Previous Article in Journal
Numerical Simulation of the Isoparaffins Dehydrogenation Process in Fluidized Bed Reactor: From Laboratory to Industry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Design of Acridone-Core Energetic Materials: Assessment of Detonation Properties and Potential as Insensitive, Thermally Stable High-Energy Materials

by
Jelena Tamuliene
1,* and
Jonas Sarlauskas
2,*
1
Physics Faculty, Institute of Theoretical Physics and Astronomy, Vilnius University, Sauletekio av. 3, LT-10257 Vilnius, Lithuania
2
Life Sciences Center, Department of Xenobiotics Biochemistry, Institute of Biochemistry, Vilnius University, Sauletekio av. 7, LT-10257 Vilnius, Lithuania
*
Authors to whom correspondence should be addressed.
ChemEngineering 2025, 9(6), 130; https://doi.org/10.3390/chemengineering9060130
Submission received: 6 October 2025 / Revised: 4 November 2025 / Accepted: 11 November 2025 / Published: 13 November 2025

Abstract

In this study, we investigated the impact of incorporating energetic substituents such as –NO2, –NH2, –Cl, –F, N-methyl-N-nitro (CH3–N–NO2), and picryl on the detonation performance and stability of acridone-based compounds. The B3LYP/cc-pVTZ approach was applied to investigate the influence of substitutions on the stability and detonation properties of acridone derivatives. The results obtained exhibit the significant influence of both the type and position of substituents on the energetic performance and stability of the compounds studied. Notably, the acridone derivative bearing a picryl group and four –NH2 substituents exhibited energetic properties superior to those of 2,4,6-trinitrotoluene (TNT). Its calculated velocity lies in the range [7.45–7.66] km/s, and its detonation pressure is [22.49–24.36] GPa; however, its stability is lower than that of core compounds. This reduction, however, is dependent on both the nature and number of substituents introduced.

1. Introduction

Acridone, a heterocyclic aromatic scaffold featuring a carbonyl group and a secondary amine, possesses a unique, planar heterocyclic ring system containing nitrogen, which can exist in both protonated and unprotonated forms [1,2]. Owing to these structural characteristics, the acridone core has attracted considerable interest from medicinal chemists worldwide. In the past years, increased knowledge and understanding of the mode of action of acridone derivatives have driven sustained and dynamic research into their potential, mostly as anti-cancer agents. However, the use of acridone in co-crystallization technology for the design of high-energy materials has drawn growing attention as a promising strategy to balance explosive performance with safety considerations, including sensitivity to heat, impact, shock, and friction. Recently, Sen et al. demonstrated that co-crystallization of trinitrotoluene (picryl) and picric acid (PIC) with acridine (ACR) enhanced the properties of these high-energy materials, suggesting their potential as insensitive explosives [3]. They also state that the newly formed compounds exhibit excellent thermal stability and propose strategies to enhance it: (i) salt formation and (ii) crystallization. Another important finding from the study is the critical role of conformers in determining the performance of the target products.
The advanced compound 3-nitro-1,2,4-triazol-5-one (NTO), when co-crystallized with acridine (ACR), demonstrated good thermal stability and excellent impact insensitivity. Moreover, the obtained detonation velocity and pressure values confirm its potential as a high-performance explosive [4]. Despite the potential of co-crystals to unite high heats of formation, ambitious detonation performance, strong resistance to thermal degradation, and reduced susceptibility to impact and shock, few studies have been conducted in this direction. Therefore, we designed and conducted theoretical studies to evaluate the effect of substitutions on the stability and detonation performance of acridone-based polynitro derivatives to predict their potential as safe and powerful high-energy materials. In our investigation, the energetic and stability properties of Tetranitroacridone, or 2,4,5,7-Tetranitro-10H-acridin-9-one according to universally-recognized authority on chemical nomenclature and terminology (UIPAC) nomenclature (Acri1), were evaluated upon substitution with selected functional groups such as NO2, –NH2, –Cl, –F, N-methyl-N-nitro (CH3–N–NO2), and Picryl. The novelty of this work lies in the selection of acridone as a core molecule for the development of new high-energy materials and, consequently, new co-crystals. To the best of our knowledge, we are among the first to investigate acridones as high-energy materials, i.e., materials that can store and release significant amounts of energy. The selection of functional groups was considered based on their expected impact on detonation performance. For example, it has been widely reported that the introduction of –NH2 leads to enhanced thermal stability, while –NO2 improves the energetic characteristics of high-energy compounds [5,6,7,8,9,10,11,12,13,14,15]. The effects of fluorine and chlorine substitutions were also established, as these halogens are known to improve detonation performance through bond reinforcement, altered decomposition pathways, greater energy output, and enhanced structural stability, often accompanied by reduced sensitivity to unintended initiation [16,17]. Methylation of the compound reduces sensitivity to mechanical stimuli and thermal decomposition, making energetic materials safer and more manageable [18,19]. On the other hand, nitrogen-rich compounds typically exhibit high sensitivity to external stimuli, but they also demonstrate excellent detonation performance [20]. Additionally, 1,3,5-trinitrobenzene derivatives, such as picric acid, are used in the formulation of energetic materials. Owing to their high thermal stability and elevated melting points, compounds containing trinitrobenzene groups are particularly suitable for applications in explosives and related energetic systems, but they have not been investigated as thoroughly [21].
We hope that our study will help reveal the most promising strategy for synthesizing co-crystals to create advanced, safe, and cost-effective materials that meet hazardous material requirements.

2. Methods

Computational investigations employed the B3LYP functional in combination with the cc-pVTZ basis set, as implemented in GAUSSIAN [22,23,24]. This approach yielded results consistent with experimental data [25,26,27,28,29,30,31,32,33]. The Berny optimization was applied to find the most stable conformations. The procedure was followed by vibrational frequency analysis to confirm true minima. The total energy was an indicator used to identify the most stable conformers. The selected conformers were then further analyzed.
Stability and sensitivity were assessed through key descriptors: cohesive energy per atom, HOMO–LUMO gap, chemical hardness, electronegativity, hardness index, and oxygen balance [34]. Cohesive energy (BDE) indicated thermal stability (the ability to maintain physical and chemical properties at high temperature), while the HOMO–LUMO gap and hardness reflected electronic stability. Electronegativity was used to gauge bond-forming propensity and predict the aging compounds.
Three computational methods were employed to predict the densities of the compounds:
  • Politzer’s equations [35], which incorporate the 0.001 e/bohr3 isosurface and surface potential balance;
  • Calculations from molecular weight by B3LYP/cc-pVTZ and molar volume;
  • ACD/ChemSketch [36], where density is estimated using a machine learning algorithm trained on compounds with known densities, expressed as a function of molecular volume, electronic properties, and intermolecular interaction patterns. The obtained densities were compared.
The energetic performance of the materials was assessed through detonation pressure and velocity, calculated using several approaches developed for Cl/F-containing compounds [37,38,39]. Predictions obtained with the method of [39] were notably higher than those from the equations in [37,38]. The latter yielded values for CaHbOcNd in close agreement with the established Kamlet–Jacobs approach. Therefore, in this study, we present detonation parameters evaluated using the equations described in [37,38]. Sensitivity to shock was estimated from the oxygen balance, calculated according to formulae presented in [40,41]. It is used for compounds that contain F or Cl.
Aiming to combine two valuable practical properties of modern high-energy materials, the ortho-amino-nitro groups present in the acridone moiety (responsible for the decreased sensitivity to mechanical stimuli) were additionally modified at N-9 with specific Picryl or 3,5-diamino-picryl fragments, known for markedly increasing thermal stability [42]. Similar methods in molecular design were previously successfully applied for the design of famous thermally resistant energetic molecules such as 3,5-Dinitro-N,N’-bis(2,4,6-trinitrophenyl)-2,6-pyridinediamine (PYX) (therm. decomp. 360 °C) and PL-1 (2,4,6-Tris(3’,5’-Diamino-2’,4’,6’-trinitrophenylamino)-1,3,5-triazine, therm. decomp. 337 °C) [43,44,45].

3. Results and Discussion

3.1. Stability

3.1.1. Thermal Stability

Figure 1 depicts the molecular skeleton together with the electrostatic potential, which was used to investigate the influence of substitutions on stability and detonation properties.
This compound contains –NO2 groups, making it a potential high-energy material. Moreover, there are several areas, marked as I and II in Figure 1, for the substitutions. Based on the electron distribution in the compounds, we predict that the more favorable position for substitution is site I, which is slightly more positive than site II and is surrounded by less negative regions (Figure 1). So, the substitutions firstly filled site I, and later site II. From a chemical perspective, the preferential occupation of site I may also be related to reduced steric hindrance and increased stabilization through electronic effects. Additionally, Picryl or other energetic fragments can be introduced in place of the hydrogen atom of the N–H group [3,4]. This introduction allows us to examine how stability and detonation properties change when not only the molecular skeleton but also the introduced molecule is substituted. It is worth emphasizing that, in some cases, the introduction of substitutions leads to major alterations in the skeleton molecule. The core (benzene rings) bends and becomes non-planar. To illustrate these structural changes, we provide the coordinates of each studied compound in the Supplementary Materials, enabling more detailed analysis, while in Appendix A, there are abbreviations, chemical compositions, and skeletons of each compound. The importance of molecular structure for stability is well known and is also confirmed by our results. For example, the difference between conformers Acri2 and Acri2_II lies in the position of the hydrogen atoms in the OH substituent. While the thermal stability of these compounds is the same, their chemical stability differs (Table 1). Notably, the hardness index of the investigated compounds is higher than or close to 0.8 (Table 1), which indicates their overall high stability, although variations reveal which substitutions may reduce or increase this stability.
Our results for cohesive energy (BDE) show that only F substitutions leave the thermal stability of Acri1 unchanged. In most cases, the thermal stability of the compounds is slightly lower, although it decreases significantly in the case of Acri15 (Table 1). Moreover, compounds with fewer substitutions are more thermally stable, whereas more complex substituents tend to reduce this stability to a greater extent than simpler ones (Table 1, Figure 2). For instance, the compound with four NCH3NO2 substituents shows the lowest thermal stability of all compounds investigated. These results are in good agreement with previous research [46,47].
Notably, the thermal stability of the Picryl-containing compound (Acri12) decreases significantly when the Picryl group is substituted by –NH2 (Acri15) but remains approximately the same when the core of the compound is substituted by –NH2 (Acri13 and Acri14). The results coincide well with the conclusions drawn from the experimental research presented in [48,49,50,51].
In summary, we conclude that Acri1 is thermally stable, and substitutions such as –Cl, –F, –NH2, –NO2, or -Picryl do not significantly affect this property. However, thermal stability may decrease when the substituent contains –NO2 groups (e.g., N–NO2–CH3 or Picryl with additional amino groups), most likely because strong electron-withdrawing effects weaken the substituent’s structure and increase its susceptibility to thermal degradation. The analysis further revealed that the thermal stability of Acri1 is not affected by the position of the substituents.

3.1.2. Chemical Stability

HOMO–LUMO gap and chemical hardness analyses reveal the same trend in the influence of substitutions on the chemical stability of the studied compounds (Table 1, Figure 3). While the HOMO–LUMO gap reflects stability and chemical hardness indicates reactivity, both parameters describe the reactivity potential of the compounds. Hence, we focus our discussion on the changes in the HOMO–LUMO gap due to substitutions.
Generally, Acri1 becomes less chemically stable upon substitution, except when four –OH or –F groups are present (Figure 3). An increasing number of hydrogen-bond-forming substitutions enhances chemical stability, provided that other effects, such as steric hindrance, do not prevail. This trend is clearly reflected in the HOMO–LUMO gap and chemical hardness of the compounds containing Picryl: substitution of the core molecule with two –NH2 groups slightly increases stability, whereas further increasing the number of such substitutions leads to decreased chemical stability. The importance of both the presence and the placement of hydrogen bonds is also illustrated by the results obtained for Acri2, Acri2_II, and Acri7. Acri2 and Acri2_II differ only in the position of the OH hydrogen: in Acri2 the hydrogen bonds form a V-shaped arrangement, while in Acri2_II they adopt a cyclic motif. The Acri7, where four hydrogen bonds form cycle arrangements, is chemically more stable than Acri2_II and Acri1. Since larger substituents cause greater steric hindrance, which can reduce chemical stability, an increasing number of Cl substituents leads to lower stability. Additionally, relatively weaker bonding compared to fluorine may also be the reason for the higher chemical stability of compounds with F substitutions, which tends to increase chemical stability through strong, stable bonds and favorable electronic effects [52]. A compound combining high chemical stability with highly electronegative substituents typically exhibits strong covalent bonding, low susceptibility to spontaneous degradation, and resistance to aging, while maintaining the potential for selective reactivity under extreme conditions.
The electronegativity of compounds such as Acri2, Acri4, Acri6, Acri11, Acri9, and Acri12 is higher than that of Acri1 (Figure 4), reflecting a stronger tendency to attract electrons.
Considering the low chemical and thermal stability, as well as the high electronegativity, of Acri2, Acri6, Acri11, and Acri12, we propose that these compounds are likely to degrade more rapidly than Acri1. In contrast, the selective reactivity and degradation of Acri4 and Acri9 are expected to depend on the environmental and chemical context. The analysis of the charge distribution illustrates these observations: in the compounds Acri2, Acri6, Acri11, and Acri12 the electron-deficient areas are located on the outer part of the compounds, whereas in Acri4 and Acri9, these regions are concentrated in the inner molecular structure (see Supplementary Materials).
The degradation of the remaining compounds could also be faster than that of Acri1 due to their higher instability despite their electronegativity. Thus, only –F substitutions enhance Acri1 stability and reduce its degradation rate. Notably, the position of the substituents plays a significant role in chemical stability. The compound bearing a picryl group and the one substituted with –NO2 group are markedly more reactive compared to the others.

3.2. Sensitivity

We obtained oxygen balance (OB) as a preliminary assessment to predict the sensitivity of the compounds under study (Figure 5). A highly negative oxygen balance is often associated with reduced sensitivity, which may be attributed to the generation of unbalanced reactive species during thermal decomposition [53,54,55,56]. This statement is supported by our results, which show that an increase in reactive fragments leads to increased sensitivity. For instance, compounds bearing four chlorine or fluorine substituents exhibit higher sensitivity compared to those with only two such groups.
It is important to note that OB of TNT is −73.97%, while that of 1,3,5-Trinitro-1,3,5-triazinane (RDX) is −21.61%. The comparison of these values with those of the studied compounds shows that all compounds investigated are substantially less sensitive than RDX, but only Acri1, Acri5, and Acri10 exhibit lower sensitivity than TNT. This implies that the compounds are more resistant to external stimuli such as impact or friction than RDX, although only a few outperform TNT in terms of insensitivity. Given that moderately sensitive materials generally have an OB between −30% and −20%, it can be stated that all of the compounds studied fall within the insensitive category.

3.3. Detonation Properties

3.3.1. Density

Material density plays a key role in determining the detonation products, which directly influence detonation pressure and velocity [57]. It was mentioned in the Section 2 that density was calculated using three different approaches. Among them, Politzer’s method, combined with rigorously validated B3LYP quantum chemical calculations, captures intermolecular interactions that are critical for accurate density modeling [58]. The results are summarized in Table 2. As expected, different approaches yield different values. In the majority of cases, these differences do not exceed 5–6%, which is considered an acceptable deviation. However, in some instances, the densities calculated using the ACD/ChemSketch approach exceed those from other methods by more than 9%. Considering that experimentally measured densities can vary by more than 10% under different conditions, we speculate that the computational approaches applied here provide reasonably trustworthy results. For confirmation, the density of TNT is generally stable but can vary from approximately 1.57 to 1.67 g/cm3 depending on temperature, pressure, and physical state. Similarly, the experimentally determined density of CF3N3 has been reported in the range of 1.716–1.816 g/cm3, while the value estimated using ACD/ChemSketch is 1.77 g/cm3 [59,60]. The calculated densities must fall within the range of 1.62 to 1.94 g/cm3, which is consistent with the reported values for substituted acridone derivatives [61].
Analysis of the calculated densities indicates that the density of Acri1 increases upon substitution, most likely due to a greater rise in overall molecular weight compared to volume. Indeed, most compounds exhibit larger calculated volumes compared to Acri1, and only the volumes of both conformers of Acri2 and Acri5 are found to be lower. In these cases, the increase in density appears to result from tighter molecular packing, promoted by molecular shape and intermolecular interactions such as hydrogen bonding or dipole–dipole forces, which reduce molecular volume and thereby enhance density [62]. Notably, contradictory results were obtained in the case of Acri10. The density calculated using molecular weight and B3LYP/cc-pVTZ-derived molar volume, as well as the value estimated via Politzer’s equation, is lower than that of Acri1. However, the density obtained using the ACD/ChemSketch approach is higher. These findings indicate that the ACD/ChemSketch method may more accurately account for empirical influences—such as molecular packing, steric effects of substituents, and crystal lattice interactions—that affect density in the solid state. On the other hand, the observed discrepancy may also reflect the variability in the density of the same compound when crystallized under different conditions.
No clear trend was observed between density and the type or number of substitutions. However, the results confirm that density increases when the rise in molecular weight outweighs the increase in molecular volume, particularly when the substituents are flexible enough to adapt their shape and promote intermolecular interactions.

3.3.2. Detonation Velocity and Pressure

High-energy materials generally exhibit detonation velocities in the range of 1.01 to 9.89 km/s. TNT, a widely used reference explosive, has a typical detonation velocity of ~6.9 km/s. In contrast, RDX and 1,3,5,7-Tetranitro-1,3,5,7-tetrazocane (HMX) achieve higher detonation velocities of 8.7 km/s and 9.1 km/s, with associated detonation pressures of approximately 33.8 GPa and 39.3 GPa, respectively, compared to ~21.0 GPa for TNT [63,64]. The detonation pressure and velocity of compounds consisting of one or two benzene rings annelated with N-heterocycle fall within the range of 20–25 GPa and 6.6–8.0 km/s, respectively [63,65].
The compounds under investigation generally exhibit lower detonation pressure compared to TNT (Table 3). An exception is Acri14, whose detonation pressure exceeded that of TNT and was equivalent to that of the aforementioned nitrogen-rich compounds, but considerably lower than that of RDX and HMX. Thus, based on the analysis of the detonation pressure, a parameter reflecting shockwave intensity, the compounds under study are unlikely to be effective explosives. While the detonation velocities of the compounds fall within the range typical of high-energy materials, only Acri14 and Acri15 exhibit values significantly exceeding those of TNT (Table 4). These compounds, along with -NH2 substitutions, contain picryl groups, and such substituents are often associated with detonation velocities and pressures similar to, or intermediate between, those of TNT and RDX. Furthermore, the significant exothermic decomposition observed in compounds with picryl substituents suggests a high energy output. So, the high detonation pressure and velocity of Acri14 and Acri15 could be related to the decomposition of picryl. This implies that when these compounds are initiated, chemical reactions will begin when the bonds between the core molecules and picryl and its substitutions begin to break. These first reactions generate highly reactive intermediate species that promote further decomposition of the core molecule and the series of exothermic reactions that cause thermal runaway. The comparison of the detonation pressure and velocity of the rest compound supports this finding. First, the detonation pressures and velocities of Acri12 and Acri13 are lower than those of Acri14, despite all three containing picryl groups. The primary difference between these compounds lies in the number of –NH2 substituents. The absence or an insufficient number of these substituents may result in an unbalanced formation of highly reactive intermediate species, which is inadequate to promote the effective decomposition of the molecular core. Second, the analysis of detonation pressures and velocities indicates lower values for compounds with the strongest bonds between the core molecule and its substituents. For example, Acri3 and Acri4—both containing strong C–Cl or C–F bonds—exhibit some of the lowest detonation pressures and velocities among the compounds studied.
From the results obtained, it follows that the detonation performance of acridone-based compounds is potentially improved by incorporating picryl-like groups along with appropriate substitutions on the core structure.

4. Conclusions

Studies were undertaken to assess the effects of substitutions on the stability and detonation performance of acridone-based polynitro derivatives and to assess their potential as safe, high-energy materials. The results show that Acri1 is thermally stable, and substitutions with –Cl, –F, –NH2, –NO2, or –Picryl do not significantly affect this property, although compounds with substituents containing –NO2 groups are less thermally stable. We also found that the position of electron-deficient areas influences chemical stability. Based on this, Acri2, Acri6, Acri11, and Acri12 are predicted to degrade more rapidly than Acri1, while only –F substitution improves Acri1 stability and reduces its degradation rate. Notably, compounds bearing a picryl group or substituted with –NH2 are markedly more reactive than the others. It is also worth mentioning that picryl itself does not affect the thermal stability of the compounds under study, although its modification (substitution) leads to a marked decrease in stability.
Based on the OB analysis, all of the studied compounds can be classified as insensitive. The detonation performance of acridone-based compounds is potentially enhanced by incorporating picryl-like groups together with appropriate substitutions on the core structure. However, while substitutions on the core molecule do not significantly affect the stability of acridone-based compounds, modifications to the picryl-like substituents lead to a marked decrease in stability.
The planar structure of Acri1 and its proposed ability to form hydrogen bonds suggest that it could be used for co-crystal production. Moreover, the properties of acridone-based co-crystals may be further improved through substitutions on this compound or on the substituted fragment.
In summary, we may state that a preliminary molecular design possessing heterocyclic acridone skeletons is proposed. These types of compounds haven‘t been explored until today as new energetic materials integrating a fused benzene and pyridine ring. Our selected preliminary compact set (15) of acridone polynitroderivatives can be further studied and expanded for their practical relevance to the creation of heat-resistant charges for conducting drilling and blasting operations in deep, high-temperature and geothermal wells, sampling rocks, and eliminating accidents during drilling of oil and gas wells, and also in the destruction of hot slag heaps of metallurgical plants. These new high-energy materials can also potentially be used as high-energy sources in military equipment where insensitive and thermostable charges are required.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/chemengineering9060130/s1.

Author Contributions

Conceptualization, J.T. and J.S.; methodology, J.T. and J.S.; formal analysis, J.T. and J.S.; investigation, J.T. and J.S.; writing—original draft preparation, J.T. and J.S.; writing—review and editing, J.T. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The numerical calculations with the GAUSSIAN09 package were performed using the resources of the Information Technology Research Center of Vilnius University and the supercomputer “VU HPC” of Vilnius University in the Faculty of Physics location.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
BDECohesive energy per atom
HOMO-LUMO gapDifference between the highest occupied and the lowest unoccupied orbitals
OBOxygen Balance (calculated to CO2)
PYX3,5-Dinitro-N,N’-bis(2,4,6-trinitrophenyl)-2,6-pyridinediamine
Acri1C13H5N5O92,4,5,7-Tetranitroacridin-9(10 H)-one
Acri2C13H5N5O113,6-Dihydroxy-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri3C13H3Cl2N5O73,6-Dichloro-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri4C13H3F2N5O73,6-Difluoro-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri5C13H7N7O93,6-Diamino-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri6C15H9N9O133,6-Bis[methyl(nitro)amino]-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri7C13H5N5O131,3,6,8-Tetrahydroxy-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri8C13HCl4N5O91,3,6,8-Tetrachloro-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri9C13HF4N5O91,3,6,8-Tetrafluoro-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri10C13H9N9O91,3,6,8-Tetraamino-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri11C17H13N13O171,3,6,8-Tetrakis[methyl(nitro)amino]-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri12C19H6N8O1510-Picryl-2,4,5,7-tetranitroacridin-9(10 H)-one
Acri13C19H8N10O1510-Picryl-3,6-Diamino-2,4,5,7-tetranitroacridin -9(10 H)-one
Acri14C19H10N12O1510-Picryl-1,3,6,8-tetraamino-2,4,5,7-tetranitroacridin -9(10 H)-one
Acri15C19H12N14O1510-(3’,5’-Diamino-picryl)-1,3,6,8-tetraamino-2,4,5,7-tetranitroacridin -9(10 H)-one

Appendix A

Structural formulas of designed acridone-based energetic materials, their molecular weights, and calculated elemental composition data are presented in Table A1.
Table A1. Structural formulas of designed acridone-based energetic materials, their molecular weights (MW), and calculated elemental composition data.
Table A1. Structural formulas of designed acridone-based energetic materials, their molecular weights (MW), and calculated elemental composition data.
AssignationStructureMolecular FormulaMWCalculated Elemental Composition Data
C, %H, %F or Cl, %N, %O, %
Acri1Chemengineering 09 00130 i001C13H5N5O9375.2141.621.34018.6738.36
Acri2Chemengineering 09 00130 i002C13H5N5O11407.2138.341.24017.2043.22
Acri3Chemengineering 09 00130 i003C13H3Cl2N5O9444.1035.160.6815.97 (Cl)15.7732.42
Acri4Chemengineering 09 00130 i004C13H3F2N5O9411.1937.970.749.24
(F)
17.0335.02
Acri5Chemengineering 09 00130 i005C13H7N7O9405.2438.531.74024.1935.53
Acri6Chemengineering 09 00130 i006C15H9N9O13523.2934.431.73024.0939.75
Acri7Chemengineering 09 00130 i007C13H5N5O11407.2138.341.24017.2043.22
Acri8Chemengineering 09 00130 i008C13HCl4N5O9512.9930.440.2027.64
(Cl)
13.6528.07
Acri9Chemengineering 09 00130 i009C13HF4N5O9447.1734.920.2316.99
(F)
15.6632.20
Acri10Chemengineering 09 00130 i010C13H9N9O9435.2735.872.08028.9633.08
Acri11Chemengineering 09 00130 i011C17H13N13O17671.3730.411.95027.1240.51
Acri12Chemengineering 09 00130 i012C19H6N8O15586.3038.921.03019.1140.93
Acri13Chemengineering 09 00130 i013C19H8N10O15616.3337.031.31022.7338.94
Acri14Chemengineering 09 00130 i014C19H10N12O15646.3635.311.56026.0037.13
Acri15Chemengineering 09 00130 i015C19H12N14O15676.3933.741.79028.9935.48

References

  1. Minotti, G.; Menna, P.; Salvatorelli, E.; Cairo, G.; Gianni, L. Anthracyclines: Molecular advances and pharmacologic developments in antitumor activity and cardiotoxicity. Pharmacol. Rev. 2004, 56, 185–229. [Google Scholar] [CrossRef]
  2. Baguley, B.C.; Wakelin, L.P.G.; Jacintho, J.D.; Kovacic, P. Mechanisms of action of DNA intercalating acridine-based drugs: How important are contributions from electron transfer and oxidative stress? Curr. Med. Chem. 2003, 10, 2643–2649. [Google Scholar] [CrossRef]
  3. Şen, N.; Pons, J.-F.; Kurtay, G.; Yüksel, B.; Nazir, H.; Khumsri, A.; Atakol, O. New explosive materials through covalent and non-covalent synthesis: Investigation of explosive properties of trinitrotoluene: Acridine and picric acid: Acridine. J. Mol. Struct. 2023, 1292, 136080. [Google Scholar] [CrossRef]
  4. Şen, N.; Pons, J.-F.; Zorlu, Y.; Dossi, E.; Persico, F.; Temple, T.; Aslan, N.; Khumsri, A. Synthesis, structure characterization, Hirshfeld surface analysis, and computational studies of 3-nitro-1,2,4-triazol-5-one (NTO): Acridine. Struct. Chem. 2024, 35, 1865–1879. [Google Scholar] [CrossRef]
  5. Drake, G.W.; Hawkins, T.; Brand, A.; Hall, L.; Mckay, M.; Vij, A.; Ismail, I. Energetic, low-melting salts of simple heterocycles. Propellants Explos. Pyrotech. 2003, 28, 174–180. [Google Scholar] [CrossRef]
  6. Darwich, C.; Klapötke, T.M.; Sabaté, C.M. 1,2,4-Triazolium-cation-based energetic salts. Chem. Eur. J. 2008, 14, 5756–5771. [Google Scholar] [CrossRef]
  7. Tao, G.H.; Tang, M.; He, L.; Ji, S.P.; Nie, F.D.; Huang, M. Synthesis, structure and property of 5-aminotetrazolate room-temperature ionic liquids. Eur. J. Inorg. Chem. 2012, 2012, 3070–3078. [Google Scholar] [CrossRef]
  8. Liu, Q.; Yuan, M.; He, J.; Yu, P.; Guo, X.; Liu, Y.; Gao, H.; Yin, P. Exchanging of NH2/NHNH2/NHOH groups: An effective strategy for balancing the energy and safety of fused-ring energetic materials. Chem. Eng. J. 2023, 466, 143333. [Google Scholar] [CrossRef]
  9. Man, T.; Wang, K.; Zhang, J.; Niu, X.; Zhang, T. Theoretical Studies of High-nitrogen-containing Energetic Compounds Based on the s-Tetrazine Unit. Cent. Eur. J. Energ. Mater. 2013, 10, 171–189. [Google Scholar]
  10. Zhang, Y.; Parrish, D.; Shreeve, J. Derivatives of 5-nitro-1,2,3-2H-triazole—High performance energetic materials. J. Mater. Chem. A 2013, 1, 585–593. [Google Scholar] [CrossRef]
  11. Cao, W.; Qin, J.; Zhang, J.; Sinditskii, V. 4,5-Dicyano-1,2,3-Triazole—A Promising Precursor for a New Family of Energetic Compounds and Its Nitrogen-Rich Derivatives: Synthesis and Crystal Structures. Molecules 2021, 26, 6735. [Google Scholar] [CrossRef]
  12. Gu, H.; Xiong, H.; Yang, H.; Cheng, G. Tricyclic nitrogen-rich cation salts based on 1,2,3-triazole: Chemically stable and insensitive candidates for novel gas generant. Chem. Eng. J. 2021, 408, 128021. [Google Scholar] [CrossRef]
  13. Feng, S.; Yin, P.; He, C.; Pang, S.; Shreeve, J. Tunable Dimroth rearrangement of versatile 1,2,3-triazoles towards high performance energetic materials. J. Mater. Chem. A 2021, 9, 12291–12298. [Google Scholar] [CrossRef]
  14. Lai, Q.; Fei, T.; Yin, P.; Shreeve, J. 1,2,3-Triazole with linear and branched catenated nitrogen chains—The role of regiochemistry in Energetic Materials. Chem. Eng. J. 2020, 410, 128148. [Google Scholar] [CrossRef]
  15. Yao, W.; Xue, Y.; Qian, L.; Yang, H.; Cheng, G. Combination of 1,2,3-triazole and 1,2,4-triazole frameworks for new high-energy and low-sensitivity compounds. Energ. Mater. Front. 2021, 2, 131–138. [Google Scholar] [CrossRef]
  16. Yadav, A.K.; Kukreja, S.; Dharavath, S. Highly Promising Primary Explosive: A Metal-Free, Fluoro-Substituted Azo-Triazole with Unmatched Safety and Performance. JACS Au 2025, 5, 1031–1038. [Google Scholar] [CrossRef] [PubMed]
  17. Lease, N.; Spielvogel, K.D.; Davis, J.V.; Tisdale, J.T.; Klamborowski, L.M.; Cawkwell, M.J.; Manner, V.W. Halogenated PETN derivatives: Interplay between physical and chemical factors in explosive sensitivity. Chem. Sci. 2023, 14, 7044–7056. [Google Scholar] [CrossRef]
  18. Nie, X.; Lei, C.; Xiong, H.; Cheng, G.; Yang, H. Methylation of a triazole-fused framework to create novel insensitive energetic materials. Energ. Mater. Front. 2020, 1, 165–171. [Google Scholar] [CrossRef]
  19. Chen, F.; Song, S.; Zhang, Q.; Wang, Y. Modification of an N-methyl group toward a new energetic melt-castable material with a good energy–stability balance. Synlett 2024, 35, 2015–2021. [Google Scholar] [CrossRef]
  20. Guo, D.; Wei, Y.; Zybin, S.V.; Liu, Y.; Huang, F.; Goddard, W.A. III Detonation performance of insensitive nitrogen-rich nitroenamine energetic materials predicted from first-principles reactive molecular dynamics simulations. JACS Au 2024, 4, 1605–1614. [Google Scholar] [CrossRef]
  21. Lease, N.; Spielvogel, K.D.; Cawkwell, M.J.; Manner, V.W. Synthesis and investigation into explosive sensitivity for a series of new picramide explosives. J. Energ. Mater. 2024, 1–21. [Google Scholar] [CrossRef]
  22. Becke, A. Density functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  23. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations, I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  24. 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 A.2; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  25. Cardia, R.; Malloci, G.; Mattoni, A.; Cappellini, G. Effects of TIPS-Functionalization and Perhalogenation on the Electronic, Optical, and Transport Properties of Angular and Compact Dibenzochrysene. J. Phys. Chem. A 2014, 118, 5170–5177. [Google Scholar] [CrossRef]
  26. Cardia, R.; Malloci, G.; Rignanese, G.; Blasé, X.; Molteni, E.; Cappellini, G. Electronic and optical properties of hexathiapentacene in the gas and crystal phases. Phys. Rev. B 2016, 93, 235132. [Google Scholar] [CrossRef]
  27. Dardenne, N.; Cardia, R.; Li, J.; Malloci, G.; Cappellini, G.; Blasé, X.; Charlier, J.; Rignanese, G. Tuning Optical Properties of Dibenzochrysenes by Functionalization: A Many-Body Perturbation Theory Study. J. Phys. Chem. C 2017, 121, 24480–24488. [Google Scholar] [CrossRef]
  28. Antidormi, A.; Aprile, G.; Cappellini, G.; Cara, E.; Cardia, R.; Colombo, L.; Farris, R.; d’Ischia, M.; Mehrabanian, M.; Melis, C.; et al. Physical and Chemical Control of Interface Stability in Porous Si–Eumelanin Hybrids. J. Phys. Chem. C 2018, 122, 28405–28415. [Google Scholar] [CrossRef]
  29. Mocci, P.; Cardia, R.; Cappellini, G. Inclusions of Si-atoms in Graphene nanostructures: A computational study on the ground-state electronic properties of Coronene and Ovalene. J. Phys. Conf. Ser. 2018, 956, 012020. [Google Scholar] [CrossRef]
  30. Mocci, P.; Cardia, R.; Cappellini, G. Si-atoms substitutions effects on the electronic and optical properties of coronene and ovalene. New J. Phys. 2018, 20, 113008. [Google Scholar] [CrossRef]
  31. Kumar, A.; Cardia, R.; Cappellini, G. Electronic and optical properties of chromophores from bacterial cellulose. Cellulose 2018, 25, 2191–2203. [Google Scholar] [CrossRef]
  32. Szafran, M.; Koput, J. Ab initio and DFT calculations of structure and vibrational spectra of pyridine and its isotopomers. J. Mol. Struct. 2001, 565–566, 439–448. [Google Scholar] [CrossRef]
  33. Begue, D.; Carbonniere, P.; Pouchan, C. Calculations of Vibrational Energy Levels by Using a Hybrid ab Initio and DFT Quartic Force Field: Application to Acetonitrile. J. Phys. Chem. A 2005, 109, 4611–4616. [Google Scholar] [CrossRef] [PubMed]
  34. Kaya, S.; Kaya, C. A New equation based on ionization energies and electron affinities of atoms for calculating of group electronegativity. Comput. Theor. Chem. 2015, 1052, 42–46. [Google Scholar] [CrossRef]
  35. Politzer, P.; Martinez, J.; Murray, J.; Concha, M.; Toro-Labbé, A. An electrostatic interaction correction for improved crystal density prediction. Mol. Phys. 2009, 107, 2095–2101. [Google Scholar] [CrossRef]
  36. Spessard, G.O. ACD Labs/LogP dB 3.5 and ChemSketch 3.5. J. Chem. Inf. Comput. Sci. 1998, 38, 1250–1253. [Google Scholar] [CrossRef]
  37. Keshavarz, M.; Zamani, A. A simple and reliable method for predicting the detonation velocity of CHNOFCl and aluminized explosives. Cent. Eur. Energ. Mater. 2015, 12, 13–33. [Google Scholar]
  38. Keshavarz, M.; Pouretedal, H. An empirical method for predicting detonation pressure of CHNOFCl explosives. Thermochim. Acta 2004, 414, 203–208. [Google Scholar] [CrossRef]
  39. Zahari, N.; Montazeri, M.; Hoseini, S. Estimation of the Detonation Pressure of Co-crystal Explosives through a Novel, Simple and Reliable Model. Cent. Eur. J. Energ. Mater. 2020, 17, 492–505. [Google Scholar] [CrossRef]
  40. Meyer, R.; Köhler, J.; Homburg, A. Explosives, 6th ed.; Wiley-VCH: Weinheim, Germany, 2007; ISBN 978-3-527-31656-4. [Google Scholar]
  41. Rajakumar, B.; Arathala, P.; Muthiah, B. Thermal Decomposition of 2-Methyltetrahydrofuran behind Reflected Shock Waves over the Temperature Range of 1179–1361 K. J. Phys. Chem. A 2021, 125, 5406–5423. [Google Scholar] [CrossRef]
  42. Klapötke, T.M.; Witkowski, T.G. 5, 5′-Bis (2, 4, 6-trinitrophenyl)-2, 2′-bi (1, 3, 4-oxadiazole)(TKX-55): Thermally Stable Explosive with Outstanding Properties. ChemPlusChem 2016, 81, 357–360. [Google Scholar] [CrossRef]
  43. Pagoria, P. A comparison of the structure, synthesis, and properties of insensitive energetic compounds. Propellants Explos. Pyrotech. 2016, 41, 452–469. [Google Scholar] [CrossRef]
  44. Gutowski, Ł.; Cudziło, S. Synthesis and properties of novel nitro-based thermally stable energetic compounds. Def. Technol. 2021, 17, 775–784. [Google Scholar] [CrossRef]
  45. Agrawal, J.P.; Hodgson, R. Organic Chemistry of Explosives; John Wiley & Sons Ltd.: Chichester, UK, 2007; pp. 163–167. ISBN 0-470-02967-6. [Google Scholar]
  46. Ali, H.M.; Ali, I.H. Thermal and thermo-oxidative stability of a series of palladium and platinum ferrocenylamine sulfides and selenides. J. Chem. Res. 2022, 46, 17475198221141979. [Google Scholar] [CrossRef]
  47. Green, S.P.; Wheelhouse, K.M.; Payne, A.D.; Hallett, J.P.; Miller, P.W.; Bull, J.A. Thermal stability and explosive hazard assessment of diazo compounds and diazo transfer reagents. Org. Process Res. Dev. 2020, 24, 67–84. [Google Scholar] [CrossRef] [PubMed]
  48. Khunnawutmanotham, N.; Jumpathong, W.; Eurtivong, C.; Chimnoi, N.; Techasakul, S. Synthesis, cytotoxicity evaluation, and molecular modeling studies of 2,N10-substituted acridones as DNA-intercalating agents. J. Chem. Res. 2020, 44, 410–425. [Google Scholar] [CrossRef]
  49. Oxley, J.C.; Smith, J.L.; Ye, H.; McKenney, R.L.; Bolduc, P.R. Thermal stability studies on a homologous series of nitroarenes. J. Phys. Chem. 1995, 99, 9593–9602. [Google Scholar] [CrossRef]
  50. Jadhav, H.S.; Talawar, M.B.; Sivabalan, R.; Dhavale, D.D.; Asthana, S.N.; Krishnamurthy, V.N. Synthesis, characterization and thermolysis studies on new derivatives of 2,4,5-trinitroimidazoles: Potential insensitive high energy materials. J. Hazard. Mater. 2007, 143, 192–197. [Google Scholar] [CrossRef]
  51. Yang, W.; Lu, H.; Liao, L.; Fan, G.; Ma, Q.; Huang, J. Synthesis and single crystal structure of fully substituted polynitrobenzene derivatives for high energy materials. RSC Adv. 2018, 8, 2203–2208. [Google Scholar] [CrossRef]
  52. Nguyen, M.T.; Vansweevelt, H.; Vanquickenborne, L.G. Remarkable effect of fluorine and chlorine atoms on the stability of 1H-phosphirene. Chem. Berichte 1992, 125, 923–927. [Google Scholar] [CrossRef]
  53. Rice, B.M.; Byrd, E.F.C. Relationships with oxygen balance and bond dissociation energies. Comput. Theor. Chem. 2022, 22, 67–75. [Google Scholar] [CrossRef]
  54. Zeman, S.; Jungova, M. Sensitivity and Performance of Energetic Materials. Propellants Explos. Pyrotech. 2016, 41, 426–451. [Google Scholar] [CrossRef]
  55. Li, G.; Zhang, C. Review of the molecular and crystal correlations on sensitivities of energetic materials. J. Hazard. Mater. 2020, 398, 122910. [Google Scholar] [CrossRef] [PubMed]
  56. Politzer, P.; Murray, J.S. High Performance, Low Sensitivity: Conflicting or Compatible? Propellants Explos. Pyrotech. 2016, 41, 414–425. [Google Scholar] [CrossRef]
  57. Džingalašević, V.; Antić, G.; Mlađenović, D. Ratio of detonation pressure and critical pressure of high explosives with different compounds. Sci. Tech. Rev. 2004, 5, 72–76. [Google Scholar]
  58. Lal, S.; Gao, H.; Shreeve, J.M. New approach for predicting crystal densities of energetic materials. New J. Chem. 2024, 48, 17947–17952. [Google Scholar] [CrossRef]
  59. Wen, L.; Wang, B.; Yu, T.; Lai, W.; Shi, J.; Liu, M.; Liu, Y. Accelerating the search of CHONF-containing highly energetic materials by combinatorial library design and high-throughput screening. Fuel 2022, 310, 122241. [Google Scholar] [CrossRef]
  60. Tamuliene, J.; Sarlauskas, J. Polynitrobenzene Derivatives, Containing -CF3, -OCF3, and -O(CF2)nO- Functional Groups, as Candidates for Perspective Fluorinated High-Energy Materials: Theoretical Study. Energies 2024, 17, 6126. [Google Scholar] [CrossRef]
  61. Tselinskii, I.V. Applications of energetic materials in engineering, technology and national economy. Soros Educat. J. 1997, 11, 46–52. [Google Scholar]
  62. Fallas, J.A.; González, L.; Corral, I. Density functional theory rationalization of the substituent effects in trifluoromethyl-pyridinol derivatives. Tetrahedron 2009, 65, 232–239. [Google Scholar] [CrossRef]
  63. Kumar, P.; Ghule, V.D.; Dharavath, S. Advancing energetic chemistry: The first synthesis of sulfur-based C–C bonded thiadiazole-pyrazine compounds with a nitrimino moiety. Dalton Trans. 2024, 53, 19112–19115. [Google Scholar] [CrossRef]
  64. Chandler, J.; Ferguson, R.E.; Forbes, J.; Kuhl, A.L.; Oppenheim, A.K.; Spektor, R. Confined combustion of TNT explosion products in air. In Proceedings of the 8th International Colloquium on Dust Explosions, Schaumburg, IL, USA, 21–25 September 1998; Available online: https://www.osti.gov/biblio/3648 (accessed on 8 February 2025).
  65. Mahajan, T.; Bhargava, G.; Sharma, H. Theoretical Exploration on Polynitrobenzo/Heteroaryl Pyrazines as Useful High-Energy-Density Materials. ChemistrySelect 2025, 10, e0227. [Google Scholar] [CrossRef]
Figure 1. The molecular skeleton (on the left) together with the electrostatic potential (on the right). The deep red regions indicate areas of the highest electron density, whereas the deep blue regions correspond to the lowest electron density. I and II indicates the position of the substitutions.
Figure 1. The molecular skeleton (on the left) together with the electrostatic potential (on the right). The deep red regions indicate areas of the highest electron density, whereas the deep blue regions correspond to the lowest electron density. I and II indicates the position of the substitutions.
Chemengineering 09 00130 g001
Figure 2. Effect of substitutions and their number on thermal stability, as reflected by the cohesive energy.
Figure 2. Effect of substitutions and their number on thermal stability, as reflected by the cohesive energy.
Chemengineering 09 00130 g002
Figure 3. Effect of substitutions and their number on chemical stability, as reflected by the HOMO–LUMO gap.
Figure 3. Effect of substitutions and their number on chemical stability, as reflected by the HOMO–LUMO gap.
Chemengineering 09 00130 g003
Figure 4. Effect of substitutions and their number on electronegativity.
Figure 4. Effect of substitutions and their number on electronegativity.
Chemengineering 09 00130 g004
Figure 5. Effect of substitutions and their number on the impact of the stimuli. The abbreviated compounds, along with substitutions, are given.
Figure 5. Effect of substitutions and their number on the impact of the stimuli. The abbreviated compounds, along with substitutions, are given.
Chemengineering 09 00130 g005
Table 1. The parameters describe the chemical and thermal stability. Here, BDE—cohesive energy; Hard—chemical hardness; Elec—electronegativity; Gap—the gap between the highest occupied and the lowest unoccupied orbitals; Ind—hardness index.
Table 1. The parameters describe the chemical and thermal stability. Here, BDE—cohesive energy; Hard—chemical hardness; Elec—electronegativity; Gap—the gap between the highest occupied and the lowest unoccupied orbitals; Ind—hardness index.
AbbreviationSubstitutionsGap, eVHard, eVElec, eVIndBDE
Acri1None3.791.905.960.865.94
Acri22 OH3.461.736.040.835.83
Acri2_II2 OH3.641.825.920.855.83
Acri74 HO3.881.945.880.875.75
Acri32 Cl3.821.915.960.865.82
Acri84 Cl3.201.605.580.805.68
Acri42 F3.901.956.070.875.93
Acri94 F4.062.036.160.885.92
Acri52 NH23.471.745.580.835.84
Acri104 NH23.651.835.130.855.76
Acri62 NCH3NO23.681.846.080.855.55
Acri114 NCH3NO23.411.716.260.835.35
Acri12Picryl3.211.606.340.815.86
Acri13Picryl and 2 NH23.221.615.850.815.81
Acri14Picryl and 4 NH23.111.555.540.795.78
Acri152 NH2 in Picryl and 4 NH22.991.505.240.784.76
Table 2. The density obtained. Here, ρgaus, is calculated from molecular weight and B3LYP/cc-pVTZ molar volume, ρKev is evaluated by Politzer’s equations [35], and ρacd is obtained by the approach implemented in ACD/ChemSketch [36].
Table 2. The density obtained. Here, ρgaus, is calculated from molecular weight and B3LYP/cc-pVTZ molar volume, ρKev is evaluated by Politzer’s equations [35], and ρacd is obtained by the approach implemented in ACD/ChemSketch [36].
AbbreviationSubstitutionsρgaus, g/cm3ρKev, g/cm3ρacd, g/cm3
Acri1None1.701.601.82
Acri22 OH1.981.892.01
Acri2_II2 OH1.951.862.01
Acri74 HO1.911.812.20
Acri32 Cl1.811.721.93
Acri84 Cl2.102.022.02
Acri42 F1.841.751.92
Acri94 F1.831.732.01
Acri52 NH21.881.791.92
Acri104 NH21.661.562.02
Acri62 NCH3NO21.851.751.91
Acri114 NCH3NO21.711.611.96
Acri12Picryl1.751.651.95
Acri13Picryl and 2 NH21.941.852.02
Acri14Picryl and 4 NH22.142.062.09
Acri152 NH2 in Picryl and 4 NH21.961.872.15
Table 3. Detonation pressures calculated using different density estimation methods. Here, Pgaus, is the pressure calculated using density derived from molecular weight and B3LYP/cc-pVTZ-optimized molar volume; PKev is the pressure obtained using density estimated via Politzer’s equation [35], and Pacd is the pressure based on density calculated with the ACD/ChemSketch approach [36].
Table 3. Detonation pressures calculated using different density estimation methods. Here, Pgaus, is the pressure calculated using density derived from molecular weight and B3LYP/cc-pVTZ-optimized molar volume; PKev is the pressure obtained using density estimated via Politzer’s equation [35], and Pacd is the pressure based on density calculated with the ACD/ChemSketch approach [36].
AbbreviationSubstitutionsPgaus, GPaPKev, GPaPacd, GPa
Acri1None9.768.4911.43
Acri22 OH14.4513.0614.91
Acri2_II2 OH13.9812.5814.91
Acri74 HO14.0312.5619.05
Acri32 Cl11.5310.3313.28
Acri84 Cl15.9314.6714.70
Acri42 F12.2210.8613.35
Acri94 F12.5611.0915.31
Acri52 NH214.2912.8115.07
Acri104 NH212.4010.7919.03
Acri62 NCH3NO216.1514.4417.26
Acri114 NCH3NO216.6414.6222.19
Acri12Picryl12.9211.3316.33
Acri13Picryl and 2 NH211.5310.3313.28
Acri14Picryl and 4 NH224.3622.4923.02
Acri152 NH2 in Picryl and 4 NH218.9117.1223.06
Table 4. Detonation velocity calculated using different density estimation methods. Here, Dgaus, is the detonation velocity calculated using density derived from molecular weight and B3LYP/cc-pVTZ-optimized molar volume; DKev is the velocity obtained using density estimated via Politzer’s equation [35], and Dacd is the velocity based on density calculated with the ACD/ChemSketch approach [36].
Table 4. Detonation velocity calculated using different density estimation methods. Here, Dgaus, is the detonation velocity calculated using density derived from molecular weight and B3LYP/cc-pVTZ-optimized molar volume; DKev is the velocity obtained using density estimated via Politzer’s equation [35], and Dacd is the velocity based on density calculated with the ACD/ChemSketch approach [36].
AbbreviationSubstitutionsDgaus, km/sDKev, km/sDacd, km/s
Acri1None5.705.485.98
Acri22 OH6.436.236.49
Acri2_II2 OH6.366.166.49
Acri74 HO6.376.157.04
Acri32 Cl5.995.806.26
Acri84 Cl6.636.466.46
Acri42 F6.105.896.27
Acri94 F6.155.926.55
Acri52 NH26.416.196.52
Acri104 NH26.135.887.04
Acri62 NCH3NO26.666.436.81
Acri114 NCH3NO26.736.457.42
Acri12Picryl6.215.966.69
Acri13Picryl and 2 NH25.995.806.26
Acri14Picryl and 4 NH27.667.457.51
Acri152 NH2 in Picryl and 4 NH27.026.797.52
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tamuliene, J.; Sarlauskas, J. Theoretical Design of Acridone-Core Energetic Materials: Assessment of Detonation Properties and Potential as Insensitive, Thermally Stable High-Energy Materials. ChemEngineering 2025, 9, 130. https://doi.org/10.3390/chemengineering9060130

AMA Style

Tamuliene J, Sarlauskas J. Theoretical Design of Acridone-Core Energetic Materials: Assessment of Detonation Properties and Potential as Insensitive, Thermally Stable High-Energy Materials. ChemEngineering. 2025; 9(6):130. https://doi.org/10.3390/chemengineering9060130

Chicago/Turabian Style

Tamuliene, Jelena, and Jonas Sarlauskas. 2025. "Theoretical Design of Acridone-Core Energetic Materials: Assessment of Detonation Properties and Potential as Insensitive, Thermally Stable High-Energy Materials" ChemEngineering 9, no. 6: 130. https://doi.org/10.3390/chemengineering9060130

APA Style

Tamuliene, J., & Sarlauskas, J. (2025). Theoretical Design of Acridone-Core Energetic Materials: Assessment of Detonation Properties and Potential as Insensitive, Thermally Stable High-Energy Materials. ChemEngineering, 9(6), 130. https://doi.org/10.3390/chemengineering9060130

Article Metrics

Back to TopTop