Next Article in Journal
Additives Enhancing the Catalytic Properties of Lipase from Burkholderia cepacia Immobilized on Mixed-Function-Grafted Mesoporous Silica Gel
Previous Article in Journal
Isolation, Structural Analyses and Biological Activity Assays against Chronic Lymphocytic Leukemia of Two Novel Cytochalasins — Sclerotionigrin A and B
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Hydractinia echinata Test-System. III: Structure-Toxicity Relationship Study of Some Azo-, Azo-Anilide, and Diazonium Salt Derivatives

by
Sergiu Adrian Chicu
1,
Melania Munteanu
2,*,
Ioana Cîtu
3,*,
Codruta Şoica
4,
Cristina Dehelean
4,
Cristina Trandafirescu
4,
Simona Funar-Timofei
1,†,
Daniela Ionescu
4,† and
Georgeta Maria Simu
4,†
1
Institute of Chemistry Timisoara of the Romanian Academy, B-dul Mihai Viteazul 24, RO-300223 Timişoara, Romania
2
Department of Clinical Laboratory and Sanitary Chemistry, "Vasile Goldis" University, 1 Feleacului Str., Arad 310396, Romania
3
Faculty of Medicine, "V. Babes" University of Medicine and Pharmacy, 2 Eftimie Murgu, 300041 Timisoara, Romania
4
Faculty of Pharmacy, "V. Babes" University of Medicine and Pharmacy, 2 Eftimie Murgu, 300041 Timisoara, Romania
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2014, 19(7), 9798-9817; https://doi.org/10.3390/molecules19079798
Submission received: 17 April 2014 / Revised: 29 June 2014 / Accepted: 3 July 2014 / Published: 8 July 2014

Abstract

:
Structure-toxicity relationships for a series of 75 azo and azo-anilide dyes and five diazonium salts were developed using Hydractinia echinata (H. echinata) as model species. In addition, based on these relationships, predictions for 58 other azo-dyes were made. The experimental results showed that the measured effectiveness Mlog(1/MRC50) does not depend on the number of azo groups or the ones corresponding to metobolites, but it is influenced by the number of anilide groups, as well as by the substituents’ positions within molecules. The conformational analysis pointed out the intramolecular hydrogen bonds, especially the simple tautomerization of quinoidic (STOH) or aminoidic (STNH2) type. The effectiveness is strongly influenced by the “push-pull” electronic effect, specific to two hydroxy or amino groups separated by an azo moiety (double alternate tautomery, (DAT), to the –COOH or –SO3H groups which are located in ortho or para position with respect to the azo group. The levels of the lipophylic/hydrophilic, electronic and steric equilibriums, pointed out by the Mlog(1/MRC50) values, enabled the calculation of their average values Clog(1/MRC50) (“Köln model”), characteristic to one derivative class (class isotoxicity). The azo group reduction and the hydrolysis of the amido/peptidic group are two concurrent enzymatic reactions, which occur with different reaction rates and mechanisms. The products of the partial biodegradation are aromatic amines. No additive or synergic effects are noticed among them.

1. Introduction

The importance of azo derivatives is due to broad range of applications, which practically cover all human activities, both as dyes [1] and pigments [2]. In the medical-pharmaceutics field, the use of azo-derivatives is of interest to assess the permeability of the blood-brain barrier (BBB) to macromolecules [3]. Several affections, such as Prion-, Alzheimer’s-, Chorea major- or Parkinson’s diseases, are due to protein misfolding and aggregation (amiloide) processes [4]. The azoic dye Congo Red detects and interfere with their formation process, stabilizes native protein monomers or partially folded intermediates, but generally little is known about the exact mechanism of the inhibitor’s action on protein aggregation [5].
From a structural point of view, an azo derivative molecule contains an azo group –N=N– (chromophore), which binds together two substituted aromatic nuclei. The coplanarity of the dye molecule allows an extended conjugation of the π-electrons of the azo group, –N=N– and of the aromatic nucleus with the p electrons of some +M effect auxochromes (–OH, –OR, –NH2, –NR2) or of some –M effect antiauxochromes (–NO2, –COOH, –SO3H) [6].
The main metabolic process to which the azo dyes are subjected in the human organism is the reduction of the azo group. In the first stage, a hydrazo derivative appears, by the action of the electron-withdrawing substituents, such as the sulfonic groups, even when these groups are more separated from the azo group [1]. The second step, e.g., the decomposition of the hydrazo derivative to its corresponding free amines, is dependent on the charge difference among nitrogen atoms, as a direct consequence of some electron-donating substituents [7].
Most azo derivatives contain, beside the azo group, one or more amidic (anilidic) groups (chromophores). The hydrolysis of these compounds occurs enzymatically [8], like the naphthols AS’s hydrolysis [9], involving the formation of an oxyanionic intermediate during the rate-determining step.
The reduction of the azo group and the hydrolysis of an amido group are two different enzymatic reactions, which probably occur concurrently. The biodegradation rate will be given by the slowest step of the reaction and is affected when the coplanarity of the molecule is disturbed by steric hindrances [10].
The reaction products are aromatic amines [11], some of them exhibiting mutagenic effects even via skin microflora, according to in vitro experiments with Staphylococus aureus [12]. Although the mutagenic or the carcinogenic character may be significantly decreased by sulphonation [13], the presence of the strongly polar sulfonic groups, confers to the molecules a marked hydrophilicity and an increased solubility, reduces very much the permeability of the lipophilic cellular membrane [14] and consequently their biodegradability is decreased to a great extent [15].
The first aim of the present work was the comparative determination of the toxicity of some azo, azo- anilide, and diazonium salts using the biological H. echinata test-system (HeTS). The larvae of H. echinata have an elongated spindle shaped body of about 1 mm in length and a diameter of 100 µm. They consist of about 10,000 cells. They have no mouth, no gut, no extremities, and no sense organs other than nerve cells which may serve to sense environmental signals. As a consequence, one could estimate a xenobiotic’s direct action at the cellular level. This matter of fact could be also the reason why the test system proved a higher sensibility for the identification of the toxicity of some nonyl- phenolic isomers in comparison to the Daphnia-immobilisation test, as well as to the duckweed-growth inhibition test [16].
H. echinata represent more than substitutes for tests on superior organisms. Further, the influence of structural modifications on effectiveness, the presentation of some hydrolysis/reduction reaction mechanisms of azo-anilide derivatives, and some prediction possibilities for a series of non-tested yet derivatives, were taken into account. Conformational analysis was performed by molecular mechanics and quantum chemical calculations in order to obtain the title compound intramolecular hydrogen bonds, especially the simple tautomerization of quinoidic (STOH) or aminoidic (STNH2) type.

2. Results and Discussion

In Table 1 the logarithm of the reciprocal value (log1/MRC50) values are presented. The average (C) concentrations of these values were calculated too. The MRC50 value calculations (Chicu et al. [17]) result from the graphical representation of the metamorphosis variation, M (%), (Y axis) function of the xenobiotic’s concentration (mol/L) (X axis), where the metamorphosis decreases with the rise of the xenobiotic’s concentration. Thus, the MRC50 value represents the xenobiotic’s concentration (mol/L) necessary for a 50% decrease of metamorphosis, with respect to control.
The Clog(MRC50) value was obtained as the sum of the least squares differences between the Mlog(MRC50) measured values and Clog(MRC50) (Microsoft Office Excel 2003) based on the experimental data/algorithm according to which +/− 0,50 log.u differences characterize the xenobiotics with identical toxicity (isotoxicity). The accuracy of experimental (log1/MRC50) data was checked by the standard deviations, which were determined by the Excel 2003 program. For each concentration of substance experiments were performed in triplicate and were repeated at least twice.
Table 1. Hydractinia echinata test system: experimental Mlog/(1/MRC50) (M) and calculated average Clog(MRC50) (C) toxicities of some azo-, azo-anilide, diazonium salts derivatives.
Table 1. Hydractinia echinata test system: experimental Mlog/(1/MRC50) (M) and calculated average Clog(MRC50) (C) toxicities of some azo-, azo-anilide, diazonium salts derivatives.
No.Compound NameMCNo.Compound NameMC
1 a(4-Methoxyphenyl) (phenyl)diazene5.204.3670 cAcid Red 151 3.68
2 b4-(2-Phenyldiazen-1-yl)benzoic acid3.673.6871 cSudan Red 7B (Solvent Red 19) 3.68
3 bMethyl Red3.683.6872 cAcid Red 73 (Crocein Scarlet 3B) 3.68
4 bMethyl Red (sodium salt)3.793.6873 cChlorantine Fast Red 5B 3.68
5 a2-Amino-5-[(E)-2-(2-carboxyphenyl)diazen-1-yl]benzoic acid4.584.3674 aAcid Black 1 (Amido Black 10B)4.334.38
6 a4-{2-[4-(Diethylamino)phenyl]diazen-1-yl}benzoic acid4.774.3675 aAcid Black 1 (Salt)4.414.38
7 b4-[2-(4-Hydroxyphenyl)diazen-1-yl]benzoic acid3.323.6876 cNaphthalene Blue Black CS 4.38
8 b4-[(E)-2-(2,4-dihydroxy-phenyl)diazen-1-yl]benzoic acid3.313.6877 bDirect Black 323.893.68
9 b5-[2-(4-Carboxyphenyl)diazen-1-yl]-2-hydroxybenzoic acid3.353.6878 cDirect Black 22 (Pontamine Fast Black PGR) 3.68
10 c2-{2-[4-(Phenylamino)-phenyl]diazen-1-yl}benzoic acid 4.3679 cAcid Blue 113 4.36
11 c4-{2-[4-(Phenylamino)-phenyl]diazen-1-yl}benzoic acid 4.3680 aSudan Black B (Solvent Black 3)4.204.36
12 aMetanil Yellow (sodium salt)4.174.3681 bBordo Direct AN3.863.68
13 c4-{2-[4-(Phenylamino)-phenyl]diazen-1-yl}benzene-1-sulfonic acid 4.3682 aCongo Red4.494.36
14 bMethyl Orange (Orange III)3.743.6883 aDirect Black 384.264.36
15 c4-{2-[4-(Diethylamino)-phenyl]diazen-1-yl}benzene-1-sulfonic acid 4.3684 cDirect Green 6 4.36
16 c4-[2-(4-Hydroxyphenyl)diazen-1-yl]benzene-1-sulfonic acid 3.6885 cDirect Green 1 3.68
17 cTropaeolin O 3.6886 cCongo Corinth G 4.36
18 c2-Hydroxy-5-[2-(4-sulfophenyl)diazen-1-yl]benzoic acid 3.6887 aAcid Red 974.114.36
19 cFast Yellow AB 3.6888 cAcid Yellow 42 4.36
20 a5-Methyl-2-[2-(4-nitrophenyl)-diazen-1-yl]phenol4.114.3689 aDirect Blue 53 (Evans Blue)5.134.36
21 a3-[(E)-2-(2,4-dihydroxyphenyl)-diazen-1-yl]-4-hydroxybenzene-1-sulfonic acid4.274.3690 aTrypanblue (Direct Blue 14)4.094.36
22 cSudan Orange G 3.6891 bDirect Blue 15 (Aizen Direct Sky Blue 5B)3.693.68
23 b4-[2-(2-Hydroxynaphthalen-1-yl)diazen-1-yl]benzoic acid3.613.6892 bDirect Blue 13.793.68
24 c4-[2-(1-Hydroxynaphthalen-2-yl)diazen-1-yl]benzoic acid 3.6893 cDirect Blue 151 3.68
25 b4-[2-(4-Carboxyphenyl)diazen-1-yl]-3-hydroxynaphthalene-2-carboxylic acid3.453.6894 aFast Blue B4.304.36
26 a4-[2-(1-Aminonaphthalen-2-yl)diazen-1-yl]benzoic acid4.734.3695 aTetrazolium Blue Chloride4.044.36
27 c4-[2-(2-Aminonaphthalen-1-yl)diazen-1-yl]benzoic acid 4.3696 bBrilliant Yellow3.843.68
28 bAcid Orange 73.173.6897 b5-[(E)-2-(2,4-Dihydroxyphenyl)-diazen-1-yl]-2-[(E)-2-{4-[(E)-2-(2,4-dihydroxyphenyl)-diazen-1-yl]-2-sulfophenyl}ethenyl]-benzene-1-sulfonic acid3.733.68
29 c4-[2-(1-Hydroxynaphthalen-2-yl)diazen-1-yl]benzene-1-sulfonic acid 3.6898 b2-{2-[4-(2-{4-[2-(3-Carboxy-4-hydroxy-phenyl)diazen-1-yl]-2-sulfophenyl}ethenyl)-3-sulfophenyl]diazen-1-yl}-5-hydroxybenzoic acid3.393.68
30 c3-Hydroxy-4-[2-(4-sulfophenyl)-diazen-1-yl]naphthalene-2-carboxylic acid 3.6899 cDirect Yellow 12 3.68
31 cDirect Yellow 11 3.68100 a5-[2-(2-Hydroxy-naphthalen-1-yl)diazen-1-yl]-2-(2-{4-[2-(2-hydroxynaphthalen-1-yl)diazen-1-yl]-2-sulfophenyl}ethenyl)-benzene-1-sulfonic acid4.554.36
32 cAcid Violet R 4.36101 b4-{2-[4-(2-{4-[2-(3-Carboxy-2-hydroxy-naphthalen-1-yl)diazen-1-yl]-2-sulfophenyl}-ethenyl)-3-sulfo-phenyl]diazen-1-yl}-3-hydroxynaphthalene-2-carboxylic acid3.873.68
33 cSudan I 3.68102 b4-Amino-3-{2-[4-(2-{4-[2-(1-amino-4-sulfonaphthalen-2-yl)diazen-1-yl]-2-sulfophenyl}ethenyl)-3-sulfophenyl]diazen-1-yl}naphthalene-1-sulfonic acid3.743.68
34 c4-(2-Phenyldiazen-1-yl)naphthalen-1-ol 3.68103 bFluorescent Brightener 283.813.68
35 cSolvent Orange 7 (Sudan II) 3.68104 a2-Hydroxy-4-[(E)-2-(4-{4-[(E)-2-(2-hydroxy-phenyl)diazen-1-yl]-benzamido}phenyl)-diazen-1-yl]benzoic acid4.524.36
36 cAcid Ponceau 2G 3.68105 b5-[(E)-2-(4-{4-[(E)-2-{2,4-dihydroxy-5-[(E)-2-(2-hydroxy-5-sulfo-phenyl)diazen-1-yl]-phenyl}diazen-1-yl]-benzamido}phenyl)-diazen-1-yl]-2-hydroxybenzoic acid3.643.68
37 cOil Yellow AB 3.68106 b5-[2-(4-{4-[2-(3-Carboxy-4-hydroxyphenyl)diazen-1-yl]benzamido}-phenyl)diazen-1-yl]-2-hydroxybenzoic acid3.443.68
38 aAcid Red 26 (Ponceau Xilidine)4.064.36107 b5-{2-[4-({4-[2-(6-Amino-1-hydroxy-3-sulfonaphthalen-2-yl)diazen-1-yl]-phenyl}carbamoyl)-phenyl]diazen-1-yl}-2-hydroxybenzoic acid3.663.68
39 bAcid Orange 10 (Orange G)3.343.68108 b5-{2-[4-({4-[2-(7-Amino-1-hydroxy-3-sulfonaphthalen-2-yl)diazen-1-yl]-phenyl}carbamoyl)-phenyl]diazen-1-yl}-2-hydroxybenzoic acid3.673.68
40 bThorin3.463.68109 b4-Amino-3-[2-(4-{4-[2-(1-amino-4-sulfonaphthalen-2-yl)-diazen-1-yl]-benzamido}phenyl)-diazen-1-yl]-naphthalene-1-sulfonic acid3.783.68
41 bAcid Red 29 (Chromotrope 2R)3.653.68110 b3-Hydroxy-4-[2-(4-{4-[2-(2-hydroxy-3,6-disulfonaphthalen-1-yl)diazen-1-yl]-benzamido}phenyl)-diazen-1-yl]-naphthalene-2,7-disulfonic acid3.713.68
42 bAcid Red 1 (Azophloxin)3.663.68111 b5-Amino-3-{2-[4-({4-[2-(7-amino-1-hydroxy-3-sulfonaphthalen-2-yl)diazen-1-yl]phenyl}-carbamoyl)phenyl]diazen-1-yl}-4-hydroxy-naphthalene-2,7-disulfonic acid3.583.68
43 bAcid Red 353.633.68112 a4-(3-Methyl-4-{2-[4-(4-{2-[5-methyl-3-oxo-2-(4-sulfophenyl)-2,3-dihydro-1H-pyrazol-4-yl]diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}-5-oxo-2,5-dihydro-1H-pyrazol-1-yl)benzene-1-sulfonic acid4,024.36
44 c5-Amino-4-hydroxy-3-[2-(4-nitrophenyl)diazen-1-yl]naphthalene-2,7-disulfonic acid 4.36113 b5-[2-(4-{[4-(2-{8-Amino-1-hydroxy-7-[2-(4-nitrophenyl)diazen-1-yl]-3,6-disulfonaphthalen-2-yl}diazen-1-yl)-phenyl]carbamoyl}-phenyl)diazen-1-yl]-2-hydroxybenzoic acid3.953.68
45 c5-Amino-4-hydroxy-3-[2-(2-hydroxy-5-nitrophenyl)diazen-1-yl]naphthalene-2,7-disulfonic acid 3.68114 b5-[2-(4-{[4-(2-{8-Amino-7-[2-(2-carboxy-4-nitrophenyl)diazen-1-yl]-1-hydroxy-3,6-disulfonaphthalen-2-yl}diazen-1-yl)phenyl]-carbamoyl}phenyl)-diazen-1-yl]-2-hydroxybenzoic acid3.893.68
46 c2-[2-(1-Amino-8-hydroxy-3,6-disulfonaphthalen-2-yl)diazen-1-yl]-5-nitrobenzoic acid 4.36115 b5-{2-[4-({4-[2-(6-Amino-1-hydroxy-5-{2-[2-(1-hydroxy-ethenyl)-4-nitro-phenyl]diazen-1-yl}-3-sulfonaphthalen-2-yl)diazen-1-yl]phenyl}-carbamoyl)phenyl]-diazen-1-yl}-2-hydroxybenzoic acid3.863.68
47 c2-[2-(2-Amino-5-hydroxy-7-sulfonaphthalen-1-yl)diazen-1-yl]-5-nitrobenzoic acid 4.36116 b4-Amino-3-{2-[4-({4-[2-(2,4-diamino-phenyl)diazen-1-yl]phenyl}carbamoyl)-phenyl]diazen-1-yl}-5-hydroxy-6-(2-phenyl-diazen-1-yl)-naphthalene-2,7-disulfonic acid3.783.68
48 cAcid Red 4 3.68117 b2-Hydroxy-5-{2-[4-(4-{2-[4-hydroxy-3-(phenylcarbamoyl)phenyl]diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}-N-phenylbenzamide3.783.68
49 cSolvent Red 1 3.68118 b7-Benzamido-4-hydroxy-3-{2-[4-(4-{2-[4-hydroxy-3-(phenyl-carbamoyl)phenyl]diazen-1-yl}benzamido)-phenyl]diazen-1-yl}naphthalene-2-sulfonic acid3.653.68
50 cPigment Orange 5 4.36119 a5-{2-[4-({4-[2-(6-Benzamido-1-hydroxy-3-sulfonaphthalen-2-yl)diazen-1-yl]-phenyl}carbamoyl)phenyl]diazen-1-yl}-2-hydroxybenzoic acid4.654.36
51 cAcid Violet 7(Pontacyl Carmine 6B) 4.36120 a4-Amino-3-{2-[4-(4-{2-[4-hydroxy-3-(phenyl-carbamoyl)phenyl]diazen-1-yl}benzamido)phenyl]-diazen-1-yl}naphthalene-1-sulfonic acid4.604.36
52 cAcid Brown 4 3.68121 a6-Amino-4-hydroxy-3-{2-[4-(4-{2-[4-hydroxy-3-(phenyl-carbamoyl)phenyl]diazen-1-yl}benzamido)-phenyl]diazen-1-yl}naphthalene-2-sulfonic acid4.504.36
53 cSolvent Orange 2 3.68122 a5-Amino-4-hydroxy-3-{2-[4-(4-{2-[4-hydroxy-3-(phenyl-carbamoyl)phenyl]-diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}-naphthalene-2,7-disulfonic acid4.504.36
54 cSolvent Red 3 3.68123 c6-Hydroxy-5-{2-[4-(4-{2-[4-hydroxy-3-(phenylcarbamoyl)phenyl]diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}-naphthalene-2-sulfonic acid 4.36
55Acid Yellow 23 (Tartrazine, Food Yellow 4)NOTE 124 c7-Amino-4-hydroxy-3-{2-[4-(4-{2-[4-hydroxy-3-(phenylcarbamoyl)-phenyl]diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}naphthalene-2-sulfonic acid 4.36
56 bAcid Red 88 (Fast Red AV)3.313.68125 c4-Amino-5-hydroxy-6-{2-[4-(4-{2-[4-hydroxy-3-(phenylcarbamoyl)-phenyl]diazen-1-yl}benzamido)phenyl]-diazen-1-yl}naphthalene-1,3-disulfonic acid 4.36
57 bAcid Red 27 (Amaranth, Food Red 9)2.333.68126 c4,5-Dihydroxy-3-{2-[4-(4-{2-[4-hydroxy-3-(phenylcarbamoyl)-phenyl]diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}-naphthalene-2,7-disulfonic acid 4.36
58 aEriochrom Black T4.134.36127 a3-Hydroxy-4-{2-[4-(4-{2-[4-hydroxy-3-(phenylcarbamoyl)phenyl]diazen-1-yl}-benzamido)phenyl]-diazen-1-yl}-naphthalene-2,7-disulfonic acid4.234.36
59 cCrystal Ponceau 6R 3.68128 a2-Hydroxy-5-{2-[4-({4-[2-(2-hydroxynaphthalen-1-yl)diazen-1-yl]-phenyl}carbamoyl)-phenyl]diazen-1-yl}-N-phenylbenzamide4.274.36
60 bBismark Brown Y (Vesuvin)3.873.68129 bDirect Brilliant Orange3.653.68
61 cBasic Brown 41 3.68130 cDirect Red 23 (sodium salt) 3.68
62 cBasic Brown 4 (Bismark Brown R) 3.68131 cDirect Orange 26 3.68
63 bSudan IV (Scarlet Red)3.053.68132 bDirect Red 80 (Saturn Rot F3B)3.803.68
64 bSolvent Red 25 (Sudan Red B)3.423.68133 cDirect Red 75 3.66
65 bSolvent Red 23 (Sudan III, Sudan V)3.643.68134 cDirect Yellow 50 4.36
66 cSolvent Red 26 3.68135 aFast Blue RR Salt4.094.36
67 cSolvent Red 27 (Oil Red O) 3.68136 aFast Blue BB Salt4.174.36
68 bAcid Red 66 (Biebrischer Scharlach)3.953.68137 aFast Red B Salt5.254.36
69 bPonceau S3.973.68138 cFast Black K Salt 4.36
Compound structures are presented in the Supplementary Material; a – “up”; b – “down”; c – “prediction”; NOTE – no observable toxic effect.
Considering that toxicity is constant in a range of ±0.50 log units (Köln model), the measured toxicity values Mlog(1/MRC50) are compared to the calculated average value Clog(1/MRC50). The introduction of this parameter points out the “class isotoxicity” character of the tested derivatives, which is also noticed in the case of AS-naphthols [9].
Depending on the experimental values, one could consider two calculated average values, and consequently an isotoxicity situated on two levels: Clog(1/MRC50) = 3.68 for “down” substrate (S) - receptor (R) interactions and 4.36 for “up” interactions (Table 1).
Just as in case of the Naphthol-AS [9], the measured toxicity values, F(1+2+…)e could be higher than the individual values of the reaction products F(1), F(2) etc., especially in the case of “strong” interactions. However, in this situation, one could not notice the additivity F(1+2+…)e > F(1+2+...)t (e = experimental; t = theoretical), as suggested by Backhaus [18] (Table 2). No synergistic effects were noticed.
For all the tested compounds, SUAD-subadditive toxicity values [9], e.g., (F(1 + 2+..)e < F(1 + 2+..)t, were found equal to zero. It is relevant to highlight the fact that none of the studied cases presented synergism, considering the individual toxicities of the metabolites. The lack of synergism is probably due to the metabolites’ reduced concentrations as compared to the initial xenobiotic, even if the individual toxicity was very high. For instance: 2,4-DHA: 2,4-dihydroxy aniline with M = 6.26.
The results obtained so far indicate that the dye (xenobiotic of first generation) is able to establish the reaction mechanism and the major antagonic effect through its chromophore. This effect is not influenced by the number, concentration, or the individual effectivenesses of the reaction products as simple functions (metabolites or second generation of xenobiotics).
Table 2. Hydractinia echinata test system: experimental (F(1+2+…)e), theoretical (F(1+2+...)t) and individual (F(1), F(2),…) toxicities of some azo- and azo-anilide derivatives and their products of enzymatic reduction and hydrolysis reaction.
Table 2. Hydractinia echinata test system: experimental (F(1+2+…)e), theoretical (F(1+2+...)t) and individual (F(1), F(2),…) toxicities of some azo- and azo-anilide derivatives and their products of enzymatic reduction and hydrolysis reaction.
No.F(1+2+...)eF(1)F(1)F(2)F(2)F(3)F(3)F(4)F(4)F(1+2+...)t
15.20A2.93 *p-Methoxy-aniline3.32----6.21
23.67A2.93 *APAB3.34----6.23
73.32APAB3.34p-AF5.78 *----9.12
83.31APAB3.342,4-DHA6.26----9.60
93.35APAB3.34Ac.p-AS2.97----6.31
143.74APABS2.81p-DMAA3.50 *----6.31
204.11p-NA2.852-Aminocresol3.21----6.06
963.84ADASDS3.43p-AF( 2M)5.78 *----14.99
973.73ADASDS3.432,4-DHA (2M)6.26----15.95
1044.52Ac p-AmS2.97APAB3.342,4-DHA6.26p-FDA2.93 *15.50
1063.44Ac p-AmS (2M)2.97APAB3.34 p-FDA2.93 *9.24
1173.78Ac p-AmS (2M)2.97A (2M)2.93 *APAB3.34p-FDA2.93 *18.07
F(1), F(2),.. etc.: individual log(1/MRC50) toxicity values; SUAD-subadditive toxicity values (for which F(1 + 2+..)e < F(1 + 2+..)t); e = experimental; t = theoretical; A: aniline; APAB: p-aminobenzoic acid; p-NA: p-nitroaniline; ADASDS: diaminostilbene disulfonic acid; 4,4'-DABA: 4,4'-diaminobenzanilide; APABS: p-aminobenzenesulfonic acid; Ac.p-AmS: p-aminosalicylic acid; p-DMAA: p-dimethylaminoaniline; p-AF: p-aminophenol; 2,4-DHA: 2,4-dihydroxyaniline; p-FDA : p-phenylendiamine; molar concentration (2M) - in parenthesis; * Calculated values.
This idea is supported by: (a) the effectiveness of compounds 74 and 75, which are identical even if one contains the cyclohexyl-ammonium salt (simple function) as first generation xenobiotic. The ssme phenomenon was observed in case of derivatives 135 and 136; (b) the effectiveness of acetylene alcohols, which is determined by the presence of a reactive carbonyl group in the structure of intermediate metabolites and not by the hydroxy group included in the initial xenobiotic [19].
The energetically most stable structures are those in which an intramolecular hydrogen bond is formed between the hydroxyl, aminic hydrogen and nitrogen azo atoms, respectively, as derived from the restricted Hartree-Fock calculations. Usually the moiety including the azo group and the attached phenyl rings to this group are coplanar, except compounds 23, 25, 30, 31, 33, 35, 36, 38, 40, 41, 42, 43, 44, 45, 49, 50, 53, 56, 57, 59, 63, 64, 65, 66, 67, 68, 69, 70, 72, 101, 110, 123, 127, 128 in which the azo group is attached at position 1 to the naphthyl fragment, including a hydroxyl group in position 2 or 8, or a bulky (carboxylic, sulfonic acid, etc.) group is present in the ortho phenyl (naphthyl fragment) position with respect to the azo group e.g.,: 46, 47, 48, 51, 58, 71, 73, 74, 75, 77, 83, 84, 85, 107, 108, 111, 113, 114, 115, 116, 118, 119, 121, 122, 124, 126, 130, 131, 132, 133, 136, 137) where slight deviations from planarity are noted.
According to the tested derivatives’ structures, the following main reactions are possible: the azo group’s reductive cleavage, as well as the anilidic and ureic group’s hydrolysis. In biological systems which contain azo-reductases, the azo group can be easily reduced with formation of the corresponding amines [20].
In the case of the anilide group, its hydrolysis involves the formation of an oxyanionic intermediate during the rate determining step [9]. The electronic shift in the carbonyl oxygen’s direction, increases the C=O bond length from 1.2 Å (fundamental state) to 1.4 Å (transition state). The rehybridization of the carbonyl carbon from a planar sp2 structure to the tetrahedral sp3 structure, is associated with a translation movement of about 1.2 Å as compared to the double C=O bond, the formation of the oxyanionic structure, as well as with the intervention of the nucleophilic Ser195 agent [8]. The transition stage is stabilized by the presence of a water molecule [21], which is further used in the hydrolysis process.
For ureic groups’ hydrolysis, the molecule is separated into two parts, which can further react independently [10].
The effectiveness is not decisively influenced by the intramolecular H-bonds formation, but is dependent on the substituents’ and chromophores’ reciprocal positions, as well as by the molecule’s electronic delocalization possibilities. Thus, the presence of the hydroxyl group in the ortho position of the azo group, induces the simple quinoidic tautomery STOH, through formation of a hydrazo derivative intermediate [22]: the hydrogen bond links the OH group to the non-adjacent nitrogen atom of the aromatic nucleus on which it is located.
According to the interatomic distance determinations STOH (* means partially) appears in derivatives 41, 42, 43, 58, 77*, 91*, 107, 111, 115, 116*, 118, 119, 121, 122 and 132* respectively, seems “unfavourable” in the presence of SO3H–azo H-bonds (39, 68), respectively OH–SO3H (69, 127), SO3H–SO3H (97, 100), OH–COOH (25, 101*), and do not “appear” in case of compound 128 (OH-NH). One notices some other cases: the “hydrogen in equilibrium” among OH–azo–OH (58) or OH–azo–SO3H bonds (91, 92, 132*), the simultaneous involvement of OH–SO3H in the H-bonds at the same azo group (107), as well as the formation of non-tautomeric H-bonds for derivatives 89, 91, 92, 110* and 132* (hydrogen bonds are presented in Table S3, Supplementary Information).
Gregory [20] has pointed out the presence of the ortho azo-aminoidic STNH2, but due to its instability, the amino derivative remains as the azo form. In the present work, this matter of fact is specific for derivative 26, and probably for 83. Non-tautomeric H-bonds are noticed in the case of derivatives 60, 77*, 81, 82, 83*, 102, 109, 113, 114, 115*, 116 and 120, respectively.
The formation of adjacent cycles through multiple tautomeric-type H-bonds (in which both hydrogen atoms of the amino group are involved) of azo–OH–NH2–azo (74, 75, 83, 90, 111, 113, 114, 116) or azo–OH–NH2–SO3H (83, 89) is also possible.
The presence of two hydroxy or amino groups located on different aromatic nuclei separated by an azo group, or present on the same nucleus, but located between two azo groups (e.g., H acid), determines electronic shifts of the “push-pull” type and the double alternate tautomery (DAT). This fact is clearly specific in case of compounds 21 and 87. Due to the fact that the OH– group is more reactive than the NH2, one could also imagine “push-pull” type electronic shift in case of structures 74, 75 and 83*, respectively, and probably in the case of 113 and 116*.
The predicted effectiveness of 58 azo dyes were estimated taking into account the structural similarities of these compounds with the tested derivatives.

2.1. The “Down”Substrat-Receptor Interactions

One could notice that the absence of internal H-bonds in compounds 2–4, 7, 14 and, the presence of an internal H-bond between –OH and –COOH of derivatives 9, 25, 101, 105, 106, 117, the singular STOH of 8, 23, 28 and 108, the involvement of -SO3H group in STOH of 107 and 118 or in STNH2 of 116, the non tautomeric amino-azo bonds of 81, 102, 109, 113–116, the steric hindrances of sulfonic groups at the double stilbenic bond level for 96–98, respectively, could generally contribute to some slightly reduced Mlog(1/MRC50) values.
The test system points out the steric effects, even if related to methyl substituents, which are less bulky [1]. This fact is more obvious in the case of Sudan IV (63), which has two methyl groups in the ortho position as compared to Sudan Red B (64) (with one methyl group), and Sudan III (65) (which does not have any methyl group and thus has the greatest effectiveness). This “sensibility” of the test-system was also noticed from the toxicity values of some nonyl-phenolic derivatives [23], but was not observed in case of Daphnia magna and duckweed test systems [16].
The effectiveness of derivatives of Sudan type can be correlated even with logP: the low reactivity of Sudan IV is due to its high lipophilicity (Table S2, Supplementary Information).
The steric hindrances corresponding to the dianisidinic’ –OCH3 group are stronger than those corresponding to the benzidinic’ –CH3 group, and this is why derivatives 91 and 92, with multiple H-bonds, involving even the methoxy group, exhibit lower toxicities than 89 and 90.
The identical inductive effects, but of opposite +M sign, exerted by two amino groups which can form non-tautomeric H-bonds, are responsible for the “down” interaction of derivative 60. It seems that the replacement of one –OH group by –NHCOCH3 in 41, or in 42 and 43, does not lead to modifications of their effectiveness.
The Amaranth molecule (57) exhibits a special polarity [24], and thus its lipophilic/hydrophilic balance (Table S2, Supplementary Information) is preponderantly in favor of the hydrophilic property. Therefore, the effectiveness of compound 56 which does not possess any sulfonic group is higher by 1 log unit.
According to the experimental values, the presence of the –SO3H group located in the ortho position to an azo group increases the effectiveness values close to “up” of derivatives 68 and 69, both by a negative mesomeric (–M) effect, as well as by affecting the molecules’ coplanarity.
In the case of azo-anilidic derivatives the general reaction mechanism involves the strong electron attracting effect of –M type of the carbonyl group, and the formation of a hydrazo derivative. This reduction is probably preceded by the anilide hydrolysis, since the carbonyl group is the permanently active reaction partner for Ser-195, and hence the reason of the electronic lack of balance. Moreover, one could emphasize that the anilide hydrolysis is faster than the azo group’s reduction, due to the fact that under identical experimental conditions, the total average effectiveness of naphthols-AS compounds is about 4.54 [9] as compared to the azo derivative’s value, which is equal to 4.01.
Because an anilide group can activate a single azo group, the anilide hydrolysis and the reduction of the intermediate hydrazo derivative are processes which occur fast. The reaction rate-determining step is represented by the reduction of the second and the third azo group. This is the case of derivatives 105–111 and 113–116, even if 113–115, also possess NO2 substituents. In the case of derivatives 117 and 118, except for the internal H-bonds, the three anilide groups exhibit antagonistic electronic effects, the azo group’s activation takes place successively, and the reaction rate is also lower. Thorin I (40), can be included in the present series of Mlog(1/MRC50) values, and this fact leads to the conclusion that the arsenic and the sulfonic groups are equivalent.

2.2. The “Up”Substrate-Receptor Interactions

The effectivenesses of the monoazo, dis- and polyazo derivatives are dependent on different combinations of stereo-electronic effects: a direct action of the ortho or para –COOH group by –M (mesomeric) effect for compounds 5 and 26, lipophilic influences for 6, 12, 38, the –M effect of –NO2 in 20, 74 and 75, as well as the DAT effect in the case of 21 and 58. More energic electronic shifts can also appear when H-bonds are involved in the formation of adjacent cycles, where atoms exhibit different electronegativities, e.g., in: 74, 75, 89, 90, 119 and 122.
Congo Red (82) (characterized by a total symmetry), Direct Black 38 (83) (has four H-bonds) and 100 (characterized by DAT) have very close effectiveness values.
The distinguished effectiveness of 89* can be explained by its total symmetry and DAT. In case of 87 and 90 derivatives, the steric hindrances caused by the –SO3H groups located in the vicinity of the azo groups yield the lowering of the effectiveness by 1 log unit.
Except compound 104 whose azo:anilidic group ratio is 2:1, in the case of derivatives 119–122, 127 and 128, this ratio is equal to 2:2. In this situation, two anilide groups (e.g., in salicylanilide and diaminobenzanilide) are each able to activate an azo group, and thus the hydrolysis and the reduction processes occur fast, and the compounds’ toxicity is higher. As compared to compound 104 who’s carbonylic –M effect is intensified by its two hydroxy resorcinol groups, in the case of derivative 105 their influence is diminished by the coupling with 2-aminophenol-4-sulfonamide.
4-Methoxyazobenzene has a coplanar structure and, considering the influence of the positive mesomeric (+M) effect of the –OCH3 group, an extended conjugation favours the formation of the corresponding hydrazo derivative. Though the para alkoxy derivatives exhibit carcinogenic effects [25], the introduction of alkoxy groups in the ortho position of aminoazobenzene dyes decreases their mutagenic effect [26].

2.3. Azo-Heterocyclic Derivatives and Diazonium Salts

The presence of some heterocyclic components can influence differently the effectiveness. In a first analysis, the increased effectiveness exhibited by Sudan Black B (80) and the compound 112 could be assigned to their 2,3-dihydro-1H-perimidine (80), or pyrazolone rings (112), and in the case of derivatives which do not possess azo groups such as Tetrazolium Blue Chloride (95) and Fluorescent Brightener (103), to the combinations of tetrazolium-ortho-dianisidine type, as well as of triazinic-diaminostilbene-disulfonic acid type. The diazonium salts such as Fast Blue B (94), Fast Blue RR Salt (135) and Fast Blue BB Salt (136), can be characterized by a stronger electrophilic character, and this fact is probably due to the –M effect of the nitro group in case of the derivative Fast Red B Salt (137). The pyrazole-monoazo dye Tartrazine (55) is not toxic at all, even at a concentration of 0.11 mol/L, in accordance to its increased hydrophilicity.

3. Experimental Section

3.1. Test Substances

The test conditions and method were identical to those described in a previous work [9]. Some of the test substances (Table 1) were purchased from catalogues, or were synthesized at the Institute of Chemistry of Timisoara of the Romanian Academy, Romania (derivatives No.: 1, 2–4, 7–9, 12, 20, 21, 23, 25, 26, 77, 87, 96–98, 100–102, 104–122, 127–129, 132).

3.2. Test Organism

Colonies of H. echinata (Biologische Anstalt, Helgoland, Germany) were used to obtain eggs and larvae. The culture medium was artificial seawater (980 mosmol, pH 8.2, 18 °C). In laboratory an artificial metamorphosis can be synchronically started by the introduction of Cs+ ions or by using seawater without Mg2+ ions; it then lasts only 24 h. Under the action of external stimulus of Cs+ ions or Cs+ ions together with the tested compounds, one part of larvae further lives as such and another one is metamorphosized to the polyp form. The evaluation of the influence of the tested substance is very clear this way, the proposed method being based on this aspect.

3.3. Toxicity Test: Induction of Metamorphosis and Treatment with Test Substances

H. echinata larvae were exposed to seawater containing Cs+ and simultaneously one of the test substances for 3 h. The percentage of animals that underwent metamorphosis (development into polyps) was determined after 24 h. During the following days the frequency of inductions did not further increase. We chose a concentration of inducers which caused about three half to three quarters of the larvae to metamorphose in order to have conditions which are highly sensitive against an inhibitory influence. The concentration of the test substances (expressed in mol/L) was varied in such a way that we were able to determine the concentration at which the frequency of induction was reduced by 50% with respect to a control. This concentration was termed MRC50 (for Metamorphosis Reduction Concentration) and is similar to the effective EC50 concentration that gives half maximal effective response [9,23].

3.4. Theory/Calculation

The neutral molecular structures of azo dyes (Table 1) were modeled by the conformational search ability of the Omega v.2.4.3 (OpenEye Scientific Software, Santa Fe, NM. USA) program [27,28,29] (only structures having toxic effect on H. echinata test system were considered). SMILES notation was used as program input.
The following parameters were used for the conformer generation with Omega: a maximum of 400 conformers per compound, an energy cutoff of 10 kcal/mol, relative to a global minimum, which was identified from the search. The force field used was the 94s variant of the Merck Molecular force field (MMFF) with Coulomb interactions and the attractive part of the van der Waals interactions. To avoid redundant conformers, any conformer having a RMSD fit outside the range between 0.1 and 0.5 Å to another conformer was removed.
The minimum energy conformers thus obtained for each dye compound were used as input structures and were fully geometry optimized at restricted Hartree-Fock (3-21G basis set) level of theory (ab initio data are presented in Table S1, Supplementary Material) using the Gaussian 2009 software [30]. All optimized structures were characterized as true minima by frequency calculations (NImag = 0 for each compound).
Dye hydrophobicity, solubility and other molecular properties (Table S2, Supplementary Material) were evaluated by several programs. Thus, the logarithm of the octanol-water partition (logP) coefficient was calculated by the InstantJChem 5.12.4 (2013, Chemaxon) software [31], as well as other molecular properties: average polarisability, the number of donor, respectively acceptor H-bonds. Instant JChem was used for structure database management, search and prediction. The solubility (logS) was predicted by the AlogPS 2.1 program [32].

4. Conclusions

In conclusion, this study demonstrates the possibility of toxicity determination of some azo dyes and diazonium salts using the HeTS, which does not affect environmental equilibrium because the animals used in these experiments continue to live as larvae or polyps (as a result of metamorphosis). The included toxicity predictions illustrate the economy of time and research resources which could be done by adopting this procedure, in accordance to the “3Rs” concept as alternative method of investigation [33] or examination in advance, which leads to the replacement of other animal species in toxicity testing [34]. The test-system is characterized by reproducibility; it is fast, accessible, simple and offers a practical alternative in the research activity of different fields of interest, such as drugs or new classes of chemical derivatives. The main enzymatic reactions are the reduction of the azo group and the hydrolysis of the amido group. They are probably competitive and dependent on the limited or total coplanarity of the molecules, as well as by the existing electronic delocalization possibilities. Thus, it is possible that hydrolysis could occur prior to the reduction reaction, since the carbonyl anilide group is an active reaction center, as well as a permanent partner for Ser195. The S-R interaction, and therefore the effectiveness, are not dependent on the number of azo groups, but is influenced by the number of anilide groups. The toxicity is not dependent directly on the molecular dimension, as Protic and Sabljic [35] observed for commercial chemicals. The close Mlog/(1/MRC50) values lead to the determination of the average Clog(MRC50) value, which is characteristic for a class of compounds. One could emphasize the “class isotoxicity”. The reduction and the hydrolysis reaction products are mixtures of aromatic amines. The problem of the reaction mixture’s effectiveness and the influence of the methoxy substituents are not fully clarified. The calculated non-empiric molecular parameters could be used in different QSAR computations, especially for the determination of the molecules’ degree of penetration across the vascular endothelium into the brain (BBB). This issue could be one of the most important solutions in the research of new specific therapeutic agents [36].
The study does not involve clinical studies or patient data. Even if the biological differences between H. echinata and the superior organisms do not allow simple extrapolations, this test-system points out the toxicity trends at a cellular level. They represent more than surrogates for tests on superior organisms due to fact that it was discovered by experimental system with bacterium E. coli (Jacob and Monod, The Nobel Prize in Physiology or Medicine, 1965) that is fundamental to cellular regulation for all organisms, or: “what is valid for bacteria, is also valid for mammals” [37].

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/19/7/9798/s1.

Acknowledgments

This project was financially supported by Project 1.1 of the Institute of Chemistry Timisoara of the Romanian Academy. The authors are indebted to the OpenEye and Chemaxon companies for giving access to their software and to Dr. Markus Molis and Margret Krüß from Biologische Anstalt Helgoland (Alfred-Wegener-Institut für Polar- und Meersforschung, Germany) for offering us the possibility to continue the H. echinata experiments, and for the development of the “Köln-model” of toxity determinations. Gaussian acquisition was funded by CNCSIS-UEFISCSU project PN II-RU PD_502/2010.

Author Contributions

SAC and MM designed the study; SAC, SFT, CS, CT and GMS conducted the chemical calculations and conformational analysis; SAC, DI and CD realized the toxicity tests. All authors contributed to the draft of the article and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moir, D.; Masson, S.; Chu, I. Structure-activity relationship study on the bioreduction of azo dyes by Clostridium paraputrificum. Environ. Toxicol. Chem. 2001, 20, 479–484. [Google Scholar]
  2. Moller, P.; Wallin, H. Genotoxic hazards of azo pigments and other colorants related to 1-phenylazo-2-hydroxynaphthalene. Mut. Res. 2000, 462, 13–30. [Google Scholar] [CrossRef]
  3. Hawkins, B.T.; Egleton, R.D. Fluorescence imaging of blood-brain barrier disruption. J. Neurosci. Meth. 2006, 151, 262–267. [Google Scholar] [CrossRef]
  4. Tillement, J.-P.; Lecanu, L.; Papadopoulos, V. Amyloidosis and neurodegenerative diseases: Current treatments and new pharmacological options. Pharmacology 2010, 85, 1–17. [Google Scholar] [CrossRef]
  5. Lendel, C.; Bertoncini, C.W.; Cremades, N.; Waudby, C.A.; Vendruscolo, M.; Dobson, C.M.; Schenk, D.; Christodoulou, J.; Toth, G. On the mechanism of nonspecific inhibitors of protein aggregation: Dissecting the interaction of a α-synuclein with Congo red and lacmoid. Biochemistry 2009, 48, 8322–8334. [Google Scholar] [CrossRef]
  6. Zollinger, H. Color Chemistry; VCH Verlagsgesellschaft: Weinheim, Germany, 1987. [Google Scholar]
  7. Van den Mooter, G.; Maris, B.; Samyn, C.; Augustijns, P.; Kinget, R. Use of azo polymers for colon-specific drug delivery. J. Pharm. Sci. 1997, 86, 1321–1327. [Google Scholar] [CrossRef]
  8. Sigala, P.A.; Kraut, D.A.; Caaveiro, M.M.; Pybus, B.; Ruben, E.A.; Ringe, D.; Petsko, G.A.; Herschlag, D. Testing geometrical discrimination within an enzyme active site: Constrained hydrogen bonding in the ketosteroid isomeraze oxyanion hole. J. Am. Chem. Soc. 2008, 130, 13696–13708. [Google Scholar] [CrossRef]
  9. Chicu, S.A.; Funar-Timofei, S.; Simu, G.-M. Hydractinia echinata test system. II. SAR toxicity study of some anilide derivatives of Naphthol-AS type. Chemosphere 2011, 82, 1578–1582. [Google Scholar] [CrossRef]
  10. Skowronek, M.; Roterman Konieczny, L.; Stopa, B.; Rybarska, J.; Piekarska, B.; Gorecki, A.; Krol, M. The conformational characteristics of Congo red, Evans blue and Trypan blue. Comput. Chem. 2000, 24, 429–450. [Google Scholar] [CrossRef]
  11. Nam, S.; Tratnyek, P.G. Reduction of azo dyes with zero-valent iron. Wat. Res. 2000, 34, 1837–1845. [Google Scholar] [CrossRef]
  12. Platzek, T.; Lang, C.; Grohmann, G.; Gi, U.S.; Baltes, W. Formation of a carcinogenic aromatic amine from an azo dye by human skin bacteria in vitro. Hum. Exp. Toxicol. 1999, 18, 552–559. [Google Scholar] [CrossRef]
  13. Chung, K.-T.; Chen, S.-C.; Claxton, L.D. Review of the Salmonella typhimurium mutagenicity of benzidine, benzidine analogues, and benzidine-based dyes. Mutat. Res. 2006, 612, 58–76. [Google Scholar] [CrossRef]
  14. Stolz, A. Basic and applied aspects on the microbial degradation of azo dyes. Appl. Microbiol. Biotechnol. 2001, 56, 69–80. [Google Scholar] [CrossRef]
  15. Schwitzguebel, J.-P.; Aubert, S.; Grosse, W.; Laturnus, L. Sulphonated Aromatic Pollutants. Limits of Microbial Degradability and Potential of Phytoremediation. Environ. Sci. Pollut. Res. Int. 2002, 9, 62–72. [Google Scholar] [CrossRef]
  16. Preuß, G.T.; Ratte, H.T. Ökotoxikologische Charakterisierung von Nonylphenol Isomeren. UWSF – Z Umweltchem Okotox 2007, 19, 227–233. [Google Scholar] [CrossRef]
  17. Chicu, S.A.; Herrmann, K.; Berking, S. An approach to calculate the toxicity of simple organic molecules on the basis of QSAR analysis in Hydractinia echinata (Hydrozoa, Cnidaria). Quan. Struct.-Act. Rel. 2000, 19, 227–236. [Google Scholar] [CrossRef]
  18. Backhaus, T.; Scholze, M.; Grimme, L.H. The single substance and mixture toxicity of quinolones to the bioluminiscent bacterium Vibrio fischeri. Aquat. Toxicol. 1999, 49, 49–61. [Google Scholar] [CrossRef]
  19. Mekenyan, O.G.; Veith, G.D.; Bradbury, S.P.; Russom, C.L. Structure-toxicity relationships for α, β-unsaturated alcohols in fish. QSAR 1993, 12, 132–136. [Google Scholar]
  20. Levine, W.G. Metabolism of azo dyes: Implication for detoxification and activation. Drug Metab. Rev. 1991, 23, 253–309. [Google Scholar] [CrossRef]
  21. Ladbury, J.L. Just add water! The effect of water on the specificity of protein-ligand binding sites and its potential application to drug designe. Chem. Biol. 1996, 3, 973–980. [Google Scholar] [CrossRef]
  22. Gregory, P. Dyes and Dye Intermediates. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons.: New Jersey, USA, 2009; pp. 1–66. [Google Scholar] [CrossRef]
  23. Chicu, S.A.; Simu, G.M. Hydractinia eschinata test system. I. Toxicity determination of some benzenic, biphenylic and naphthalenic phenols. Comparative SAR-QSAR study. Rev. Roum. Chim. 2009, 54, 659–669. [Google Scholar]
  24. Zbaida, S.; Lewine, W.G. Characteristics of two classes of azo dye reductase activity associated with rat liver microsomal cytochrome P450. Biochem. Pharmacol. 1990, 40, 2415–2423. [Google Scholar] [CrossRef]
  25. Sztandera, L.; Trachman, M.; Bock, C.; Velga, J.; Garg, A. Soft computing in the design of nontoxic chemicals. J. Chem. Inf. Comput. Sci. 2003, 43, 189–198. [Google Scholar] [CrossRef]
  26. Esancy, J.F.; Freeman, H.S.; Claxton, L.D. The effect of alkoxy substituents on the mutagenicity of some aminoazobenzene dyes and their reductive-cleavage products. Mutat. Res. 1990, 238, 1–22. [Google Scholar]
  27. Nicholls, A. OpenEye Scientific Software. Available online: http://www.eyesopen.com (accessed on 31 January 2014).
  28. Hawkins, P.C.D.; Skillman, A.G.; Warren, G.L.; Ellingson, B.A.; Stahl, M.T. Conformer Generation with OMEGA: Algorithm and Validation Using High Quality Structures from the Protein Databank and Cambridge Structural Database. J. Chem. Inf. Model. 2010, 50, 572–584. [Google Scholar] [CrossRef]
  29. Hawkins, P.C.D.; Nicholls, A. Conformer Generation with OMEGA: Learning from the Data Set and the Analysis of Failures. J. Chem. Inf. Model. 2012, 52, 2919–2936. [Google Scholar]
  30. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  31. Chemaxon. Available online: http://www.chemaxon.com/ (accessed on 5 July 2014).
  32. Tetko, I.V.; Tanchuk, V.Y. Application of associative neural networks for prediction of lipophilicity in ALOGPS 2.1 program. J. Chem. Inf. Comp. Sci. 2002, 42, 1136–1145. [Google Scholar] [CrossRef]
  33. Vanhaecke, T.; Snykers, S.; Rogiers, V.; Garthoff, B.; Castell, J.V.; Hengstler, J.G. EU research activities in alternative testing strategies: Current status and future perspectives. Arch. Toxicol. 2009, 83, 1037–1042. [Google Scholar] [CrossRef]
  34. Vinken, M.; Pauwels, M.; Ates, G.; Vivier, M.; Vanhaecke, T.; Rogiers, V. Screening of repeated dose toxicity data present in SCC(NF)P/SCCS safety evaluations of cosmetic ingredients. Arch. Toxicol. 2012, 86, 405–412. [Google Scholar] [CrossRef]
  35. Protic, M.; Sabljic, A. Quantitative structure-activity relationships of acute toxicity of commercial chemicals on fathead minnows: Effect of molecular size. Aquat. Toxicol. 1989, 14, 47–64. [Google Scholar] [CrossRef]
  36. Vilar, S.; Chakrabarti, M.; Costanzi, S. Prediction of passive blood-brain partitioning: Straghtforward and effective classification models based on in silico derived physicochemical descriptors. J. Mol. Graph. Model. 2010, 28, 899–903. [Google Scholar] [CrossRef]
  37. Friedmann, H.C. From ‘Butyribacterium’ to ‘E. coli’: An Essay on Unity. Biochem. Perspect. Biol. Med. 2004, 47, 47–66. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds 98, 100, 106122 are available from the authors.

Share and Cite

MDPI and ACS Style

Chicu, S.A.; Munteanu, M.; Cîtu, I.; Şoica, C.; Dehelean, C.; Trandafirescu, C.; Funar-Timofei, S.; Ionescu, D.; Simu, G.M. The Hydractinia echinata Test-System. III: Structure-Toxicity Relationship Study of Some Azo-, Azo-Anilide, and Diazonium Salt Derivatives. Molecules 2014, 19, 9798-9817. https://doi.org/10.3390/molecules19079798

AMA Style

Chicu SA, Munteanu M, Cîtu I, Şoica C, Dehelean C, Trandafirescu C, Funar-Timofei S, Ionescu D, Simu GM. The Hydractinia echinata Test-System. III: Structure-Toxicity Relationship Study of Some Azo-, Azo-Anilide, and Diazonium Salt Derivatives. Molecules. 2014; 19(7):9798-9817. https://doi.org/10.3390/molecules19079798

Chicago/Turabian Style

Chicu, Sergiu Adrian, Melania Munteanu, Ioana Cîtu, Codruta Şoica, Cristina Dehelean, Cristina Trandafirescu, Simona Funar-Timofei, Daniela Ionescu, and Georgeta Maria Simu. 2014. "The Hydractinia echinata Test-System. III: Structure-Toxicity Relationship Study of Some Azo-, Azo-Anilide, and Diazonium Salt Derivatives" Molecules 19, no. 7: 9798-9817. https://doi.org/10.3390/molecules19079798

Article Metrics

Back to TopTop