Next Article in Journal
The Electronic Structure of the Nitrogen Atoms of Allyl (5- pyridin-2-yl-[1,3,4]-thiadiazol-2-yl)-amine
Previous Article in Journal
QSAR of Some N1-Aryl/Heteroarylaminomethyl/ethyl-1,2,4-Triazoles Part II: Antimicrobial Activity Against BacillusSubtilis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reaction of 2,2-Diphenyl-1-picrylhydrazyl with HO•, O2•, HO ,and HOO Radicals and Anions

1
Institute of Physical Chemistry “Ilie Murgulescu” of the Romanian Academy, Splaiul Independentei 202, 060021 Bucharest, Romania
2
Institute of Organic Chemistry “C. D. Nenitzescu” of the Romanian Academy, Splaiul Independentei 202B, Bucharest 15-256, Romania
3
National Agency for Medicinal Drugs, Aviator Sanatescu Street 48, Bucharest, Romania
4
University of Bucharest, Department of Physical Chemistry, Bulevardul Elisabeta 4-12, Bucharest, Romania
5
Texas A&M University, 5007 Avenue U, Galveston, TX 77551, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2006, 7(5), 130-143; https://doi.org/10.3390/i7050130
Submission received: 24 March 2006 / Accepted: 15 May 2006 / Published: 15 May 2006

Abstract

:
Using electronic absorption spectra and thin layer chromatography, the reaction of 2,2-diphenyl-1-picrylhydrazyl (DPPH) with O2•–, HO, HO, and HOO anions and free radicals revealed the formation of the para-nitro- and para-hydroxy-derivatives of 2,2-diphenyl-1-picrylhydrazine (DPPH-H) and of DPPH fragmentation products (diphenylamine, tetraphenylhydrazine). The reaction of DPPH with the O2•– anion-radical (from KO2 in benzene solution at room temperature in the presence of 18-crown-6 ether) is pseudo-first-order during the first 25 minutes.

1. Introduction

The free radical 2,2-diphenyl-1-picrylhydrazyl (DPPH) 1a [1,2,3] has found many applications due to its high stability and intense purple color that changes whenever it reacts [4,5,6,7,8,9,10,11]. Its reduction affords 2,2-diphenyl-1-picrylhydrazine (DPPH-H) 2a, or the corresponding anion, 3a (DPPH) in basic medium, as shown in Scheme 1. The DPPH radical acts as a scavenger for other odd-electron species which afford para-substitution products at phenyl rings: thus NO2 yields the mono-nitro-DPPH-H (2b) or a dinitro-DPPH-H [12,13,14,15]; the hydroxy free radical HO affords the hydroxy-DPPH-H (7) or its betainic oxidation product 8 [16]; halogen atoms can also be trapped similarly [4,13,14]. Other reactions that have been observed involve polynitroaniline [17,18], ipso-substitution of a nitro group of the picryl moiety, and meta-attack on this group [19].
Scheme 1. Redox and protolytic reactions for compounds 1, 2 and 3.
Scheme 1. Redox and protolytic reactions for compounds 1, 2 and 3.
Ijms 07 00130 g003
The biochemical process known as oxidative stress involves chemical species presented in eq. 1 [20,21,22].
Ijms 07 00130 i001
In the oxidative stress syndrome, H2O2 as well as the radicalic species HO and O2•– are supposed to play an important part. Enzymes such as catalase, xanthine-oxidase, peroxidase, and superoxide-dismutase convert such species into less aggressive compounds. If DPPH (1a) may be used for monitoring some of the above reactions as indicated in the literature [11,16,23], a study of each reaction should bring interesting information. Depending on the pH, the anions may interact with DPPH via redox processes (electron transfer) [8,11,15,19,23,24,25,26,27].
For exploring the possibility of using colorimetric methods in monitoring such processes, we investigated the reaction of DPPH with the HO radical, O2•– radical-anion, HO, and HOO anions using qualitative and densitometric thin-layer chromatography (TLC) for detecting and separating the reaction products, and spectrophotometric methods for following the reaction kinetics.

2. Results and discussion

All investigations were carried out in the same solvent (benzene), either in homogeneous or heterogeneous—liquid/liquid (ℓ/ℓ) or solid/liquid systems (s/ℓ). Spectral methods were used directly with the mixture of reaction products in the benzene solution, whereas TLC separations were carried out after extracting the benzene solution with 1N hydrochloric acid.
The reactions of DPPH with the three anions (O2•–, HO, and HOO) will be reported sequentially.

2.1. Reaction with potassium superoxide in the presence of crown ether 18C6

The O2•– radical-anion [28,29,30,31,32,33] may act either as reducing or oxidizing agent in the presence of copper-containing enzymes such as superoxide-dismutase. Under the influence of this enzyme the radical anion undergoes a dismutation into O2 and H2O2 (eqs. 2 and 3).
Ijms 07 00130 i002
Ijms 07 00130 i003
On using commercial potassium superoxide (K+ O2•–) [31] in the presence of crown ether 18C6 one obtains a homogeneous benzene solution in which the nucleophilic reactivity of the anion is enhanced [30,34,35,36,37]. Preliminary investigation revealed that the intensity of the characteristic absorption band of DPPH (1a) at λmax = 520 nm decreases due both to the formation of the anion (3a) with λmax = 428 nm [8,38] via a redox process involving the superoxide radical-anion (eqs. 4 and 5), and to other reactions. In all equations, subscripts s, o and w denote solid, organic (benzene) and, aqueous phases, respectively.
Ijms 07 00130 i004
Ijms 07 00130 i005
A kinetic study was performed. Determinations were carried out at 25ºC with benzene solutions of DPPH and 18C6, and solid KO2 in molar ratios 1a:18C6:KO2 that were 1:1:3, 1:5:5, and 1:5:9 by monitoring the DPPH λmax = 520 nm band. An excess of KO2 (hence also of 18C6) is needed. After 24 h there is no further decrease of 1a absorption (Table 1) and the concentration of the anion 3a reaches a plateau.
The spectrophotometric study revealed two isosbestic points at 344 and 496 nm (Fig.1).
During the first 25 minutes, the reaction is pseudo-first-order, as seen from Fig. 2. Compounds 2b, 4, and 6 appear after about 30 minutes.
Table 1. Kinetics of unreacted DPPH (1a) in the reaction with 18C6 and KO2 (molar ratio = 1 : 5 : 5).
Table 1. Kinetics of unreacted DPPH (1a) in the reaction with 18C6 and KO2 (molar ratio = 1 : 5 : 5).
Time
(min)
Unreacted DPPH (%)
595.66
2593.59
4591.94
6590.70
9087.40
12083.05
1440 (24h)71.90
Figure 1. Monitoring by UV-VIS in benzene the formation of 3amax = 428 nm) from 1amax = 520 nm) as described by eqs. 4 and 5 (for molar ratio 1a : 18C6 : KO2 = 1:5:5).
Figure 1. Monitoring by UV-VIS in benzene the formation of 3amax = 428 nm) from 1amax = 520 nm) as described by eqs. 4 and 5 (for molar ratio 1a : 18C6 : KO2 = 1:5:5).
Ijms 07 00130 g001
Figure 2. Pseudo-first order kinetics of the reaction between DPPH and KO2 in benzene in the presence of 18C6 at 25ºC (the first 25 minutes).
Figure 2. Pseudo-first order kinetics of the reaction between DPPH and KO2 in benzene in the presence of 18C6 at 25ºC (the first 25 minutes).
Ijms 07 00130 g002
The reaction (eqs. 4 and 5) was also investigated by qualitative and densitometric TLC, allowing the observation of secondary products that are not detected by spectrophotometry. During the first 30 minutes, secondary products are not detectable either in the crude reaction mixture or after acid extraction. However, after 2 h the formation of mononitro-DPPH-H (2b) and of diphenylamine (4) could be observed, and after 24 h also tetraphenylhydrazine (6) was found. Neither the bis-(p-nitro)- DPPH-H (3b) nor the p-hydroxy- DPPH-H (7) and/or its oxidized betainic form (8) could be detected. The identity of these two products was certified by comparing the Rf values in TLC and the NMR spectra (1H and 13C) with those of authentic compounds.
Scheme 2. Redox process involving diphenylamine (4) and tetraphenylhydrazine (6).
Scheme 2. Redox process involving diphenylamine (4) and tetraphenylhydrazine (6).
Ijms 07 00130 g004
Scheme 3. Redox process involving 4-hydroxy- DPPH-H (7) and the betainic oxidized product (8).
Scheme 3. Redox process involving 4-hydroxy- DPPH-H (7) and the betainic oxidized product (8).
Ijms 07 00130 g005
In Table 2 we present the products that have been identified by quantitative (densitometric) TLC analysis after 24 h at room temperature by treating DPPH (1a) with: HO (from benzyltrimethylammonium hydroxide, reaction A); with HO (from the photochemical reaction of ortho-hydroxy-acetophenoneoxime also called FotoFentonTM2, reaction B), with O2•– (from potassium superoxide, reaction C), and with HOO (from hydrogen peroxide, reaction D).
Table 2. Quantitative TLC data on the reactions of DPPH (1a) with HO, HO, O2•− and HOO in benzene after 24 h.
Table 2. Quantitative TLC data on the reactions of DPPH (1a) with HO, HO, O2•− and HOO in benzene after 24 h.
Reaction cDPPH-H b
2a
O2N-DPPH-Hb
2b
Ph2NHa
4
Ph2N-NPh2a
6
HO-DPPH-H b
7
Betaineb
8
Rf%Rf%Rf%Rf%Rf%Rf%
A (HO)0.67460.44220.6810.5980.046
B (HO)0.68200.31120.44250.57110.049
C (O2•–)0.67650.30120.4340.670.4
D (HOO)0.67700.29100.4430.035
a,b TLC on silica gel: a without acid extraction using as eluent n-hexane : toluene = 7 : 3 (v/v); b after extraction with 1N hydrochloric acid using toluene as eluent.c Reaction A: 1a + benzyltrimethylammonium hydroxide (molar ratio 1:2); Reaction B: 1a + hν + hydroxy-acetophenoneoxime (molar ratio 1:2); Reaction C: 1a + 18C6 + KO2 (molar ratio 1:5:5); Reaction D: 1a + Kryptofix 222 + H2O2 (molar ratio 1:2:2).
The formation of DPPH anion (3a) and redox processes may explain the production of the observed products, as indicated by eqs. (6) – (9).
Ijms 07 00130 i006
Ijms 07 00130 i007
Ijms 07 00130 i008
Ijms 07 00130 i009
One may speculate that DPPH (1a) may split into the diphenylamino radical (5) that dimerizes [1,39,40] to tetraphenylhydrazine (6), and dinitrobenzofuroxan (from picrylnitrene), but as yet there is no experimental evidence for this conjecture. However, we intend to look for such evidence among the many side products detected by TLC. Also, σ-Meisenheimer complexes (eq. 7) may be involved in the formation of some of the observed reaction products [23,24,41].

2.2. Reaction of DPPH with the hydroxide anion and the hydroxy free radical

Previous papers reported that on treating DPPH (1a) with hydroxide anions (from powdered alkali hydroxides in the presence of crown ethers or polyethylene glycol) a redox process takes place affording hydroxyl free radicals (HO) and resulting in the formation of 4-nitro-DPPH-H (2b) and 4-hydroxy-DPPH-H (7) [25,42]. Also the reaction between DPPH and photolytically-generated HO [43] radicals was investigated, leading to 4-hydroxy-DPPH-H (7) and its betainic oxidized product (8), which were isolated [16].
The present paper reports results for the reaction between DPPH and HO radicals generated by three methods: (i) photochemically, (ii) from solid potassium hydroxide, DPPH, and 18C6, and (iii) from DPPH, 18C6, and benzyltrimethylammonium hydroxide.
(i)
As shown in Table 2, in the photochemical reaction, in addition to compounds 2b, 7, and 8, diphenylamine (4) was also identified and determined quantitatively.
(ii)
From the reaction with solid potassium hydroxide the same four products were detected qualitatively.
(iii)
The reaction with benzyltrimethylammonium hydroxide afforded 2b, 6, 7, and tetraphenylhydrazine (4).
Eqs. (7 – 9) and (12) make plausible the formation of NO2 radicals, which could explain the formation of 2b from DPPH. Then eqs. (10 – 14) could explain the formation of compounds 4 and 6 (along perhaps with dinitro-benzofuroxan among the products).
Ijms 07 00130 i010
Ijms 07 00130 i011
Ijms 07 00130 i012
Ijms 07 00130 i013
Ijms 07 00130 i014

2.3. Reaction of DPPH with the hydrogen peroxide anion

As discussed above, eq. 1 indicates that hydrogen peroxide is formed in living cells, and rapidly decomposed by the defense enzymatic system (catalase, peroxidase). Numerous reagents may be used for assaying hydrogen peroxide and derived oxidizing species in organisms [43]. From the reaction of DPPH with the Fenton reagent (hydrogen peroxide and ferrous sulfate, which afford hydroxy free radicals HO), quantitative TLC determinations certified the formation of mononitro- and monohydroxy-DPPH-H derivatives 2b and 7, respectively [16].
Now we report the reaction of DPPH (1a) with the HOO anion formed from hydrogen peroxide and the basic compound Kryptofix 222 (4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo[8.8.8] hexacosane), abridged here as K222, (eq. 15) [44]. The reaction was carried out with DPPH and K222 (molar ratios 1:2) with an excess of 30% H2O2 in a liquid/liquid biphasic system (water/benzene) at room temperature under stirring. During the first three hours the process was monitored by qualitative TLC, and after 48 h by quantitative TLC. The mono-nitro derivative 2b and the betaine 8 were identified after 3 h, implying that the mono-hydroxy derivative 7 must have been formed initially (Scheme 3). After 24 h, in addition to 2b and 8, the presence of diphenylamine (4) was detected, and all these three products were assayed quantitatively (Table 2). These results may be explained as being due to attack on DPPH by the HOO / HO radicals or the HOO / HO anions formed by K222, the latter being basic enough to react with hydrogen peroxide or water and the corresponding anions then becoming oxidized by DPPH (eqs. 15 – 18). Finally, the reactions between DPPH and the free radicals afford the observed products (eqs. 11 – 14 and 19 – 20) [44].
Ijms 07 00130 i015
Ijms 07 00130 i016
Ijms 07 00130 i017
Ijms 07 00130 i018
Ijms 07 00130 i019
Ijms 07 00130 i020

2.4. Comparison between reactions of DPPH with O2•–, HO, HO, and HOO

On investigating the room-temperature reaction of the stable free radical DPPH with oxidizing species O2•–, HO, and HOO that appear in biological processes, it was found that in all cases the electronic absorption spectra could provide information during the first 25 minutes, but TLC methods were more adequate for monitoring the reaction at longer times, when various products interfered with the spectral determinations.
The concentration decrease of the starting material is highest for HOO and lowest for HO, leading to the following reaction rate order: HOO > O2•– > HO > HO. Excepting the reaction A with hydroxide anions, the mono-nitro-DPPH-H compound (2b) was formed in comparable concentrations in all experiments (Table 2). Diphenylamine (4) was formed in different amounts: HO ≈ HO > O2 ≈ HOO, and the derived tetraphenylhydrazine (6) was identified only in very small concentrations in the reactions with O2•– and HO. The mono-hydroxy-DPPH-H compound (7) and its betainic oxidized derivative (8) were detected in low amounts in all reactions except for reaction C with O2•–. It may be possible that all products were always formed, but in some cases at lower amounts than could be detected by TLC techniques.
A mechanism involving σ-Meisenheimer complexes of DPPH with the hydride anion was proposed earlier [41]. In Scheme 4 we propose two distinct mechanisms for the reactions reported in the present paper. On the left-hand side, the homolytic attack of the hydroxyl radical HO explains the formation of mono-hydroxy-DPPH-H (7) and its oxidized betainic derivative (8). A different pathway for the formation of compounds 7 and 8 is suggested by eq. 20. On the right-hand side the nucleophilic attack of anions Nu: (HOO and HO or radical-anion O2•–) may lead to the formation [16,23] of an unstable σ-Meisenheimer complex [DPPH, 2Nu:]. In turn, this complex is supposed to cleave either a nitro group explaining the formation of the mono-nitro-DPPH-H (2a), or (supposedly) dinitro-benzofuroxan and diphenylamine (4), which is then further oxidized to tetraphenylhydrazine (6).
Scheme 4. Mechanisms proposed for the reactions of DPPH with HO free radicals and nucleophilic agents (Nu:).
Scheme 4. Mechanisms proposed for the reactions of DPPH with HO free radicals and nucleophilic agents (Nu:).
Ijms 07 00130 g006
Taking into account that DPPH is able to abstract a hydrogen atom from amines with a significant reaction rate, as determined by K. U. Ingold and his coworkers [45], even more complicated mechanisms may be envisaged, in addition to the process converting 4 into 6.

3. Conclusions

Several conclusions can be drawn from the study of reactions between DPPH and anions or oxidizing species: O2•–, HO, and HOO. Redox processes involving DPPH convert such anions into oxidants (HO, HOO). The formation of minor amounts of reaction products can be detected well by TLC, but less well (and only during the first half-hour at room temperature) by spectrophotometric techniques. The following reactions products were detected and assayed quantitatively: DPPH-H (2a), its mono-nitro- (2b) and mono-hydroxy-derivatives (7), the betainic oxidized form (8) of the latter, diphenylamine (4) and tetraphenylhydrazine (6) resulted by cleavage and oxidation of DPPH. The reaction mechanisms are fairly complicated, involving protolytic equilibria and redox processes, accompanied by molecular fragmentations.

4. Experimental part

Commercial chemicals were purchased from Aldrich (DPPH and KO2), Merck (18C6, Kryptofix-222, diphenylamine, benzyltrimethylammonium hydroxide, glass plates with silica gel GF254 silanized and nonsilanized), and Molecular Probes (ortho-hydroxy-acetophenoneoxime – FotoFentonTM2). The mono-nitro-DPPH-H (2b) [46], the solid supramolecular complex [18C6…K]+DPPH [47], and tetraphenylhydrazine (6) [39], were prepared according to literature data.
Electronic absorption spectra were recorded with a Unicam-UV-VIS spectrophotometer using “Vision Software V.3.33”. NMR spectra were recorded with a Varian Gemini 300 BB spectrometer (300 MHz for 1H-NMR and 75 MHz for 13C-NMR) using TMS as internal standard. We used Camag Software 1992 scanner II – Switzerland for densitometric TLC analysis.

Reactions of 2,2-diphenyl-1-picrylhydrazyl, DPPH (1a) with KO2, KOH, benzyltrimethylammonium hydroxide, FotoFentonTM2, and H2O2 at room temperature.

A1. The reaction with hydroxide anions from solid KOH in the presence of 18C6 was carried out in benzene for 2 hrs under stirring with molar ratio 1a:18C6:KOH = 1:2:2. After liquid/liquid (ℓ/ℓ) extraction with 1N hydrochloric acid, the separated organic layer was dried with anhydrous sodium sulfate and analyzed by TLC (silica gel, toluene). For the quantitative TLC analysis the chromatograms were scanned at λmax = 254 nm, and the densitometric results are presented in Table 2.
A2. The reaction with hydroxide anions from benzyltrimethyl ammonium hydroxide was carried out in benzene under stirring wither for 2 hrs or 48 hrs with molar ratio between DPPH and the quaternary ammonium salt 1:2. Without ℓ/ℓ extraction, the solution was analyzed qualitatively and quantitatively by scanning (λmax = 254 nm) and results are indicated in Table 2.
B. The reaction with hydroxy free radicals (HO) was performed in benzene under stirring using DPPH and ortho-hydroxy-acetophenoneoxime (FotoFentonTM2) with ultraviolet irradiation in a quartz flask. Without ℓ/ℓ extraction, the solution was analyzed qualitatively and quantitatively by scanning (λmax = 254 nm) and results are indicated in Table 2.
C. The reaction with solid KO2 in the presence of 18C6 was carried out in benzene under stirring with three molar ratios of the reactants: 1a:18C6:KO2 = 1:1:3, 1:5:5, and 1:5:9. Spectrophotometric analysis for the decrease of DPPH concentration (λmax = 520 nm) and the increase of DPPH concentration (λmax = 428 nm) allowed us to follow the reaction kinetics (Figure 1 and Figure 2). Product formation was analyzed by two methods: (i) without ℓ/ℓ extraction by TLC (silanized silica gel, n-hexane; and nonsilanized silica gel, n-hexane:toluene 7:3 v/v); (ii) with ℓ/ℓ extraction, as for reaction A.
D. The reaction with HOO anions from H2O2 and K222 was performed in a biphasic benzene-aqueous ℓ/ℓ system under stirring with molar ratio 1a:K222:H2O2 = 1:2:2. After 2 hrs, the organic phase was separated, dried, and without ℓ/ℓ extraction the solution was analyzed qualitatively and quantitatively by scanning (λmax = 254 nm); the results are indicated in Table 2.

Kinetics of the reaction between DPPH (1a) and potassium superoxide.

For studying the kinetics of the reaction 1a + 18C6 + KO2, the temperature was fixed at 25ºC, and two concentrations were kept constant (1a and 18C6); the concentration of potassium superoxide was varied according to molar ratios 1a:18C6:KO2 = 1:5:2; 1:5:3; 1:5:4 and 1:5:5. The formation of the DPPH anion (3a) was monitored at λmax = 428 nm. For determining the order of the reaction, the averages of five results for each of the five spectral measurements were plotted as kobs vs. concentration of KO2 (Fig. 2) according to eq. 21.
kobs = 1/t ln [a/(ax)]
where t is the time (seconds), a is the initial concentration of 3a (absorbance A0 at t = t0, and (ax) is the concentration at time t.

Reaction of the supramolecular complex [18C6…K]+DPPH with potassium superoxide.

To the supramolecular complex [18C6…K]+DPPH [46] dissolved in benzene, solid KO2 and 18C6 (molar ratio Complex:18C6:KO2 = 1:4:5). The concentrations of 1amax = 520 nm) and of the DPPH anion 3amax = 428 nm) were monitored spectrophotometrically. Qualitative TLC analysis after 2h revealed the presence of DPPH (1a) and diphenylamine (4) using silanized silica gel (n-hexane), nonsilanized silica gel (n-hexane:toluene, 7:3 v/v) and nonsilanized silica gel (toluene).

NMR Identification of reaction products mono-nitro-DPPH-H (2b), diphenylamine (4), and tetraphenylhydrazine (6).

Using reactions that afforded significant yields of products 2b, 4, and 6 (Table 2) and preparative conditions (starting with 0.1 g of DPPH), the reaction products were separated and purified as follows: for 2b, the preparative TLC plate was extracted with a mixture of methylene chloride and methanol 9:1 v/v. For 4 and 6 the extractions were performed with methylene chloride and the solvent was removed under vacuum and in argon atmosphere in order to avoid oxidation by air oxygen.
Diphenylamine 4: 1H-NMR(CDCl3, δ ppm, J Hz): 5.61(bs, 1H, NH, deuterable); 6.90(tt, 2H, H-para, 1.1, 7.3); 7.03(dd, 4H, H-ortho, 1.1, 8.5); 7.23(dd, 4H, H-meta, 7.3, 8.5); 13C-NMR(CDCl3, δ ppm): 143.08(C-1); 129.27(C-meta); 120.92(C-para); 117.79(C-ortho).
Tetraphenylhydrazine 6 : 1H-NMR(CDCl3, δ ppm, J Hz): 6.98(tt, 1H, H-para, 1.1, 7.2); 7.20(dd, 2H, H-meta, 7.2); 7.31(dd, 2H, H-ortho, 7.2, 8.2); 13C-NMR(CDCl3, δ ppm): 142.91(C-1); 129.07(C-meta); 122.08(C-para); 118.18(C-ortho).

Acknowledgements

We express our thanks to Molecular Probes for samples of FotoFentonTM2.

References

  1. Forester, A. R.; Hay, J. M.; Thomson, R. H. Organic Chemistry of Free Radicals; Academic Press: London, 1968; pp. 110, 137, and references cited therein. [Google Scholar]
  2. Balaban, A. T. Stable nitrogen free radicals. Rev. Roum. Chim. 1971, 16, 725–737, and references cited therein. [Google Scholar]
  3. Hazell, J. E.; Russel, K. E. The reaction of 2,2-diphenyl-1-picrylhydrazyl with secondary amines. Can. J. Chem. 1958, 36, 1729–1734. [Google Scholar] [CrossRef]
  4. Gille, L.; Prosch, U.; Stoesser, R. Reactions of 1,1-diphenyl-2-picrylhydrazyl with NO2, Cl2 and F2. Radiat. Phys.Chem. 1992, 40, 461–468. [Google Scholar]
  5. Fargere, T.; Abdennadher, M.; Delmas, M.; Boutevin, B. Determination of peroxides and hydroperoxides with 2,2-diphenyl-1-picrylhydrazyl (DPPH). Application to ozonized ethylene vinyl acetate copolymers (EVA). Eur. Polym. J. 1995, 31, 489–497. [Google Scholar] [CrossRef]
  6. Nanjo, F.; Goto, K.; Seto, R.; Suzuki, M.; Sakai, M.; Hara, Y. Scavenging effects of tea catechins and their derivatives on 1,1-diphenyl-2-picrylhydrazyl radical. Free Radicals Biol. Med. 1996, 21, 895–902. [Google Scholar] [CrossRef]
  7. Yokozawa, T.; Chen, C. P.; Dong, E.; Tanaka, T.; Nonaka, G.; Nishioka, I. Study on the inhibitory effect of tannins and flavonoids against the 1,1-diphenyl-2-picrylhydrazyl radical. Biochem. Pharmacol. 1998, 56, 213–222. [Google Scholar] [CrossRef]
  8. Popescu, D. O.; Ionita, P.; Zarna, N.; Covaci, I.; Stoica, A.; Zarna, A.; Nourescu, D.; Spafiu, F.; Badea, F.; Luca, C.; Caproiu, M. T.; Constantinescu, T.; Balaban, A. T. Hydrazyl and aminyl analogs, liquid membranes and stationary phases for gas-chromatography. Roum. Quart. Rev. 1998, 6, 271–282, and references cited therein. [Google Scholar]
  9. Nanjo, F.; Mori, M.; Goto, K.; Hara, Y. Radical scavenging activity of tea catechins and their related compounds. Biosci. Biotechnol. Biochem. 1999, 63, 1621–1623. [Google Scholar] [CrossRef]
  10. Sawai, Y.; Moon, J. H.; Sakata, K.; Watanabe, N. Effects of structure on radical-scavenging abilities and antioxidative activities of tea polyphenols: NMR analytical approach using 1,1-diphenyl-2-picrylhydrazyl radicals. J. Agric. Food Chem. 2005, 53, 3598–3604. [Google Scholar] [CrossRef]
  11. Ionita, P.; Gilbert, B. C.; Whitwood, A. C. Generation of oxygen-, sulfur, carbon-, nitrogen- and phosphorus-centred short-lived radicals via one-electron oxidation with stable hydrazyl radical. Org. Chem. Lett. 2004, 1, 70–74. [Google Scholar] [CrossRef]
  12. Weil, J. A.; Sane, K. V.; Kinkade, J. M. The reaction between 2,2-diphenyl-1-picrylhydrazyl and nitrogen dioxide. J. Phys. Chem. 1961, 65, 710–712. [Google Scholar] [CrossRef]
  13. Currie, P. F.; Quail, J. W.; Rusk, A. C. M.; Weil, J. A. A study of Br/NO2 substitution reactions and nuclear magnetic resonance of 2,2-diphenyl-1-picrylhydrazines. Can. J. Chem. 1983, 61, 1760–1765. [Google Scholar] [CrossRef]
  14. Currie, P. F.; Quail, J. W.; Weil, J. A. The reaction between 2,2-diphenyl-1-picrylhydrazyl and bromine. Can. J. Chem. 1980, 58, 723–726. [Google Scholar] [CrossRef]
  15. Caproiu, M. T.; Covaci, I. C.; Ionita, P.; Luca, C.; Constantinescu, T.; Balaban, A. T. Regioselectively [15NO2] labeled N-methoxypicramide and DPPH prepared by using crown ether and solid sodium [15N]-nitrite. J. Nucl. Radioanal. Chem. 1999, 242, 773–776. [Google Scholar] [CrossRef]
  16. Hristea, E. N.; Hillebrand, M.; Caproiu, M. T.; Caldararu, H.; Constantinescu, T.; Balaban, A. T. Scavenging the hydroxyl radical by 2,2-diphenyl-1-picrylhydrazyl. Arkivoc. 2002, (ii), 123–132. [Google Scholar]
  17. Constantinescu, T.; Caproiu, M. T.; Zarna, N.; Caragheorgheopol, A.; Caldararu, H.; Stanciuc, G.; Radu, M.; Badescu, V.; Balaban, A. T. Reaction of 2,2-diphenyl-1-picrylhydrazyl (DPPH) with N-alkoxy-di- and tri-nitroanilines. New J. Chem. 1997, 21, 575–579. [Google Scholar]
  18. Covaci, I. C.; Constantinescu, T.; Caproiu, M. T.; Caldararu, H.; Caragheorgheopol, A.; Constantinescu, T.; Balaban, A. T. New congeners of 1-picryl-2-phenyl-2-(para-picramidophenyl)-diazenium betaine whose picramido groups are replaced by 4-cyano-2,6-dinitrophenyl analogs. Pol. J. Chem. 2001, 75, 1427–1440. [Google Scholar]
  19. Ionita, P.; Constantinescu, T.; Caldararu, H.; Luca, C.; Caproiu, M. T.; Dumitrascu, F.; Silberg, I.; Balaban, A. T. Reaction between the DPPH free radical and potassium cyanide in the presence of crown ether 18-C-6: a correction. Rev. Roum. Chim. 1999, 44 (4), 393–396. [Google Scholar]
  20. Yoshikawa, T.; Yamamoto, Y.; Naito, Y. Free Radicals in Chemistry, Biology and Medicine; Oica International: London, 2000. [Google Scholar]
  21. Albert, M. G.; Rousselot, D. B.; Abedinzadeh, Z.; Jore, D. Especes réactives de l`oxygène: Comment l’oxygène peut-il devenir toxique? L′Actualité Chimique 2003, 11-12, 91–96. [Google Scholar]
  22. Favier, A. Le stress oxydant: Intérêt conceptuel et expérimental dans la compréhension des mécanismes des maladies et potentiel thérapeutique. L′Actualité Chimique 2003, 11-12, 108–115. [Google Scholar]
  23. Ionita, P. Is DPPH stable free radical a good scavenger for oxygen active species? Chem. Papers 2005, 59, 11–16. [Google Scholar]
  24. Ionita, P.; Constantinescu, T.; Caldararu, H.; Luca, C.; Caproiu, M. T.; Balaban, A. T. Reaction between the DPPH free radical and potassium cyanide in the presence of crown ether 18-C-6. New J. Chem. 1997, 21, 511–515. [Google Scholar]
  25. Luca, C.; Ionita, P.; Constantinescu, T.; Caldararu, H.; Caragheorgheopol, A.; Caproiu, M. T. Interphase processes involving hydrazine derivatives. III. The reaction of 2,2-diphenyl-1-picrylhydrazyl with HO- anion in the presence of crown ethers. Rev. Roum. Chim. 1997, 42, 985–992. [Google Scholar]
  26. Ionita, P.; Caproiu, M. T.; Luca, C.; Constantinescu, T.; Caldararu, H.; Balaban, A. T. Selective [15N] nitration of 2,2-diphenyl-1-(2,4- or 2,6-dinitrophenyl)-hydrazines or hydrazyls. J. Labelled. Comp. Radiopharm. 1998, 41, 791–799. [Google Scholar] [CrossRef]
  27. Ionita, P.; Spafiu, F.; Constantinescu, T.; Caldararu, H. The reaction of 2,2-diphenyl-1-picrylhydrazyl stable free radical with sodium tetra-phenylborate in the presence of 18-C-6 crown ether. Rev. Roum. Chim. 1999, 44, 497–500. [Google Scholar]
  28. Fridovich, I. Superoxide radical and superoxide dismutase. Acc. Chem. Res. 1972, 5, 321–326. [Google Scholar] [CrossRef]
  29. Fee, J. A.; Hildenbrand, G. On the development of a well-defined source of superoxide ion for studies with biological systems. FEBS Lett. 1974, 39, 79–82. [Google Scholar] [CrossRef]
  30. Valentine, J. S.; Curtis, A. A convenient preparation of solutions of superoxide anion and the reaction of superoxide anion with a cooper(II) complex. J. Amer. Chem. Soc. 1975, 97, 24–26. [Google Scholar]
  31. Sawyer, D. T.; Gibian, M. J. The chemistry of superoxide ion. Tetrahedron. 1979, 35, 1471–1481. [Google Scholar] [CrossRef]
  32. Saran, M.; Bors, W. Signalling by O2•- and NO: how far can either radical, or any specific reaction product, transmit a message under in vivo conditions. Chem.-Biol. Interact. 1994, 90, 35–45. [Google Scholar] [CrossRef]
  33. Davies, M. J.; Timmins, G. S. Biological free radicals. Electron Paramagn. Reson. 2000, 17, 1–42. [Google Scholar] [CrossRef]
  34. Vogtle, F. Supramolecular Chemistry; Wiley: Chichester, 1991; p. 27. [Google Scholar]
  35. Jackman, L. M.; Lange, B. C. Structure and reactivity of alkali metal enolates. Tetrahedron 1977, 33, 2737–2769. [Google Scholar] [CrossRef]
  36. Weber, E.; Toner, J. L.; Goldberg, I.; Vogtle, F.; Laidler, D. A.; Stoddart, J. F.; Bartsch, R. A.; Liotta, C. L. Crown Ethers and Analogs; John Wiley & Sons, 1989. [Google Scholar]
  37. Gokel, G. W. Crown Ethers and Cryptands; The Royal Society of Chemistry: London, 1991. [Google Scholar]
  38. Weil, J. A.; Janusonis, G. A. The alkali salts of 2,2-diphenyl-1-picrylhydrazine. J. Org. Chem. 1962, 27, 1248–1250. [Google Scholar] [CrossRef]
  39. Wieland, H.; Gambarjan, S. Die oxidation des diphenylamins. Ber. 1906, 39, 1499–1506. [Google Scholar] [CrossRef] [Green Version]
  40. Jackson, C.; Patel, N. K. D. Free radicals derived from phenothiazine. Tetrahedron lett. 1967, 24, 2255–2263. [Google Scholar] [CrossRef]
  41. Hristea, E. N.; Hillebrand, M.; Radutiu, A. C.; Caldararu, H.; Caproiu, M. T.; Ionita, P.; Constantinescu, T.; Balaban, A. T. Reaction of 1,1-diphenyl-2-picryl-hydrazyl or hydrazine with sodium borohydride in the presence of 18-crown-6 ether. Rev. Roum. Chim. 2000, 45, 1089–1096. [Google Scholar]
  42. Luca, C.; Ionita, P.; Constantinescu, T. The reaction of free stable 2,2-diphenyl-1-picrylhydrazyl radical with KOH in presence of polyethylene glycols. Rev. Roum. Chim. 1997, 42, 683–685. [Google Scholar]
  43. Haugland, R. P. Handbook of Fluorescent Probes and Research Chemicals; Sixth edition: Netherlands, 1996; p. 485. [Google Scholar]
  44. Luca, C.; Ionita, P.; Constantinescu, T. Interphase processes involving hydrazine derivatives. VII. The reaction of DPPH with H2O and H2O2 in presence of cryptands. Rev. Roum. Chim. 1998, 43, 129–131, and references cited therein. [Google Scholar]
  45. Mulder, P.; Litwinenko, G.; Lin, S.; MacLean, P.; Barclay, L. R. C.; Ingold, K. U. Chem. Res. Toxicol. 2006, 19, 79–85. [CrossRef] [Green Version]
  46. Luca, C.; Ionita, P.; Caproiu, M. T.; Caldararu, H.; Constantinescu, T. Interphase processes involving hydrazine derivatives. VI. The syntheses of 2-(p-nitrophenyl)-2-phenyl-1-picrylhydrazine, 2,2-bis(p-nitrophenyl)-1-picrylhydrazine, and their 15N labeled congeners. Roum. Chim. 1998, 43, 221–224. [Google Scholar]
  47. Luca, C.; Ionita, P.; Caproiu, M. T.; Constantinescu, T. Interphase processes involving hydrazine derivatives. IV. The preparation of solid supramolecular complexes between crown ethers, potassium and 2,2-diphenyl-1-picrylhydrazyl anion. Rev. Roum. Chim. 1998, 43, 25–30. [Google Scholar]

Share and Cite

MDPI and ACS Style

Hristeac, E.N.; Caproiu, M.T.; Pencu, G.; Hillebrand, M.; Constantinescu, T.; Balaban, A.T. Reaction of 2,2-Diphenyl-1-picrylhydrazyl with HO•, O2•, HO ,and HOO Radicals and Anions. Int. J. Mol. Sci. 2006, 7, 130-143. https://doi.org/10.3390/i7050130

AMA Style

Hristeac EN, Caproiu MT, Pencu G, Hillebrand M, Constantinescu T, Balaban AT. Reaction of 2,2-Diphenyl-1-picrylhydrazyl with HO•, O2•, HO ,and HOO Radicals and Anions. International Journal of Molecular Sciences. 2006; 7(5):130-143. https://doi.org/10.3390/i7050130

Chicago/Turabian Style

Hristeac, Elena N., Miron T. Caproiu, Gabriela Pencu, Mihaela Hillebrand, Titus Constantinescu, and Alexandru T. Balaban. 2006. "Reaction of 2,2-Diphenyl-1-picrylhydrazyl with HO•, O2•, HO ,and HOO Radicals and Anions" International Journal of Molecular Sciences 7, no. 5: 130-143. https://doi.org/10.3390/i7050130

Article Metrics

Back to TopTop