Abstract

The synthesis of 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione derivatives, using one-pot three-component condensation reaction of 3-nitrophthalic anhydride, hydrazine monohydrate, dimedone, and aromatic aldehydes in the presence of ZrO(NO3)2.2H2O as the novel catalyst and in reflux conditions in EtOH was reported. Quantum theoretical calculations for three structures of compounds (5a, 5b, and 5c) were performed using the Hartree–Fock (HF) and density functional theory (DFT). From the optimized structure, geometric parameters were obtained and experimental measurements were compared with the calculated data. The structures of the products were confirmed by IR, 1H NMR, 13C NMR, mass spectra, and elemental analyses. The IR spectra data and 1H NMR and 13C NMR chemical shift computations of the 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione derivatives in the ground state were calculated. Frontier molecular orbitals (FMOs), total density of states (DOS), thermodynamic parameters, and molecular electrostatic potential (MEP) of the title compounds were investigated by theoretical calculations. Molecular properties such as the ionization potential (I), electron affinity (A), chemical hardness (η), electronic chemical potential (µ), and electrophilicity (ω) were investigated for the structures. Thus, there was an excellent agreement between experimental and theoretical results.

1. Introduction

The nitrogen-containing heterocyclic compounds are extensive in nature and play a unique role in biological systems [13]. Heterocycles containing the phthalazine ring are important compounds with broad biological activities, such as anticonvulsant [4], vasorelaxant [5], cardiotonic [6], cytotoxic [7], antimicrobial [8], anticonvulsant [9], antifungal [10], anticancer [11], and anti-inflammatory [12]. Several methods have been reported in the literature for the synthesis of 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione derivatives which have been reported using Mg(HSO4)2 [13], [Bmim]Br [14], p-TSA [15], silica sulfuric acid [16], PPA–SiO2 [17], H2SO4/H2O–EtOH and H2SO4/[bmim]BF4 [18], Ce(SO4)2.4H2O [19], and starch sulfate [20], as catalysts. Given the interest in the synthesis of heterocycles [2124], we described herein a simple synthesis of 1H-indazolo[1, 2-b]phthalazinetriones by three-component condensation reaction of phthalhydrazide, dimedone, and aromatic aldehydes in the presence of a catalytic amount of ZrO(NO3)2.2H2O at reflux conditions in EtOH. In recent years, computational chemistry has become an important tool for chemists and a well-accepted partner for experimental chemistry [25]. Computational chemistry is the application of computer simulation to predict or interpret chemical reactivity. Computational organic chemistry is an important area within which occur the determination of the mechanisms of chemical reactions [26, 27] especially catalysis [28, 29]structural determination of organic compounds [30, 31] prediction of spectroscopic data such as 1H NMR and 13C NMR chemical shifts [32, 33] properties calculation of organic molecules [34, 35] and the interaction of a substrate with an enzyme [36]. This study has revealed some potential leads for possible pharmaceutical applications, and further research may help in the development of new antioxidative agents for important metabolic functions. Also, three new crystal structures of the compounds 5a, 5b, and 5c are reported. In the present work, we investigated the energetic and structural properties of three compounds of 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione derivatives (5a, 5b, and 5c), using the DFT calculations. The optimized geometries, quantum molecular descriptors, IR spectra data, 1H NMR and 13C NMR chemical shift computations, molecular electrostatic potential (MEP), thermodynamic and electronic properties, and NBO analysis were calculated through density functional theory (DFT) and Hartree–Fock (HF) methods.

2. Materials and Methods

The starting materials and solvents were obtained from Merck (Germany) and Fluka (Switzerland) and were used without further purification. The melting points were measured with an Electrothermal 9100 apparatus and were uncorrected. The IR spectra were recorded on a Jasco FT-IR 6300 spectrometer. The 1H NMR and 13C NMR spectra were measured (CDCl3 solution) with a Bruker DRX-250 Avance spectrometer at 250.0 and 62.9 MHz, respectively. Mass spectra were recorded with an Agilent Technologies 5975°C mass spectrometer. The elemental analyses were carried out using a Heraeus CHN-O-rapid analyzer.

2.1. General Procedure for the Synthesis of 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-triones

3-Nitrophthalic anhydride (1, 1 mmol) and hydrazine monohydrate (2, 1 mmol)) were refluxed in ethanol for 15 minutes to form phthalhydrazide as an intermediate. Then, we added dimedone (3, 1 mmol) and aromatic aldehydes (4, 1 mmol) to the mixture of this reaction one by one in the presence of ZrO(NO3)2.2H2O (2 mol%) and the mixture was refluxed again for 2-3 hours. The completion of the reaction was checked by TLC. The solvent was removed under reduced pressure, and the viscous residue was purified by a preparative layer chromatography (silica gel; petroleum ether–ethyl acetate (8 : 2)). The solvent was removed under a reduced pressure and the products 5ac were obtained.

2.2. Computational Studies

In the present study, we carried out quantum theoretical calculations for the compounds 5a, 5b, and 5c using the HF/6-31+G, HF/6–311+G∗∗, B3LYP/6-31+G, and B3LYP/6–311+G∗∗ levels [37] by the Gaussian 03W program package [38] and calculated their properties. During the beginning stage, we obtained an optimized structure using Gaussian 03W program (see Figure 1). Then, we calculated the 1H NMR chemical shifts using the HF/6-31+G, HF/6–311+G∗∗, B3LYP/6-31+G, and B3LYP/6–311+G∗∗ levels for the title compounds (5a, 5b, and 5c) and 13C NMR [39]. The electronic properties included energy of the highest occupied molecular orbital (EHOMO), energy of the lowest unoccupied molecular orbital (ELUMO), HOMO-LUMO energy gap (∆E), ELUMO, natural charges, and molecular properties [40]. The optimized molecular structure, HOMO and LUMO surfaces were visualized using GaussView 03 program [41].

3. Results and Discussion

The three-component reaction between 3-nitro phthalhydrazide (6), dimedone (3), and aromatic aldehydes (4) proceeded very smoothly and cleanly in the presence of a catalytic amount of ZrO(NO3)2.2H2O at reflux conditions in ethanol and afforded the corresponding 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione derivatives (5a-c) in high yields (Scheme 1 and Table 1), and no undesirable side reactions were observed. A mechanistic rationalization for this reaction is provided in Scheme 2. The structures of the products were deduced from their IR, 1H NMR, 13C NMR, mass spectra and elemental analyses. For example, the 1H NMR spectrum of 5a exhibited distinct signals arising from two CH3 groups of dimedone (0.96 and 1.00 ppm, 2 s), two CH2 groups of dimedone ring (2.07–2.33 ppm, m) and (3.65 ppm, s), one OCH3 group of 4-methoxyphenyl ring (3.80 ppm, s), aliphatic CH (5.25 ppm, s), and seven aromatic CH (7.02 ppm, d,3JHH = 8.25 Hz), (7.78 ppm, d,3JHH = 8.25 Hz), and (8.50–8.82, m). The 13C NMR spectrum of 5a shows 22 distinct resonances arising from two CH3 groups (26.91 and 28.16 ppm), C (32.24 ppm), two CH2 groups (42.49 and 50.74 ppm), OCH3 (55.81 ppm), aliphatic CH (64.46 ppm), aromatic and olefinic carbons (113.66, 114.83, 127.03, 127.99, 128.71, 129.08, 129.41, 129.96, 130.40, 131.97, 135.29, and 136.87 ppm), and C=O carbons (160.84, 162.12, and 196.47 ppm). The mass spectrum of 5a displays a molecular ion peak at m/z 447.

3.1. Characterization Data for the Synthesis of 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione Derivatives (5a-c)
3.1.1. 3,3-Dimethyl-13-(4-methoxyphenyl)-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione (5a)

Yellow solid; m.p.>250°C; Yield: 75%; Anal. Calcd for C24H21N3O6 (447.4): C, 64.42; H, 4.73; N, 9.39 %. Found: C, 64.38; H, 4.75; N, 9.33; IR (KBr) (υmax, cm−1): 3439, 2959, 1728 (C=O), 1666 (C=O), 1602 (C=O), 1535 (NO2-Asymmetric Str.), 1464, 1358 (NO2-Symmetric Str.); 1H NMR (DMSO. d6, 250.0 MHz): δH 0.96 (3H, s, CH3), 1.00 (3H, s, CH3), 2.07–2.33 (2H, m, CH2), 3.65 (2H, s, CH2), 3.80 (3H, s, OCH3), 5.25 (1H, s, CH), 7.02 (2H, d, 3JHH = 8.25 HZ, arom CH), 7.78 (2H, d, 3JHH = 8.25 HZ, arom CH), 8.50–8.82 (3H, m, arom CH); 13C NMR (DMSO. d6, 62.9 MHz): δC 26.91 and 28.16 (2CH3), 32.24 (C), 42.49 (CH2), 50.74 (CH2), 55.81 (OCH3), 64.46 (CH), 113.66, 114.83, 127.03, 127.99, 128.71, 129.08, 129.41, 129.96, 130.40, 131.97, 135.29, and 136.87 (aromatic and olefinic carbons), 160.84, 162.12, and 196.47 (3C=O); MS (EI): m/z 447 (M+, 0.02), 286 (5), 269 (100), 253 (9), 241 (27), 161 (76), 134 (27), 91 (24), 77 (33), and 63 (13).

3.1.2. 3,3-Dimethyl-10-nitro-13-phenyl-2,3,4,13-tetrahydro-1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione (5b)

Yellow solid; m.p.>250°C; Yield: 80%; Anal. Calcd for C23H19N3O5 (417.4): C, 66.18; H, 4.59; N, 10.07 %. Found: C, 66.25; H, 4.54; N, 10.00; IR (KBr) (max, cm−1): 2931, 2358, 1724 (C=O), 1617 (C=O), 1588 (C=O), 1539 (NO2-Asymmetric Str.), 1460, 1410 (NO2 - Symmetric Str.); 1H NMR (DMSO. d6, 250.0 MHz): δH 0.96 (3H, s, CH3), 1.24 (3H, s, CH3), 2.07–2.23 (2H, m, CH2), 3.87–3.92 (2H, m, CH2), 5.26 (1H, s, CH), 7.36–8.69 (9H, m, arom CH); 13C NMR (DMSO. d6, 62.9 MHz): δC 28.15 and 28.50 (2CH3), 32.47 (C), 42.49 (CH2), 50.74 (CH2), 64.44 (CH), 126.81, 128.29, 128.44, 128.56, 128.79, 129.14, 129.32, 130.78, 131.76, 131.95, 134.25, and 135.31 (aromatic and olefinic carbons), 161.82, 165.50, and 198.19 (3C=O).

3.1.3. 13-(4-Chlorophenyl)-3,3-dimethyl-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione (5c)

Yellow Solid; m.p. >250°C; Yield: 78%; Anal. Calcd for C23H18ClN3O5 (451.9): C, 61.14; H, 4.02; N, 9.30 %. Found: C, 61.10; H, 4.04; N, 9.24; IR (KBr) (υmax, cm−1): 2928, 2358, 1728 (C=O), 1611 (C=O), 1591 (C=O), 1474 (NO2- Asymmetric Str.), 1393 (NO2-Symmetric Str.), 1085 (C–Cl); 1H NMR (DMSO. d6, 250.0 MHz): δH 0.94 (3H, s, CH3), 1.22 (3H, s, CH3), 2.05–2.20 (2H, m, CH2), 3.87 (2H, s, CH2), 5.23 (1H, s, CH), 7.15–8.66 (8H, m, arom CH); 13C NMR (DMSO. d6, 62.9 MHz): δC 26.82 and 28.12 (2CH3), 32.45(C), 42.43 (CH2), 50.68 (CH2), 64.45 (CH), 114.38, 123.32, 128.22, 128.67, 128.90, 129.07, 129.26, 129.46, 129.89, 130.34, 133.01, and 134.70 (aromatic and olefinic carbons), 161.00, 163.41, and 196.43 (3C=O).

3.2. Computational Section
3.2.1. IR Spectroscopy

Harmonic vibrational frequencies of the title compounds were calculated using the B3LYP/3-21G, HF/3-21G, B3LYP/6-311+G(d), and HF/6-311+G(d) levels. The vibrational frequencies assignments were made using the GaussView program. Some of the characteristic frequencies are provided in Tables 24. The harmonic frequencies calculated by DFT are usually higher than the corresponding experimental values due to the approximate treatment of the electron correlation, anharmonicity effects, and basis set deficiencies [39].

For the title compound (5a), the strong band at 3376 cm−1 in the FT-IR spectrum is assigned as =O mode. The calculated values for this mode are 3375, 3204, 3227, and 3360 cm−1 for HF/6-31+G, HF/6–311+G∗∗, B3LYP/6-31+G, and B3LYP/6–311+G∗∗ levels, respectively. For the title compound (5a), the strong band at 3439, 2959, and 1728 cm−1 in the FT-IR spectrum is assigned as =O mode. The calculated values for this mode are the same as those calculated for HF/3-21G, HF/6-311+G∗∗, B3LYP/3-21G, and B3LYP/6-311+G∗∗ levels, respectively. The DFT computation predicts this vibrational mode in ArC=C at 1464 cm−1 for A, 1489 cm−1 for the compound 5b, and 1493 cm−1 for the compound 5c. This observed frequency coincides well with the expected value [42]. There was an excellent agreement between experimental and theoretical results for all used methods. In order to compare this agreement, the correlation graphic based on the theoretical and experimental data was investigated. A small difference between the experimental and calculated vibrational modes was observed. This difference might have been due to intermolecular hydrogen bonding formation. Also, the experimental results belong to the solid phase while theoretical calculations belong to the isolated gaseous phase.

3.2.2. NMR Parameters

The calculation of 1H NMR and 13C NMR chemical shifts of compounds 5a, 5b, and 5c is done at B3LYP/3-21G, HF/3-21G, B3LYP/6-311+G(d), and HF/6-311+G(d) levels. The experimental and calculated 1H NMR and 13C NMR chemical shifts of 1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione derivatives (5a, 5b and 5c) are presented in Tables 510. Based on our calculations and experimental spectra, we made a reliable one-to-one correspondence between our fundamentals, and all of the chemical shifts were calculated by the HF and B3LYP methods. For the title compound (5a), the aromatic CH protons appeared at δH 7.02–8.82 ppm, and the calculated amounts at HF/6-311+G∗∗ and B3LYP/6-311+G∗∗ basis set levels were at 7.30–8.19 and 7.00–8.89 ppm, respectively. Protons of two methyl groups appeared at δH 0.96 and 1.00 ppm, the calculated amounts at HF/6-311+G∗∗ and B3LYP/6-311+G∗∗ basis set levels were at 0.67 and 1.28 ppm and 0.86 and 1.19 ppm, respectively. Also, chemical shifts of three carbonyl groups appeared at δc 160.84, 162.12, and 196.47 ppm, the calculated amounts at HF/6-311+G∗∗ and B3LYP/6-311+G∗∗ basis set levels were at 149.88, 153.99, and 190.97 ppm and 159.51, 161.27, and 197.32 ppm, respectively. The same was true for other compounds.

There was an excellent agreement between experimental and theoretical results for all methods employed. In order to compare this agreement, the correlation graphic based on the theoretical and experimental data was investigated. The correlation value (R2) for compounds at HF/3-21G, B3LYP/3-21G, HF/6-311+G∗∗, and B3LYP/6-311+G∗∗ is presented in Table 11. There was an excellent agreement between experimental and theoretical results [39]. A small difference between the experimental and calculated vibrational modes was observed. This difference might have been due to the intermolecular hydrogen bonding formation. Furthermore, the experimental results belong to the solid phase while and theoretical calculations belong to the isolated gaseous phase.

3.2.3. Electronic Properties

Quantum chemical methods are important for obtaining information about molecular structure and electrochemical behavior. A frontier molecular orbital (FMO) analysis was performed for the compounds using the B3LYP/6-311+G(d) level [38]. FMO results such as EHOMO, ELUMO, and the HOMO-LUMO energy gap (∆E) of the title compounds are summarized in Table 7. The energy of the LUMO, HOMO, and their energy gaps reflected the chemical reactivity of the molecule [43]. In addition, the HOMO could act as an electron donor and the LUMO as an electron acceptor. A higher HOMO energy (EHOMO) for the molecule indicated a higher electron-donating ability to an appropriate acceptor molecule with a low-energy empty molecular orbital [44]. As shown in Figure 1 the HOMO energy of the compound 5c had the highest value (−0.29 eV). A large energy gap implied high stability for the molecule. The calculated values of the HOMO-LUMO energy gap (∆E) for the structures 5a, 5b, and 5c were 0.07, 0.08, and 0.06 eV, respectively. DOS plots [45] also demonstrated the calculated energy gaps (∆E) for the compounds 5a, 5b, and 5c (see Figure 2). It is obvious that the energy gap of the compound 5b was the highest (0.08 eV); therefore, it was less reactive than the other structures, whereas the energy gap of the compound 5c was the lowest (0.06 eV), which indicates that it was the most reactive. As presented in Figure 2, charge transfer could take place within the three molecules.

The electronic properties such as ionization potential, electron affinity, global hardness, electronic chemical potential, and electrophilicity are calculated in Table 12. The first ionization potential (I) and electron affinity (A) could be expressed through HOMO and LUMO orbital energies by connecting it with Hartree–Fock SCF theory and invoking Koopmans' theorem [46] as I = −EHOMO and A = −ELUMO. The chemical hardness (η = I−A/2) is an important property that measures the molecular stability and reactivity [47]. A hard molecule has a large energy gap (∆E) and a soft molecule has a small energy gap (∆E) [48]. The chemical hardness (η) values of the compounds 5a, 5b, and 5c were 0.170, 0.176, and 0.175 eV, respectively. Compound 5b had the highest chemical hardness (η = 0.176 eV); therefore, it was a hard molecule with less reactivity and a high energy gap (∆E = 0.08 eV). The electronic chemical potential (µ = −(I + A)/2) is a form of the potential energy that can be absorbed or released during a chemical reaction and that may also change during a phase transition [49]. The electronic chemical potential of 5c had the most negative value (−0.26 eV). The electrophilicity (ω) measures the stabilization in energy when the system acquires an additional electronic charge from the environment. The electrophilicity index (ω = µ2/2η) contains information about both electron transfer (chemical potential) and stability (hardness) and is a better descriptor of global chemical reactivity [50]. The higher value of electrophilicity index shows the high capacity of the molecule to accept electrons. The electrophilicity index for the compounds 5a, 5b, and 5c was 0.0040, 0.0046, and 0.0059 eV, respectively. The compound 5c had the highest electrophilicity index; therefore, it had a high capacity for accepting electrons. The dipole moment (µD) is a good measure of the asymmetric nature of a molecule. The size of the dipole moment depends on the composition and dimensionality of the 3D structures. As shown in Table 12, all structures had a high value of dipole moment and point group of C1, which reflected no symmetry in the structures. The dipole moment for the compound 5b (B3LYP/6-311+G(d) = 6.172 Debye) was higher than that for the compounds 5a and 5c (4.988 and 4.396 Debye, respectively).The high value for 5c was due to its asymmetric character.

3.2.4. Thermodynamic Analysis

The total energy of a molecule consists of the sum of translational, rotational, vibrational, and electronic energies. The statistical thermochemical analysis of title compounds is carried out considering the molecule to be at room temperature of 25°C and 1 atmospheric pressure. The thermodynamic parameters such as zero-point vibrational energy, rotational constant, heat capacity (C), and the entropy (S) of the title compound by B3LYP/6-311+G(d) level are listed in Table 13. According to Table 13, the calculated values for compound 5a and 5b are larger than compound 5c; therefore, compounds 5a and 5b have maximum stability compared to compound 5c due to formation of intramolecular hydrogen bonding.

3.2.5. Molecular Electrostatic Potential (MEP)

The molecular electrostatic potential (MEP) was calculated by the B3LYP/6-311+G (d) level. The MEP is related to the electronic density and is a very useful descriptor in understanding sites for electrophilic attack and nucleophilic reactions as well as hydrogen bonding interactions [45]. The negative regions (red color) of the MEP are related to electrophilic reactivity, and the positive (blue color) region is related to nucleophilic reactivity, as shown in Figure 2.

Molecular electrostatic potential (MEP) surface aims at locating the positive and negative charged electrostatic potential in the molecule. In each MEP surface, there is a color scale which indicates the negative and positive value. The red color is a sign for the negative extreme, and the blue color represents the positive extreme. The red color with a negative sign indicates the minimum electrostatic potential (which means it is bound loosely or has excess electrons), and it acts as electrophilic attack. The blue color also indicates the maximum of electrostatic potential, and it acts in the opposite manner.

Starting from the above note, if we plot all MEP surfaces with all isosurface values, we see only the top surface. It is observed from the MEP map in Figure 2 that the nitrogen-bonded oxygen atoms in the NO2 group and the oxygen atoms in the carbonyl groups (C=O) of the rings are negative regions in all compounds because in the resonance form of the nitro group, the oxygen atoms have a negative charge and the nitrogen atom has a positive charge, and in the resonance form of the carbonyl groups, the oxygen atoms have a negative charge and the carbon atom has a positive charge. Thus, oxygen atoms are sites for electrophilic activity. The nitrogen atoms in the ring, which are attached to carbonyl lethal electron groups, are also positively charged; therefore, nitrogen atoms are sites for nucleophilic attraction. As such, these sites provide information about the regions where the compounds can have strong intermolecular interactions.

4. Conclusion

In the present study, the three-component reaction between 3-nitrophthalic anhydride, hydrazine monohydrate, dimedone, and aromatic aldehydes in the presence of a novel catalytic amount of ZrO(NO3)2.2H2O to produce 1H-indazolo[1,2-b]phthalazine-1,6,11(13H)-trione derivatives was reported. The reported method offers a mild and efficient procedure for the preparation of these compounds. These compounds of the products were confirmed by IR, 1H NMR, 13C NMR, mass spectra, and elemental analyses. The IR spectra data and 1H NMR and 13C NMR chemical shift computations of the 1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione derivatives in the ground state were calculated. There was an excellent agreement between experimental and theoretical results. Frontier molecular orbitals (FMOs), total density of states (DOS), and molecular electrostatic potential (MEP) of the title compounds were investigated through theoretical calculations.

Data Availability

No data were used to support this study.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Acknowledgments

The authors are thankful to the Zanjan Branch, Islamic Azad University, for partial support of this work.

Supplementary Materials

Figure S1: 1H NMR (62.9 MHz, DMSO) spectrum of 3,3-dimethyl-13-(4-methoxyphenyl)-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione (5a). Figure S2: 13C NMR (62.9 MHz, DMSO) spectrum of 3,3-dimethyl-13-(4-methoxyphenyl)-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione (5a). Figure S3: mass spectrum of 3,3-dimethyl-13-(4-methoxyphenyl)-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione (5a). Figure S4: 1H NMR (62.9 MHz, DMSO) spectrum of 3,3-dimethyl-10-nitro-13-phenyl-2,3,4,13-tetrahydro-1H-indazolo[1,2-b] phthalazine 1,6,11(13H)-trione (5b). Figure S5: 13C NMR (62.9 MHz, DMSO) spectrum of 3,3-dimethyl-10-nitro-13-phenyl-2,3,4,13-tetrahydro-1H-indazolo[1,2-b] phthalazine-1,6,11(13H)-trione (5b). Figure S6: 1H NMR (62.9 MHz, DMSO) spectrum of 13-(4-chlorophenyl)-3,3-dimethyl-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b]-phthalazine-1,6,11(13H)-trione (5c). Figure S7: 13C NMR (62.9 MHz, DMSO) spectrum of 13-(4-chlorophenyl)-3,3-dimethyl-10-nitro-2,3,4,13-tetrahydro-1H-indazolo[1,2-b]-phthalazine-1,6,11(13H)-trione (5c) (Supplementary Materials)