syn

260 261

Scheme 2. Relative configurations of all the possible diastereoisomers of compound 4.*

262

*Diastereoisomers 4a-d were obtained as racemic mixtures and the stereochemistry showed in

263

the structures of these compounds are the relative configurations between then.

AC C

EP

TE D

264

265 266

Figure 3. 1H NMR (300 MHz, CDCl3) spectra of the diastereoisomers of compound 4. 11

267 268

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 4. 13C NMR (75 MHz, CDCl3) spectra of the diastereoisomers of compound 4.

269

The diastereoselectivity of this multicomponent reaction is probably due to the

271

mechanism of reaction and epimerization equilibrium. Scheme 3 presents a proposed

272

mechanism for this multicomponent reaction. The reaction starts with an organometallic

273

compound originating from bromobenzene (2). The organometallic compound then reacts

274

with itaconate derivative 1, to produce enolate 5. In turn, this enolate reacts with

275

benzaldehyde (3) through cyclic transition state 6, to yield intermediate 7. The latter

276

intermediate is in equilibrium with intermediate 8, thanks to exchange of the carboxylate

277

group chelating the metal. Finally, a lactonization process gives maculalactone derivative

278

4.[21]

EP

AC C

279

TE D

270

12

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Scheme 3. Mechanism proposed for the multicomponent reaction to obtain maculalactone

282

derivative 4.

283

EP

280 281

The aldol reaction, during which cyclic transition state 6 acquires a chair

285

conformation, probably underlies the diastereoselectivity at carbons 3 and 4.[79,80] Enolate 5

286

can have stereochemistry E/Z (5a and 5b in Scheme 4). Under the reaction conditions (i.e.,

287

temperature = 60 oC), the aldol reaction product can be determined by thermodynamic

288

control.[81] As shown in Scheme 4, enolate E (5a) has transition state 6a with the large

289

substituent R1 in the equatorial position and furnishes intermediate 7a, which has relative

290

configuration anti between groups R2 and R3. The opposite is observed for enolate Z (5b): it

291

has transition state 6b with the large substituent R1 in the axial position and furnishes

292

intermediate 7b, which has relative configuration syn between groups R2 and R3. Saturated

293

six-membered rings are known to exist in the chair/boat conformation. Because the largest

294

substituent is known to be more stable in the equatorial position of the chair conformation,

295

relative configuration anti is probably favored at carbons 3 and 4.

AC C

284

13

ACCEPTED MANUSCRIPT 296 O

M

R1 Ph

R2

MeO2C Ph

Ph

Ph

5a enolate E O

Ph Ph

Ph

297 298

Ph

CO2Me

5b enolate Z

R3 M

4 O R2

7a 3,4-anti

R H Ph 3

O

OMe H O M O Ph

MeO 5 Ph R2 Ph

CO2Me

6b

3

Ph 2 CO2Me 1

Ph

6a

M

H Ph

RI PT

MeO

MeO 5

SC

MeO

O

OMe H M O O Ph R3

3

4 O

2

M

CH2O2Me 1

7b 3,4-syn

Scheme 4. Origin of diastereoselectivity at carbons 3 and 4 of intermediate 7.

M AN U

299 300

After conversion of intermediate 7 to 8 and subsequent lactonization to maculalactone

301

derivative 4 (Scheme 3), an epimerization process may occur. This process is catalyzed by

302

bases like methoxide, which is generated during the lactonization reaction and is favored by

303

excess organometallic species (due to excess halide).[82,83]

Epimerization possibly accounts for the diastereoselectivity at carbons 2 and 3

305

(Scheme 5). Diastereoisomer 4c, which possesses relative configuration syn at carbons 2 and

306

3, can be converted to diastereoisomer 4a, which has relative configuration anti at carbons 2

307

and 3 and also a minor steric hindrance at the aromatic substituents on carbons 2 and 3. The

308

same epimerization equilibrium can occur with diastereoisomers 4d and 4b, as shown in

309

Scheme 5.

AC C

EP

TE D

304

14

310 311

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Scheme 5. Epimerization equilibrium between the diastereoisomers of compound 4.

312

In accordance with the considerations above, we expected that the major

314

diastereoisomer of maculalactone derivative 4 would be diastereoisomer 4a (anti:anti),

315

whereas the minor diastereoisomer of maculalactone derivative 4 would be diastereoisomer

316

4d (syn:syn). The second major diastereoisomer should be diastereoisomer 4b (anti:syn),

317

followed by diastereoisomer 4c (syn:anti). Therefore, we proposed that spectra A, B, C, and D

318

should correspond to diastereoisomers 4a, 4b, 4c, and 4d, respectively.

TE D

313

To determine the relative configurations of all the four diastereoisomers of compound

320

4 isolated by preparative HPLC, we first obtained the NOEDiff NMR spectra. On the basis of

321

these NMR experiments, we deduced that the major diastereoisomer obtained during the

322

multicomponent

323

diastereoisomer possesses syn:syn relative configuration. The relative configurations of the

324

other two diastereoisomers are not clear-cut. See supporting information for the NOEDiff

325

NMR spectra and more details.

AC C

EP

319

reaction

has

anti:anti

relative

configuration,

whilst

the

minor

326

To conclude these assignments and to increase our confidence level about the relative

327

configurations of the four diastereoisomers obtained during the multicomponent reaction, we

328

decided to compare the experimental and calculated 1H and 13C NMR data by using the CP3

329

parameter.[42] Simpler statistical parameters like R2, MAE, and CMAE cannot distinguish

330

between the four diastereoisomers. See supporting information for more details. 15

ACCEPTED MANUSCRIPT We calculated the 1H and

331

13

C NMR chemical shifts according to the procedure

described in the experimental section. Tables 4 and 5 summarize the experimental and

333

calculated 1H and 13C NMR chemical shifts, respectively. These tables only present the atoms

334

that we used to calculate CP3. We did not employ aromatic hydrogens because their signals

335

are heavily overlapped. We removed carbons 10-12, 16-18, and 23-25 because we were not

336

able to attribute them experimentally on the basis of 2D NMR experiments.

337

Table 4. Experimental and calculated 1H NMR chemical shifts used to calculate CP3.

A

B

C

D

4

5.37

5.77

5.43

5.71

6

3.50

3.73

3.19

3.71

7A

3.35

3.10

2.99

20A

3.23

2.83

20B

3.17

2 7B

Calculated diastereoisomer 4a

4b

4c

4d

6.21

6.45

6.21

6.34

4.29

4.10

3.32

4.48

2.97

4.30

4.02

4.26

3.91

3.43

3.01

3.87

3.32

4.71

3.49

2.48

3.35

2.92

3.79

3.12

3.51

3.48

3.03

3.20

3.66

3.62

3.60

4.16

4.64

4.63

2.53

2.57

2.94

2.74

3.59

4.00

3.55

3.71

TE D EP AC C

339

Experimental spectrum

SC

H

M AN U

338

RI PT

332

16

ACCEPTED MANUSCRIPT Table 5. Experimental and calculated 13C NMR chemical shifts used to calculate CP3. C

Calculated diastereoisomer

B

C

D

4a

4b

4c

4d

7

33.0

33.5

31.9

32.2

35.8

36.0

33.4

33.4

20

36.3

39.5

37.2

34.4

38.0

40.8

40.1

37.1

2

47.7

48.5

47.8

52.5

48.7

49.4

51.0

54.7

6

52.0

52.4

52,0

52.6

51.7

51.1

51.3

52.1

3

61.7

58.3

59.5

59.1

63.0

58.8

59.4

61.4

4

81.2

83.9

83.6

83.7

81.3

82.7

79.7

83.3

15

126.0

127.0

125.9

127.1

19

126.0

127.0

125.9

127.1

9

129.5

129.2

129.3

129.1

13

129.5

129.2

129.3

22

131.4

130.4

26

131.4

21

SC

RI PT

A

120.7

118.7

120.5

120.9

120.8

118.7

120.9

123.5

123.2

122.9

124.2

129.1

123.8

123.4

122.9

122.1

129.8

130,0

126.0

125.3

125.5

124.2

130.4

129.8

130,0

124.6

125.2

125.8

126.4

134.8

135.9

136.1

136.1

129.9

129.7

131.8

128.3

14

135.1

134.9

135.1

134.1

129.8

130.7

130.8

132.2

8

139.0

138.2

138.3

139.6

134.2

130.9

132.6

132.7

5

170.9

172.6

171.5

172.7

163.5

165.2

163.7

164.4

1

176.1

176.4

176.2

175.3

165.8

166.6

166.0

165.2

M AN U

120.6

EP

341

Experimental spectrum

TE D

340

CP3 was developed to compare two sets of experimental data to two possible

343

diastereomeric structures, so just two assignments should be made. In our case, we had four

344

NMR spectral datasets to compare to four possible diastereomeric structures, so the number of

345

possible assignments was much larger—actually, 144 assignments were possible.

346

AC C

342

We conducted all these assignments and computed CP3 and its probability for the 1H

347

and 13C NMR data together (“All Data”). Table 6 shows these results in a simple manner. A

348

green check indicates that CP3 is positive and its probability is high (> 95%); a yellow check

349

indicates that CP3 is positive and its probability is low, or that CP3 is negative and its

350

probability is high; and a red “X” indicates that CP3 is negative and its probability is low. For

351

more details about Table 6, see supporting information.

352

Table 6 also contains all the possible assignments. For example, the first line shows

353

the assignments attributed to diastereoisomer 4a in spectrum A and to the other 17

ACCEPTED MANUSCRIPT 354

diastereoisomers (diastereoisomers 4b, 4c and 4d) in spectra B, C, and D. This table is

355

reflected over the diagonal, which means that the first line is equivalent to the first column.

356

The last three columns are the sum of the green check, the yellow check, and the red “X”. If we analyze Table 6 over the four lines for a single spectrum; i.e., the first four lines

358

of spectrum A, for example, the line that presents the largest numbers of green and yellow

359

checks indicates that this spectrum corresponds to the diastereoisomer in question. For

360

spectrum A, the first line (diastereoisomer 4a) is the line with the largest numbers of green

361

and yellow checks (total of 9), indicating that the assignment of spectrum A to

362

diastereoisomer 4a is probably right. Another important observation is that when the other

363

spectra are attributed to diastereoisomer 4a (look at lines 2, 3, and 4 and columns 5, 9, and 13

364

for spectra B, C, and D, respectively), all the assignments present a red “X”, indicating that

365

CP3 is negative, and that the probability of the assignment is low. This confirms that the

366

former assignment is right, spectrum A really refers to diastereoisomer 4a. Hence, the best

367

assignments are spectra A, B, C, and D corresponding to diastereoisomers 4a (anti:anti), 4b

368

(anti:syn), 4c (syn:anti), and 4d (syn:syn), respectively.

M AN U

SC

RI PT

357

CP3 could provide “false positives” when the assignment is made for two spectra with

370

two diastereoisomers but just one spectrum corresponds to one of the two diastereoisomers

371

and has been attributed to the right one. For example, spectra A and C are attributed to

372

diastereoisomers 4a and 4b, respectively. This assignment furnishes a positive value for CP3,

373

and the probability that this assignment is right is high (green check). Because “false

374

positives” may happen, we must look over all the possible assignments and find out which are

375

the best.

EP

AC C

376

TE D

369

18

ACCEPTED MANUSCRIPT

377

Table 6. Possible assignments of spectra A, B, C, and D to diastereoisomers 4a, 4b, 4c, and 4d. Spectrum Diastereoisomer

4a

4b

B 4c

4d

4a

4b

C 4c

4d

4a

4b

4c

4d

 

  

4a

Sum

4b

4c

4d









 

  

7 0 2 1 1 8 2 2 2 1 5 0 1 1 2 8

2 3 2 1 0 1 0 2 2 1 3 3 1 3 0 1

0 6 5 7 8 0 7 5 5 7 1 6 7 5 7 0

378

4a    4b    A 4c    4d    4a    4b    B 4c    4d    4a       4b       C 4c       4d       4a       4b       D 4c       4d        Green check indicates that CP3 is positive and its probability is high.

379

 Yellow check indicates that CP3 value is positive and its probability is low, or that CP3 is negative and its probability is high.

380

 Red “X” indicates that CP3 is negative and its probability is low.

TE D

 

   

  

SC

M AN U

  

  

 

  



        

 

       

      

     

  



AC C

EP



Spectrum

381

  



D

RI PT

A

19

ACCEPTED MANUSCRIPT

382

Finally, we detail the assignments that are necessary and enough to conduct the

383

stereochemical assignments of all the four possible diastereoisomers of maculalactone derivative 4.

384

Table 7 gives these assignments and the CP3 values for the 1H and

385

correct assignments (A = 4a, B = 4b, C = 4c, and D = 4d) yield positive values for CP3, whereas

386

wrong assignments furnish negative values for CP3.

13

C NMR data. Presumably

Table 7. CP3 values for the 1H and 13C NMR data. CP3

Assignment

13

H

C

1

H and 13C

0.74

0.65

A = 4b and B = 4a

-0.98

-0.86

A = 4a and C = 4c

0.26

0.03

0.15

A = 4c and C = 4a

-1.27

-0.62

-0.94

A = 4a and D = 4d

0.65

A = 4d and D = 4a

-1.27

B = 4b and C = 4c

0.42

B = 4c and C = 4b

-0.95

B = 4b and D = 4d

0.69

B = 4d and D = 4b

-0.86

0.70

-0.92

0.47

0.56

-0.81

-1.04

0.30

0.36

-0.93

-0.94

0.67

0.68

-1.03

-0.94

0.26

0.66

0.46

-1.52

-0.84

-1.18

TE D

C = 4d and D = 4c

SC

A = 4a and B = 4b

C = 4c and D = 4d

389

1

M AN U

388

RI PT

387

The graphs in Figure 5 correspond to the plots of the CP3 values obtained for the 1H and 13C

391

NMR data for the assignments listed in Table 7. The horizontal lines represent one, two, and three

392

standard deviations from the expected CP3 value when the assignment is correct. For all the

393

assignments that give negative values, these values are below three standard deviations. This means

394

that element P(R2|A2) in Equation 2 will furnish a value that is smaller than 0.05, which is typical of

395

normal distribution. Actually, this value is usually responsible for the high probabilities obtained

396

with CP3. Mathematically speaking, when element P(R2|A2) in Equation 2 has value near zero, the

397

probability that a given assignment is right is close to one or 100%. Indeed, all the assignments

398

shown in Table 7 and presumed right present probabilities of 100.0%, so they are the correct ones.

AC C

EP

390

399

20

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

400 401

Figure 5. Plots of CP3 values obtained for the 1H and

402

shown in Table 7.

403

13

C NMR data regarding the assignments

As an example, figure 6 illustrates the CP3 parameter values obtained. It was generated by

405

simulating spectra with the chemical shifts obtained from the experimental and calculated 1H NMR

406

data from spectra A and B, and the diastereoisomers 4a and 4b that were used to compute the CP3

407

parameter.

TE D

404

It is possible to note that the differences in between chemical shifts (∆δ) for similar atoms in

409

calculated spectra for diastereoisomers 4a and 4b are better matched with the differences in

410

experimental chemical shifts for the assignments A = 4a and B = 4b, indicating that a positive value

411

for CP3 parameter was expected for this assignment and a negative value for the opposite one. For

412

example, the difference of chemical shifts for atom H4 in calculated spectra of diastereoisomer 4a

413

and 4b is – 0.24 (∆δcalc = 6.21 – 6.45; Table 4) and the difference for chemical shifts of atom H4

414

between experimental spectra A and B is – 0.40 (∆δexp = 5.37 – 5.77; Table 4) which contribute

415

with a positive increment for the CP3 parameter value because both differences of chemical shifts

416

possess the same algebraic sign. On the other hand, the difference of chemical shifts for atom H6 in

417

calculated spectra of diastereoisomer 4a and 4b is 0.19 (∆δcalc = 4.29 – 4.10; Table 4) and the

418

difference for chemical shifts of atom H6 between experimental spectra A and B is – 0.23 (∆δexp =

419

3.50 – 3.73; Table 4) which contribute with a negative increment for the CP3 parameter value. In

420

fact, the only difference between algebraic signs of ∆δcalc and ∆δexp in assignments A = 4a and B =

AC C

EP

408

21

ACCEPTED MANUSCRIPT

421

4b is for atom H6. The overall evaluation of chemical shift differences of corresponding atoms

422

between the two calculated spectra compared to the two experimental ones leads to a positive value

423

of which indicates that the CP3 parameter has a positive value (CP3 = 0.74) and the assignments in

424

question are correct. The opposite assignments, A = 4b and B = 4a, has a negative value for the

425

CP3 parameter value (CP3 = – 0.98) indicating that these assignments are wrong.

TE D

M AN U

SC

RI PT

426

Figure 6. Differences between the chemical shifts for spectra A and B and for diastereoisomers 4a

429

and 4b.

AC C

430

EP

427 428

431

Figure 6 also illustrate how the systematic errors in NMR chemical shift calculations can be

432

removed by comparing the differences between two calculated chemical shifts and two

433

experimental data, indicating that these differences are calculated more accurately than the chemical

434

shifts themselves.

435 436

4. Conclusion

437

Assignment of the relative configurations of the four diastereoisomers of maculalactone

438

derivative 4, initially deduced on the basis of the mechanism proposed for the multicomponent

439

reaction, was confirmed through the use of the CP3 parameter. This parameter provides high level 22

ACCEPTED MANUSCRIPT

440

of confidence as revealed by the calculated CP3 probability. In summary, DFT/GIAO calculations

441

of 1H and 13C NMR chemical shifts can be used in combination with CP3 to compare experimental

442

and calculated NMR data for complete assignment of all the four possible diastereoisomers of a γ-

443

butyrolactone ring with three stereogenic centers.

444 445

Acknowledgments The authors thank Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP,

447

grants 2013/18254-9, 2015/05454-5 and 2016/04896-7), Coordenação de Aperfeiçoamento de

448

Pessoal de Nível Superior - Brasil (CAPES) - Finance Code 001, and Conselho Nacional de

449

Desenvolvimento Científico e Tecnológico (CNPq) for financial support and fellowships.

RI PT

446

AC C

EP

TE D

M AN U

SC

450

23

ACCEPTED MANUSCRIPT

5. References

452

[1] C. Le Floch, E. Le Gall, S. Sengmany, P. Renevret, E. Léonel, T. Martens, T. Cresteil, Synthesis

453

of 2,3-di- and 2,2,3-trisubstituted-3-methoxycarbonyl-γ-butyrolactones as potent antitumor agents,

454

Eur. J. Med. Chem. 89 (2015) 654–670. doi:10.1016/j.ejmech.2014.10.074.

455

[2] C. Le Floch, E. Le Gall, E. Léonel, T. Martens, T. Cresteil, Synthesis and cytotoxic evaluation

456

of novel paraconic acid analogs, Bioorg. Med. Chem. Lett. 21 (2011) 7054–7058.

457

doi:10.1016/j.bmcl.2011.09.092.

458

[3] H. Yokoe, M. Yoshida, K. Shishido, Total synthesis of (−)-xanthatin, Tetrahedron Lett. 49

459

(2008) 3504–3506. doi:10.1016/j.tetlet.2008.03.081.

460

[4] K. Hayashi, K. Narutaki, Y. Nagaoka, T. Hayashi, S. Uesato, Therapeutic effect of arctiin and

461

arctigenin in immunocompetent and immunocompromised mice infected with influenza A virus,

462

Biol. Pharm. Bull. 33 (2010) 1199–1205. doi:10.1248/bpb.33.1199.

463

[5] R. da Silva, J. Saraiva, S. Albuquerque, C. Curti, P. Donate, T. Bianco, J. Bastos, M. Silva,

464

Trypanocidal structure–activity relationship for cis- and trans-methylpluviatolide, Phytochemistry.

465

69 (2008) 1890–1894. doi:10.1016/j.phytochem.2008.04.002.

466

[6] L. Barbosa, R. Furtado, H. Bertanha, I. Tomazella, E. Costa, J. Bastos, M. Silva, D. Tavares,

467

Chemopreventive effects of (−)-hinokinin against 1,2-dimethylhydrazine – induced genotoxicity

468

and preneoplastic lesions in rat colon, J. Nat. Prod. 77 (2014) 2312–2315. doi:10.1021/np500093u.

469

[7] R. Kitson, A. Millemaggi, R. Taylor, The renaissance of α-methylene-γ-butyrolactones: new

470

synthetic approaches, Angew. Chemie Int. Ed. 48 (2009) 9426–9451. doi:10.1002/anie.200903108.

471

[8] J. Saraiva, C. Vega, M. Rolon, R. da Silva, M. Silva, P. Donate, J. Bastos, A. Gomez-Barrio, S.

472

de Albuquerque, In vitro and in vivo activity of lignan lactones derivatives against Trypanosoma

473

cruzi, Parasitol. Res. 100 (2007) 791–795. doi:10.1007/s00436-006-0327-4.

474

[9] R. da Silva, G. Souza, A. da Silva, V. de Souza, A. Pereira, V. Royo, M. Silva, P. Donate, A.

475

Araújo, J. Carvalho, J. Bastos, Synthesis and biological activity evaluation of lignan lactones

476

derived

477

doi:10.1016/j.bmcl.2004.12.035.

478

[10] V. de Souza, R. da Silva, A. Pereira, V. Royo, J. Saraiva, M. Montanheiro, G. de Souza, A.

479

Filho, M. Grando, P. Donate, J. Bastos, S. Albuquerque, M. Silva, Trypanocidal activity of (−)-

480

cubebin derivatives against free amastigote forms of Trypanosoma cruzi, Bioorg. Med. Chem. Lett.

481

15 (2005) 303–307. doi:10.1016/j.bmcl.2004.10.079.

482

[11] M. Hughes, J. McFadden, C. Townsend, New α-methylene-γ-butyrolactones with

483

antimycobacterial

484

doi:10.1016/j.bmcl.2005.05.119.

AC C

EP

TE D

M AN U

SC

RI PT

451

from

(–)-cubebin,

properties,

Bioorg.

Bioorg.

Med.

Med.

Chem.

Chem.

Lett.

Lett.

15

15

(2005)

(2005)

1033–1037.

3857–3859.

24

ACCEPTED MANUSCRIPT

485

[12] A. Picman, Biological activities of sesquiterpene lactones, Biochem. Syst. Ecol. 14 (1986)

486

255–281. doi:10.1016/0305-1978(86)90101-8.

487

[13] M. Seitz, O. Reiser, Synthetic approaches towards structurally diverse γ-butyrolactone natural-

488

product-like

489

doi:10.1016/j.cbpa.2005.03.005.

490

[14] S. Bader, M. Luescher, K. Gademann, Synthesis of maculalactone A and derivatives for

491

environmental

492

doi:10.1039/C4OB02042A.

493

[15] A. Kar, S. Gogoi, N. Argade, Synthesis of naturally occurring bioactive butyrolactones:

494

maculalactones

495

doi:10.1016/j.tet.2005.03.065.

496

[16] G. Brown, H. Wong, N. Hutchinson, S. Lee, B. Chan, G. Williams, Chemistry and biology of

497

maculalactone A from the marine cyanobacterium Kyrtuthrix maculans, Phytochem. Rev. 3 (2004)

498

381–400. doi:10.1007/s11101-004-6552-5.

499

[17] H. Wong, G. Williams, G. Brown, Maculalactone M from the marine cyanobacterium

500

Kyrtuthrix maculans, Phytochemistry. 60 (2002) 425–429. doi:10.1016/S0031-9422(02)00138-3.

501

[18] S. Lee, G. Williams, G. Brown, Maculalactone L and three halogenated carbazole alkaloids

502

from

503

9422(99)00226-5.

504

[19]

505

tetrahydrobenzofuranones from Kyrtuthrix maculans, J. Nat. Prod. 61 (1998) 29–33.

506

doi:10.1021/np970322p.

507

[20] W. Tsui, G. Williams, G. Brown, A tribenzylbutyrolactone from Kyrtuthrix maculans,

508

Phytochemistry. 43 (1996) 1083–1085. doi:10.1016/S0031-9422(96)00296-8.

509

[21] C. Le Floch, E. Le Gall, E. Léonel, J. Koubaa, T. Martens, P. Retailleau, A cobalt-catalyzed

510

multicomponent approach to novel 2,3-di- and 2,2,3-trisubstituted 3-methoxycarbonyl-γ-

511

butyrolactones, Eur. J. Org. Chem. 2010 (2010) 5279–5286. doi:10.1002/ejoc.201000698.

512

[22] C. Le Floch, C. Bughin, E. Le Gall, E. Léonel, T. Martens, Three-component synthesis of

513

functionalized five-membered ring lactones under Barbier-like conditions, Tetrahedron Lett. 50

514

(2009) 5456–5458. doi:10.1016/j.tetlet.2009.07.038.

515

[23] M. Pinatto-Botelho, A. Crotti, J. de Souza, L. Magalhães, P. Donate, Microwave-assisted

516

synthesis and antileishmanial activity of 3-methoxycarbonyl-γ-butyrolactone derivatives, J. Braz.

517

Chem. Soc. 25 (2014) 1331–1337. doi:10.5935/0103-5053.20140113.

fate

Curr.

tracking

studies,

and

nostoclide

Chem.

Org.

I,

Biol.

Biomol.

9

Chem.

Tetrahedron.

(2005)

13

285–292.

(2015)

199–206.

61

(2005)

5297–5302.

Lee,

G.

Phytochemistry.

Brown,

52

(1999)

Tribenzylbutyrolactones

537–540.

and

doi:10.1016/S0031-

dibenzyldiphenyl-4,5,6,7-

AC C

EP

S.

maculans,

TE D

Kyrtuthrix

M AN U

SC

A–C

Opin.

RI PT

compounds,

25

ACCEPTED MANUSCRIPT

518

[24] M. Pinatto-Botelho, P. Donate, A Rapid Protocol to synthesize gamma-butyrolactone

519

derivatives via the microwave technique, Curr. Microwave. Chem. 2 (2015) 83–87.

520

doi:10.2174/221333560201150212113120.

521

[25] D. Previdi, M. Rosa, P. Donate, Fast and efficient synthesis of maculalactone derivatives via

522

the

523

doi:10.2174/2213335604666170616122835.

524

[26] A. Bagno, G. Saielli, Addressing the stereochemistry of complex organic molecules by density

525

functional theory-NMR, Wiley Interdiscip. Rev. Comput. Mol. Sci. 5 (2015) 228–240.

526

doi:10.1002/wcms.1214.

527

[27] P. Willoughby, M. Jansma, T. Hoye, A guide to small-molecule structure assignment through

528

computation of (1H and

529

doi:10.1038/nprot.2014.042.

530

[28] D. Tantillo, Walking in the woods with quantum chemistry – applications of quantum chemical

531

calculations in natural products research, Nat. Prod. Rep. 30 (2013) 1079. doi:10.1039/c3np70028c.

532

[29] M. Lodewyk, M. Siebert, D. Tantillo, Computational prediction of 1H and 13C chemical shifts:

533

a useful tool for natural product, mechanistic and synthetic organic chemistry, Chem. Rev. 112

534

(2012) 1839–1862. doi:10.1021/cr200106v.

535

[30] G. Bifulco, P. Dambruoso, L. Gomez-Paloma, R. Riccio, Determination of relative

536

configuration in organic compounds by NMR spectroscopy and computational methods, Chem.

537

Rev. 107 (2007) 3744–3779. doi:10.1021/cr030733c.

538

[31] Y. Zang, G. Genta-Jouve, A. Escargueil, A. Larsen, L. Guedon, B. Nay, S. Prado,

539

Antimicrobial oligophenalenone dimers from the soil fungus Talaromyces stipitatus, J. Nat. Prod.

540

79 (2016) 2991–2996. doi:10.1021/acs.jnatprod.6b00458.

541

[32] T. Medeiros, H. Dias, E. Silva, M. Fukui, A. Soares, T. Kar, V. Heleno, P. Donate, R. Parreira,

542

A. Crotti, Detailed 1H and

543

neolignans, J. Braz. Chem. Soc. 27 (2015) 136–143. doi:10.5935/0103-5053.20150262.

544

[33] G. Resende, E. Alvarenga, P. Willoughby, Isolation and stereochemical assignment of

545

phthalides resulting from the Diels–Alder reaction between 5-isopropoxyfuran-2(5H)-one and

546

cyclopentadiene, J. Mol. Struct. 1101 (2015) 212–218. doi:10.1016/j.molstruc.2015.08.028.

547

[34] M. Lodewyk, C. Soldi, P. Jones, M. Olmstead, J. Rita, J. Shaw, D. Tantillo, The correct

548

structure of aquatolide – experimental validation of a theoretically-predicted structural revision, J.

549

Am. Chem. Soc. 134 (2012) 18550–18553. doi:10.1021/ja3089394.

Microwave.

Chem.

4

(2017)

229–237.

C) NMR chemical shifts, Nat. Protoc. 9 (2014) 643–660.

SC

13

Curr.

RI PT

technique,

EP

TE D

M AN U

microwave

C NMR spectral data assignment for two dihydrobenzofuran

AC C

13

26

ACCEPTED MANUSCRIPT

[35] G. Saielli, K. Nicolaou, A. Ortiz, H. Zhang, A. Bagno, Addressing the stereochemistry of

551

complex organic molecules by density functional theory-NMR: vannusal B in retrospective, J. Am.

552

Chem. Soc. 133 (2011) 6072–6077. doi:10.1021/ja201108a.

553

[36] R. Rotta, A. Cunha Neto, D de Lima, A. Beatriz, G. da Silva, Configuration of stilbene

554

derivatives by 1H NMR and theoretical calculation of chemical shifts, J. Mol. Struct. 975 (2010)

555

59–62. doi:10.1016/j.molstruc.2010.03.079.

556

[37] M. Constantino, L. da Silva Filho, A. Cunha Neto, V. Heleno, G. da Silva, J. Lopes, Structural

557

assignment of Diels–Alder adducts: an experimental and theoretical approach, Spectrochim. Acta

558

Part A Mol. Biomol. Spectrosc. 61 (2005) 171–176. doi:10.1016/j.saa.2004.04.002.

559

[38] G. da Silva, A.Cunha Neto, Calculated NMR as a tool for structural elucidation of jungianol

560

and mutisianthol, Tetrahedron. 61 (2005) 7763–7767. doi:10.1016/j.tet.2005.05.101.

561

[39] G. Barone, L. Gomez-Paloma, D. Duca, A. Silvestri, R. Riccio, G. Bifulco, Structure validation

562

of natural products by quantum-mechanical GIAO Calculations of

563

Chem.

564

CHEM3233>3.0.CO;2-0.

565

[40] G. Barone, D. Duca, A. Silvestri, L. Gomez-Paloma, R. Riccio, G. Bifulco, Determination of

566

the relative stereochemistry of flexible organic compounds by Ab initio methods: conformational

567

analysis and Boltzmann-averaged GIAO

568

doi:10.1002/1521-3765(20020715)8:14<3240::AID-CHEM3240>3.0.CO;2-G.

569

[41] N. Grimblat, A. Sarotti, Computational chemistry to the rescue: modern toolboxes for the

570

assignment of complex molecules by GIAO NMR calculations, Chem. Eur. J. 22 (2016) 12246–

571

12261. doi:10.1002/chem.201601150.

572

[42] S. Smith, J. Goodman, Assigning the stereochemistry of pairs of diastereoisomers using GIAO

573

NMR shift calculation, J. Org. Chem. 74 (2009) 4597–4607. doi:10.1021/jo900408d.

574

[43] S. Smith, J. Goodman, Assigning stereochemistry to single diastereoisomers by GIAO NMR

575

calculation:

576

doi:10.1021/ja105035r.

577

[44] N. Grimblat, M. Zanardi, A. Sarotti, Beyond DP4: an improved probability for the

578

stereochemical assignment of isomeric compounds using quantum chemical calculations of NMR

579

shifts, J. Org. Chem. 80 (2015) 12526–12534. doi:10.1021/acs.joc.5b02396.

580

[45] A. Sarotti, Successful combination of computationally inexpensive GIAO

581

calculations and artificial neural network pattern recognition: a new strategy for simple and rapid

582

detection

583

doi:10.1039/c3ob40843d.

SC

RI PT

550

8

(2002)

M AN U

J.

C NMR chemical shifts,

3233.

13

doi:10.1002/1521-3765(20020715)8:14<3233::AID-

C NMR chemical shifts, Chem. Eur. J. 8 (2002) 3240.

AC C

EP

TE D

Eur.

13

the

of

DP4

structural

probability,

J.

Am.

misassignments,

Chem.

Org.

Soc.

Biomol.

132

Chem.

(2010)

11

12946–12959.

13

C NMR

(2013)

4847.

27

ACCEPTED MANUSCRIPT

[46] M. Zanardi, A. Sarotti, GIAO C–H COSY simulations merged with artificial neural networks

585

pattern recognition analysis. Pushing the structural validation a step forward, J. Org. Chem. 80

586

(2015) 9371–9378. doi:10.1021/acs.joc.5b01663.

587

[47] P. Le-Huu, D. Petrović, B. Strodel, V. Urlacher, One-pot, two-step hydroxylation of the

588

macrocyclic diterpenoid β-cembrenediol catalyzed by P450 BM3 mutants, ChemCatChem. 8 (2016)

589

3755–3761. doi:10.1002/cctc.201600973.

590

[48] C. Kim, L. Subedi, J. Oh, S. Kim, S. Choi, K. Lee, Bioactive triterpenoids from the twigs of

591

Chaenomeles sinensis, J. Nat. Prod. 80 (2017) 1134–1140. doi:10.1021/acs.jnatprod.7b00111.

592

[49] S. Qiu, E. Gussem, K. Tehrani, S. Sergeyev, P. Bultinck, W. Herrebout, Stereochemistry of the

593

tadalafil diastereoisomers: a critical assessment of vibrational circular dichroism, electronic circular

594

dichroism,

595

doi:10.1021/jm401407w.

596

[50] J. Zhang, C. Taylor, E. Bowman, L. Savage-Low, M. Lodewyk, L. Hanne, G. Wu, Inverse

597

electron demand hetero-Diels–Alder reaction in preparing 1,4-benzodioxin from o-quinone and

598

enamine, Tetrahedron Lett. 54 (2013) 6298–6302. doi:10.1016/j.tetlet.2013.09.013.

599

[51] S. Smith, J. Channon, I. Paterson, J. Goodman, The stereochemical assignment of acyclic

600

polyols: a computational study of the NMR data of a library of stereopentad sequences from

601

polyketide natural products, Tetrahedron. 66 (2010) 6437–6444. doi:10.1016/j.tet.2010.06.022.

602

[52] A. Sarotti, A. Suárez, R. Spanevello, DFT calculations induced a regiochemical outcome

603

revision of the Diels–Alder reaction between levoglucosenone and isoprene, Tetrahedron Lett. 52

604

(2011) 3116–3119. doi:10.1016/j.tetlet.2011.04.021.

605

[53] D. Hodgson, C. Villalonga-Barber, J. Goodman, S. Pellegrinet, Synthetic and computational

606

studies on the tricarboxylate core of 6,7-dideoxysqualestatin H5 involving a carbonyl ylide

607

cycloaddition–rearrangement, Org. Biomol. Chem. 8 (2010) 3975. doi:10.1039/c004496b.

608

[54] S. Mohamed, M. Trabelsi, B. Champagne, Assigning the stereochemistry of syn and anti

609

β-trimethylsiloxy-α-trimethylsilyl alkanoic acid silyl esters using GIAO 1H NMR chemical shift

610

calculations, J. Mol. Struct. 1141 (2017) 436–440. doi:10.1016/j.molstruc.2017.03.112.

611

[55] M. Stucchi, G. Lesma, F. Meneghetti, G. Rainoldi, A. Sacchetti, A. Silvani, Organocatalytic

612

asymmetric Biginelli-like reaction involving isatin, J. Org. Chem. 81 (2016) 1877–1884.

613

doi:10.1021/acs.joc.5b02680.

614

[56] C. Talotta, C. Gaeta, M. Rosa, J. Ascenso, P. Marcos, P. Neri, Alkylammonium guest induced-

615

fit recognition by a flexible dihomo-oxacalix[4]arene derivative, European J. Org. Chem. 2016

616

(2016) 158–167. doi:10.1002/ejoc.201501319.

rotatory

dispersion,

J.

Med.

Chem.

56

(2013)

8903–8914.

SC

optical

AC C

EP

TE D

M AN U

and

RI PT

584

28

ACCEPTED MANUSCRIPT

[57] F. Cen-Pacheco, A. Santiago-Benítez, C. García, S. Álvarez-Méndez, A. Martín-Rodríguez, M.

618

Norte, V. Martín, J. Gavín, J. Fernández, A. Daranas, Oxasqualenoids from Laurencia viridis:

619

combined spectroscopic–computational analysis and antifouling potential, J. Nat. Prod. 78 (2015)

620

712–721. doi:10.1021/np5008922.

621

[58] I. Hwang, J. Oh, W. Zhou, S. Park, J.- Kim, A. Chittiboyina, D. Ferreira, G. Song, S. Oh, M.

622

Na, M. Hamann, Cytotoxic activity of rearranged drimane meroterpenoids against colon cancer

623

cells via down-regulation of β-catenin expression, J. Nat. Prod. 78 (2015) 453–461.

624

doi:10.1021/np500843m.

625

[59] E. Gussem, W. Herrebout, S. Specklin, C. Meyer, J. Cossy, P. Bultinck, Strength by joining

626

methods: combining synthesis with NMR, IR, and vibrational circular dichroism spectroscopy for

627

the determination of the relative configuration in hemicalide, Chem. Eur. J. 20 (2014) 17385–

628

17394. doi:10.1002/chem.201404822.

629

[60] D. Llompart, A. Sarotti, V. Corne, A. Suárez, R. Spanevello, G. Echeverría, O. Piro, E.

630

Castellano, Asymmetric construction of substituted pyrrolidines via 1,3-dipolar cycloaddition of

631

azomethine ylides and chiral acrylates derived from biomass, Tetrahedron Lett. 55 (2014) 2394–

632

2397. doi:10.1016/j.tetlet.2014.02.113.

633

[61] T. Tran, N. Pham, R. Quinn, Structure determination of pentacyclic pyridoacridine alkaloids

634

from the australian marine organisms Ancorina geodides and Cnemidocarpa stolonifera, European

635

J. Org. Chem. 2014 (2014) 4805–4816. doi:10.1002/ejoc.201402372.

636

[62] S. Smith, J. Goodman, CP3, (2009). http://www-jmg.ch.cam.ac.uk/tools/nmr/ (accessed June

637

13, 2017).

638

[63] W. Armarego, D. Perrin, Purification of laboratory chemicals, 5th ed., Elsevier, Oxford, 2003.

639

doi:10.1016/B978-0-7506-7571-0.X5000-5.

640

[64] V. Nardini, L. Dias, V. Palaretti, G. da Silva, Citronellal assumes a folded conformation in

641

solution due to dispersion interactions: a joint NMR-DFT analysis, J. Mol. Struct. 1157 (2018) 401–

642

407. doi:10.1016/j.molstruc.2017.12.083.

643

[65] S. Silva, S. Rodrigues, V. Nardini, A. de L. Vaz, V. Palaretti, G. da Silva, R. Vessecchi, G.

644

Clososki, Conformational dynamics of 4-formylaminoantipyrine based on NMR and theoretical

645

calculations, J. Mol. Struct. 1163 (2018) 280–286. doi:10.1016/j.molstruc.2018.03.003.

646

[66] Serena Software, PCMODEL, (2002) IN 47402-3076.

647

[67] F. Neese, The ORCA program system, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2 (2012) 73–

648

78. doi:10.1002/wcms.81.

AC C

EP

TE D

M AN U

SC

RI PT

617

29

ACCEPTED MANUSCRIPT

[68] F. Weigend, R. Ahlrichs, Balanced basis sets of split valence, triple zeta valence and quadruple

650

zeta valence quality for H to Rn: design and assessment of accuracy, Phys. Chem. Chem. Phys. 7

651

(2005) 3297. doi:10.1039/b508541a.

652

[69] A. Schäfer, C. Huber, R. Ahlrichs, Fully optimized contracted Gaussian basis sets of triple zeta

653

valence quality for atoms Li to Kr, J. Chem. Phys. 100 (1994) 5829–5835. doi:10.1063/1.467146.

654

[70] P. Stephens, F. Devlin, C. Chabalowski, M. Frisch, Ab initio calculation of vibrational

655

absorption and circular dichroism spectra using density functional force fields, J. Phys. Chem. 98

656

(1994) 11623–11627. doi:10.1021/j100096a001.

657

[71] A. Becke, Density‐functional thermochemistry. III. The role of exact exchange, J. Chem.

658

Phys. 98 (1993) 5648–5652. doi:10.1063/1.464913.

659

[72] C. Lee, W. Yang, R. Parr, Development of the Colle-Salvetti correlation-energy formula into a

660

functional

661

doi:10.1103/PhysRevB.37.785.

662

[73] S. Vosko, L. Wilk, M. Nusair, Accurate spin-dependent electron liquid correlation energies for

663

local spin density calculations: a critical analysis, Can. J. Phys. 58 (1980) 1200–1211.

664

doi:10.1139/p80-159.

665

[74] F. Neese, F. Wennmohs, A. Hansen, Efficient and accurate local approximations to coupled-

666

electron pair approaches: an attempt to revive the pair natural orbital method, J. Chem. Phys. 130

667

(2009) 114108. doi:10.1063/1.3086717.

668

[75] S. Grimme, M. Steinmetz, Effects of London dispersion correction in density functional theory

669

on the structures of organic molecules in the gas phase, Phys. Chem. Chem. Phys. 15 (2013) 16031.

670

doi:10.1039/c3cp52293h.

671

[76] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate Ab initio

672

parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J.

673

Chem. Phys. 132 (2010) 154104. doi:10.1063/1.3382344.

674

[77] Gaussian 03, Revision C.02, M. Frisch, G. Trucks, H. Schlegel, G. Scuseria, M. Robb, J.

675

Cheeseman, J. Montgomery, Jr., T. Vreven, K. Kudin, J. Burant, J. Millam, S. Iyengar, J. Tomasi,

676

V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. Petersson, H. Nakatsuji, M. Hada, M.

677

Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai,

678

M. Klene, X. Li, J. Knox, H. Hratchian, J. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts,

679

R. Stratmann, O. Yazyev, A. Austin, R. Cammi, C. Pomelli, J. Ochterski, P. Ayala, K. Morokuma,

680

G. Voth, P. Salvador, J. Dannenberg, V. Zakrzewski, S. Dapprich, A. Daniels, M. Strain, O. Farkas,

681

D. Malick, A. Rabuck, K. Raghavachari, J. Foresman, J. Ortiz, Q. Cui, A. Baboul, S. Clifford, J.

682

Cioslowski, B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Martin, D. Fox, T.

Phys.

SC

density,

Rev.

B.

37

(1988)

785–789.

M AN U

electron

TE D

the

AC C

EP

of

RI PT

649

30

ACCEPTED MANUSCRIPT

Keith, M. Al-Laham, C. Peng, A. Nanayakkara, M. Challacombe, P. Gill, B. Johnson, W. Chen, M.

684

Wong, C. Gonzalez, J. Pople, Gaussian, Inc., Wallingford CT, 2004.

685

[78] S. Grimme, S. Ehrlich, L. Goerigk, Effect of the damping function in dispersion corrected

686

density functional theory, J. Comput. Chem. 7 (2011) 1456–1465. doi:10.1002/jcc.

687

[79] H. Zimmerman, M. Traxler, The stereochemistry of the Ivanov and Reformatsky reactions. I, J.

688

Am. Chem. Soc. 79 (1957) 1920–1923. doi:10.1021/ja01565a041.

689

[80] C. Heathcock, C. Buse, W. Kleschick, M. Pirrung, J. Sohn, J. Lampe, Acyclic stereoselection.

690

7. Stereoselective synthesis of 2-alkyl-3-hydroxy carbonyl compounds by aldol condensation, J.

691

Org. Chem. 45 (1980) 1066–1081. doi:10.1021/jo01294a030.

692

[81] C. Heathcock, J. Lampe, Acyclic stereoselection. 17. Simple diastereoselection in the addition

693

of medium- and long-chain n-alkyl ketone lithium enolates to aldehydes, J. Org. Chem. 48 (1983)

694

4330–4337. doi:10.1021/jo00171a035.

695

[82] J. Mulzer, M. Zippel, A study on kinetic C-H-acidity; The α-epimerization of β-lactones,

696

Tetrahedron Lett. 22 (1981) 2165–2168. doi:10.1016/S0040-4039(01)90488-5.

697

[83] Y. Aso, Y. Hayashi, S. Yoshikoda, Y. Takeda, Y. Kita, Y. Nishimura, Y. Arata, Epimerization

698

and hydrolysis of etoposide analogues in aqueous solution, Chem. Pharm. Bull. (Tokyo). 37 (1989)

699

422–424. doi:10.1248/cpb.37.422.

AC C

EP

TE D

M AN U

SC

RI PT

683

31

ACCEPTED MANUSCRIPT Stereochemical assignment of four diastereoisomers of a maculalactone derivative by computational NMR calculations

Authors: Daniel Previdi*, Viviani Nardini, Mayla Eduarda Rosa, Vinicius Palaretti, Gil

RI PT

Valdo José da Silva and Paulo Marcos Donate

Departamento de Química, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto, Universidade de São Paulo, Avenida Bandeirantes, 3900, CEP 14040-901, Ribeirão Preto,

SC

SP, Brazil.

M AN U

*Corresponding author. E-mail address: [email protected] (D. Previdi).

Highlights

• NMR chemical shifts calculation and the relative configurations of lactones

TE D

• Stereochemical assignment of diastereoisomers of a tri-benzylated γ-butyrolactone • Determination of the relative configurations of a maculalactone derivative • CP3 parameter applied to distinguish four diastereoisomers

AC C

EP

• Assignment of the relative configurations corroborated a proposed reaction mechanism