Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Quantum chemical calculations for predicting the partitioning of drug molecules in the environment

Lukas Wittmanna, Tunga Salthammer*b and Uwe Hohmc
aMulliken Center for Theoretical Chemistry, Institute for Physical and Theoretical Chemistry, University of Bonn, 53115 Bonn, Germany
bFraunhofer WKI, Department of Material Analysis and Indoor Chemistry, 38108 Braunschweig, Germany. E-mail: tunga.salthammer@wki.fraunhofer.de
cInstitute of Physical and Theoretical Chemistry, University of Braunschweig – Institute of Technology, 38106 Braunschweig, Germany

Received 8th July 2025 , Accepted 1st October 2025

First published on 3rd October 2025


Abstract

Regional and temporal trends in legal and illicit drug use can be tracked through monitoring of municipal wastewater, ambient air, indoor air, and house dust. To assess the analytical result for the selected environmental matrix, reliable information on the partitioning of the target substance between the different compartments is required. The logarithmic partition coefficients octanol/water (log[thin space (1/6-em)]KOW), octanol/air (log[thin space (1/6-em)]KOA) and air/water (log[thin space (1/6-em)]KAW) are usually applied for this purpose. Most drug molecules are semi-volatile compounds with complex molecular structures, the handling of which is subject to legal regulations. Chemically, they are often acids, bases, or zwitterions. Consequently, the physical and chemical properties are in most cases not determined experimentally but derived from quantitative structure–activity relationships (QSARs). However, the lack of experimental reference data raises questions about the accuracy of computed values. It therefore seemed appropriate and necessary to calculate partition coefficients using alternative methods and compare them with QSAR results. We selected 23 substances that were particularly prominent in European and US drug reports. Different quantum mechanical methods were used to calculate log[thin space (1/6-em)]KOW, log[thin space (1/6-em)]KOA, and log[thin space (1/6-em)]KAW for the undissociated molecule as a function of temperature. Additionally, the logarithmic hexadecane/air partition coefficient log[thin space (1/6-em)]KHdAL and the logarithmic vapor pressure of the subcooled liquid log[thin space (1/6-em)]PL were determined in the temperature range 223 < T/K < 333. Despite the sometimes high variability of the parameters, it is possible to estimate how an investigated substance distributes between air, water and organic material.



Environmental significance

According to the United Nations, drug use is growing faster than the global population. Both the increasing demand and the diversity of supply are exacerbating a social problem that is certainly cause for serious concern. A proven method in forensic drug analysis is detection in wastewater, soil, house dust, etc. However, a realistic assessment of the distribution of drugs in environmental matrices requires reliable values for partition coefficients and vapor pressure. These are often unavailable because the measurements are complex or because of legal hurdles. Therefore, 23 representative substances were selected from the vast amount of available drugs, and the physical properties were calculated using quantum mechanical methods. The results were subjected to a critical analysis for their plausibility.

1 Introduction

Psychoactive substances are closely linked to human cultural history. As early as 3000 BC, hallucinogenic mushrooms of the genus Psilocybe spp., whose main ingredient is psilocybin, were consumed in Mesoamerica.1 Also in South America, chewing coca leaves has been a popular practice for centuries to combat fatigue and hunger. The sleep-inducing and pain-relieving effects of opiates were already recognized in prehistoric times and became a frequent component of medicine from the 16th century onwards. The development of modern chemistry in the 19th century made it possible to produce amphetamines, morphine and heroin, which were often used in the military sector.2 The first prohibitions on drug trafficking in the 19th century were motivated by trade policy. Health-related measures were introduced starting in the 1920s. Since the 1980s, the illicit drug market has been flooded with synthetic products known as so-called designer drugs. For example, Gerona3 lists 21 structurally different synthetic cannabinoids. Fentanyl was first synthesized in 1960, and a wide range of fentanyl analogues are now also available.4 For years, it has been evident that synthetic chemistry is progressing faster than the drug regulations by authorities. Nitrous oxide (N2O), known as laughing gas, is not only used as an anaesthetic in medicine or in the kitchen to whip cream, but more recently also abused in some countries without restrictions as a so-called lifestyle drug.5

One method for tracking drug usage is the analysis of environmental matrices. Wastewater not only allows to identify which drugs are being consumed locally, but also to identify delivery routes and emerging or declining trends.6–8 In the indoor environment, house dust is a suitable matrix for screening purposes.9 Of equal importance is the question of how a chemical substance distributes between different environmental matrices. Outdoors, the simplest model includes the compartments air, water, and organic phase.10 In reality, however, the dynamics are considerably more complex. For example, semi volatile airborne substances require consideration of the partitioning between the gas and the particle phase11,12 or the aqueous and gas phase.8 Analogously, the same applies to indoor areas with gas phase, particle phase, house dust, clothing and other surfaces.13–17

To assess the environmental equilibrium partitioning of a substance, parameters are required that describe molecular interactions between different phases. These are essentially the octanol/water partition coefficient (KOW), the octanol/air partition coefficient (KOA), the air/water partition coefficient (KAW) and the saturation vapor pressure of the subcooled liquid (PL).10 A frequently used quantity in linear-free-energy-relationships (LFER) is the hexadecane/air partition coefficient KHdA and its logarithmic form log[thin space (1/6-em)]KHdAL, respectively.10 However, these parameters are often difficult to measure or require significant experimental effort. Therefore, early attempts were made to predict physical and chemical properties from the molecular structure.18 With the increase in computing power, these models became more complex.19,20 Advanced quantitative-structure–activity-relationships (QSAR) models use molecular descriptors, often in conjunction with machine learning.21–23 Quantum mechanical (QM) methods provide a more fundamental approach to obtain these quantities of interest. These methods can, among other things, predict the solvation energy ΔGsolv in the respective solvents of interest.

Compared to other substances, the properties of drug molecules have not been investigated as thoroughly experimentally – often due to their complex structure and respective legal requirements. The use of theoretical methods is therefore common, which naturally raises questions about their accuracy. It is known that popular prediction tools such as EpiSuite and SPARC provide unreliable values for large molecules.24 On the other hand, the aforementioned quantum chemical methods are often not suffering from such problems, but require advanced know how and computational effort.

For this study, we selected 23 molecules representing a broad spectrum of currently legally and illegally consumed drugs. This includes both the long-known and newly emerging substances. The partition coefficients log[thin space (1/6-em)]KOW, log[thin space (1/6-em)]KOA, log[thin space (1/6-em)]KAW, log[thin space (1/6-em)]KHdA, and PL were calculated using quantum chemical methods. The obtained results were compared with data from popular QSAR/QSPR tools and databases. The comparison data had partly experimental and partly predictive background. Since some molecules are weak bases or acids, the pKa values were also taken into account. Our results are not only of interest regarding the possibilities and limitations of using QM and QSAR computed parameters for estimating the partitioning of drug molecules in the environment. We additionally discuss the temperature dependence of these parameters in the interval 283–308 K based on the calculated temperature-dependent free energy of solvation ΔGsolv.

2 Compounds and methods

2.1 Drug molecules

For the selection of drug molecules, three main references were used: (i) the report of the German Bundeskriminalamt (BKA) (Federal Criminal Police Office) on drug-related crime in Germany 2023,25 (ii) the World Factbook: Illicit Drugs of the U.S. Central Intelligence Service (CIA)26 and (iii) the Commonly Used Drugs Charts of the U.S. Institute on Drug Abuse (NIDA).27 From the multitude of possibilities, we selected 23 molecules. The criteria included their relevance for production and consumption, their future significance, their diversity of structural differences, their detectability in air, water, and bio-monitoring, as well as the required computational effort. All molecules investigated are listed in Table 1 with their abbreviation and CAS number; their respective structure is shown in Fig. 1. Boiling points and pKa values were taken from the OPERA database.28
Table 1 Investigated drugs, their abbreviations and CAS numbers. Except for N2O, the boiling points and pKa were taken from OPERA.28 The boiling point of N2O was taken from Rumble et al.,29 n.a. = not available
Drug Abbr. CAS no. M (g mol−1) BP (°C) pKa
1-Benzylpiperazine BZP 2759-28-6 176.26 289 9.80
4-Hydroxybutanoic acid HBA 591-81-1 104.11 219 5.23
Amphetamine AMP 300-62-9 135.21 203 9.87
Cathinone KHAT 71031-15-7 149.19 261 6.39
Cocaine COC 50-36-2 303.35 332 7.78
Midomafetamine (formerly ecstasy) MDMA 42542-10-9 193.25 285 9.96
Fentanyl FEN 437-38-7 336.47 353 7.15
Heroin HER 561-27-3 369.42 398 8.28
Ketamine KET 6740-88-1 237.74 312 8.08
Lysergic acid diethylamide LSD 50-37-3 323.42 361 5.66
Mephedrone MPD 1189805-46-6 177.24 274 6.63
Mescaline MES 54-04-6 211.26 294 9.75
Methadone MTD 76-99-3 309.45 353 6.76
Methamphetamine METH 537-46-2 149.23 308 9.97
Methaqualone MEQ 72-44-6 250.30 332 2.23
Methylphenidate MEP 113-45-1 233.31 307 10.54
Morphine MOR 57-27-2 285.34 385 8.06
Nitrous oxide N2O 10024-97-2 44.01 −89 n.a.
Pethidine PET 57-42-1 247.33 295 8.70
Phencyclidine (angel dust) PCP 77-10-1 243.39 288 10.49
Psilocybin PSY 520-52-5 284.25 286 3.63
Δ9-Tetrahydrocannabinol THC 1972-08-3 314.47 328 7.80
Trifluoromethylphenylpiperazine TFPP 15532-75-9 230.23 289 6.91



image file: d5em00524h-f1.tif
Fig. 1 Structures of the investigated drugs with their abbreviations, see Table 1 for the chemical names.

2.2 Theory

We consider the partitioning of a substance between two adjacent phases α and β. The ratio of the equilibrium concentrations is known as the partition coefficient Kαβ = cα/cβ. In the case of missing solute–solute correlations, Kαβ is directly related to the free energy of transfer ΔGαβ of solute between the phases α and β via30,31
 
image file: d5em00524h-t1.tif(1)
where R is the molar gas constant and T the temperature. According to eqn (1), calculation of ΔGαβ allows for a direct determination of the partition coefficient Kαβ.32 If on the other hand, α denotes the pure ideal gas phase with concentration cα = pα/(RT) and β its pure condensed phase with concentration cβ = ρβ/M, eqn (1) can directly be used to calculate the vapor pressure PLpα (analogous to ref. 33) via
 
image file: d5em00524h-t2.tif(2)
where M is the molar mass and ρβ the density of phase β. If T < Tfus of the substance, PL in eqn (2) refers to the vapor pressure of the subcooled liquid. For arbitrary phases β, the density of the phase ρβ is not always known. For this reason, we set it to the density of liquid water of 997 kg m−3. This of course introduces an error, which, however, is lower than 0.1 kcal mol−1 in terms of ΔGαβ and thus negligible in terms of the accuracy of our computational workflow.33,34 In addition, we use the geometries obtained in octanol for the ones in the subcooled liquid. Most drug-like compounds in their condensed (subcooled-liquid) state have a relative dielectric permittivity εr ≈ 8–15, close to that of 1-octanol.35 We therefore optimize geometries in an octanol-like dielectric as a proxy for each compound's own liquid. We do not optimize each drug in its individual εr because experimental compound-specific values are generally unavailable; although εr can be computed in principle, there is no straightforward, robust protocol for diverse drug-like molecules, the required calculations are computationally demanding, and the setup is often substance-specific.36–38
2.2.1 Ensemble-averaged free energies. To obtain the quantities of interest herein, the free energy of a substance in all phases (gas, water, octanol, and hexadecane) has to be known. As non-rigid molecules often have multiple relevant conformers contributing to the total free energy, we need to consider the ensemble of conformers that contribute to the total free energy of that substance. The following theory and equations are based on previous work and are explained in more detail there (e.g. ref. 34, 39 and 40). The free energy of a single conformer i is given by the gas-phase electronic energy Eel, the thermostatistical correction, and the possibly needed solvation free energy contribution ΔGsolv(T).
 
Gi(T) = Eel,i + Gtrv,i(T) + ΔGsolv,i(T) (3)

The electronic energy can be obtained with any electronic structure method – and is obtained in our case with Density Functional Theory (DFT). The thermostatistical correction at temperature Gtrv,i(T) accounts for translation, rotation, and vibration degrees of freedom and including the zero-point vibrational and volume work terms. To obtain the free energy of the substance from its individual conformers, the individual conformer free energies Gi need to be Boltzmann-weighted via eqn (4).

 
image file: d5em00524h-t3.tif(4)

The free energy of a conformer ensemble should additionally include a conformational free energy part Gconf that stems from conformational entropy −TSconf as a result of mixing multiple populated conformers.34 This contribution is, however, most often neglected due to the huge amount of computation cost needed to accurately determine Sconf.41

Generally, the enthalpic contributions are very similar for gas- and solution-phases, similarly to the vibrational entropy contribution SgasvSsolvv.42 The terms left are thus just the change in rotational and translational entropy. Because this change is generally already implicitly included in the solvation free energy of quantum chemical solvation models due to the parameterization on experimental data, we neglect the additional thermostatistical correction Gtrv,i(T).

2.2.2 Accounting for solvation effects. To obtain free energy of a conformer (eqn (3)) in solution, the solvation contribution ΔGsolv (i.e., the solvation free energy) is needed. The solvation free energy describes the change in free energy when transferring a substance from the gas phase to a liquid.43 In computational chemistry, it is often decomposed into several contributions
 
image file: d5em00524h-t4.tif(5)
where ΔGES are the electrostatic solvent–solute interactions (i.e., polarization), ΔGnon-ES denote non-electrostatic contributions (e.g., cavity formation), ΔGN is the nuclear relaxation term describing the change in geometry when transferring a solute from the gas to liquid phase, and image file: d5em00524h-t5.tif is the so-called standard-state correction.44–46 This term is given by
 
image file: d5em00524h-t6.tif(6)
which depends on the ideal gas constant R, the standard concentration c = 1 mol L−1 and standard pressure p = 1 atm and amounts to 1.89 kcal mol−1 at 298.15 K.47

The nuclear relaxation contribution, ΔGN, captures the change in energy associated with the solute's structural reorganization upon transferring from the gas to the solution phase.48 Concretely, it is computed as the difference between the gas phase electronic energy of the molecule at its solution phase optimized geometry, and the gas-phase electronic energy at its gas-phase optimized geometry. Because the gas phase geometry is a minimum on the gas phase potential energy surface, any other geometry – including the one favored in solvation – can only have equal or higher gas-phase energy. Consequently, ΔGN will always be equal or larger than zero. Note that by performing full geometry optimizations both in the gas phase and in solution and using those respective electronic energies, the nuclear relaxation contribution is automatically included. Generally, it is very desirable to use the optimized geometries in all respective phases. In cases like, e.g., strong zwitter ions, the solute in solution will be zwitterionic, whereas in gas phase it will be neutral. Not using the respective optimized geometries will yield large errors for those systems. However, many implicit solvation models are parameterized on gas phase structures only, partially absorbing the nuclear relaxation into the parameterization and the model itself.44 It is, however, not possible to correctly account for the nuclear relaxation of very complex solutes in this way, which can lead to significant errors.48 In the following, we will shortly introduce the used implicit quantum-chemical solvation models and their workings.


2.2.2.1 SMD. The universal solvation model based on electron density44 is an improvement over conductor-like polarizable continuum model (CPCM)49,50 by adding a so-called cavity-dispersion solvent-structure term, that accounts for interactions that regular CPCM does not account for like the cavity creation and short-range dispersion interactions.
2.2.2.2 COSMO-RS and openCOSMO-RS. The conductor-like screening model for real solvents51,52 is also based on CPCM, however, it uses the obtained surface charges of solute and solvent. These surface charges are then used to compute pairwise interaction energies and combinatorial contributions within an empirical statistical thermodynamic framework, from which it directly predicts the solvation free energy. openCOSMO-RS is the fully open-source implementation of COSMO-RS and both differ mainly in their parameterization.

Besides the three mentioned models, the aforementioned conductor-like polarizable continuum model49,50 is tested and the results are shown in the SI. As CPCM only accounts for electrostatic interactions, it is not suitable to compute accurate solvation free energies or partition coefficients.

2.3 Computational workflow

To assemble the solvent-specific conformer–tautomer ensembles described in Section 2.2.1, we use the conformer–rotamer ensemble sampling tool, CREST, which uses meta-dynamics to gently bias molecules out of their local energy minima and into unexplored regions of the prototropic-tautomeric and conformational space.53,54 Sampling with CREST using the GFN2-xTB tight-binding method, allowing an extensive exploration at a low cost. Because GFN2-xTB is a semi-empirical tight-binding method, however, a higher level optimization of the resulting ensemble is needed afterwards. For this, we use the command-line energetic sorting tool, CENSO, which carries out full DFT geometry optimizations and re-ranking.34,55 This multi-level approach – broad, meta-dynamics-driven exploration followed by high-level DFT refinement – yields accurate ensembles at a reasonable cost. This workflow is carried out for each substance in each phase separately, yielding a refined ensemble for each phase, optimized at DFT (r2SCAN-3c) level. These structures are then used to carry out the calculation of solvation free energies. The workflow is shown in Fig. 2.
image file: d5em00524h-f2.tif
Fig. 2 Sketch of the employed CREST + CENSO workflow. The workflow is run separately for each drug and solvent combination.

2.4 Computational details

Quantum chemical calculations were performed with xTB 6.7.0[thin space (1/6-em)]56 and ORCA 6.0.1.57 The workflows used Open Babel 3.1.0,58 CREST 3.0,54,59 CENSO 2.0,34 and MolBar 1.1.0.60 The sampling and subsequent optimization was done using GFN2-xTB61 and the r2SCAN-3c composite method.62 Range-separated hybrid calculations are on the ωr2SCAN-D4/def2-TZVPPD level.63–67 The matching def and def2 effective small core potentials (ECPs)68,69 for heavy elements with Z > 36 were generally employed for all calculations. Matching general-purpose auxiliary basis sets are constructed on the fly using Stoychev et al.'s automatic generation of auxiliary basis sets in ORCA.70 The RIJCOSX71–73 approximation was used for all hybrid calculations. COSMO-RS49,51,74 is calculated using TurboMole 7.9.0[thin space (1/6-em)]75 with COSMOtherm C30-1601 and uses per default BP86/def-TZVP level of theory.76,77 openCOSMO-RS52 results are obtained using ORCA and utilize BP86/def2-TZVPD. Solvation contributions of the CPCM50,78 and SMD44 models are obtained using the r2SCAN-3c composite method. The ALPB solvation model48 was employed for all semi-empirical solution phase optimizations and the CPCM solvation model for all r2SCAN-3c-based optimizations.

2.5 Databases and literature data

The Open (Quantitative) Structure–activity/property Relationship App (OPEn(q)saRApp = OPERA)28 provides QSAR/QSPR models for chemical properties and was primarily used for comparison with the quantum chemically calculated data. The workflows behind OPERA consist of complex algorithms,79,80 whose structure is explained by accompanying documents directly in the database. The basis is an implemented dataset from which a training set and a test set are randomly generated. The desired property of the respective target substance is computed using the k-nearest neighbor (kNN) method with k = 5. This algorithm identifies the five substances from the dataset whose molecular descriptors are closest to the target substance and uses the Euclidean distance as metric.28 If the target substance itself is among the nearest neighbors, the value is marked as “experimental”. This can be checked using the CompTox database81 (note that we used Version 2.5.3 – April 8, 2025), which also provides statistical parameters for the reliability of the value calculated by OPERA. Experimental data were particularly available for log[thin space (1/6-em)]KOW, so we differentiated accordingly in this case (see SI). We also noticed that the log[thin space (1/6-em)]KOW values often match the values of Hansch et al.82 In several cases, Hansch et al.82 refer to inaccessible sources, so the quality of the log[thin space (1/6-em)]KOW data could not always be verified. For 13 substances we found experimentally determined Linear Solvation Energy Relationship (LSER) descriptors and logarithmic hexadecane/air partition coefficients in the UFZ-LSER database.83 For the other 10 substances, the descriptors and log[thin space (1/6-em)]KHdA(L) values were computed from their SMILES structures using a tool implemented in the UFZ-LSER database and described by Brown.84 Sander's85 database provides access to a compilation of Henry's law constants for organic and inorganic species in water from literature data. Additional calculated and measured literature data were considered and discussed accordingly. All comparison data extracted from the databases refer to 298 K. Experimental data for other temperatures, when available, are also listed in the SI.

3 Results and discussion

To systematically evaluate the results of the quantum mechanical calculations and the literature values, we based all comparisons on the data generated with COSMO-RS. Accordingly, Table 2 contains all partition coefficients and vapor pressures calculated with COSMO-RS for the 23 target molecules. The corresponding ΔG can be found in the SI, as well as the results of the openCOSMO-RS and SMD calculations and the associated costs for the computational workflow. The results obtained with the other QM methods are also listed, but are not discussed since they do not provide any further insights. With the exception of N2O, log[thin space (1/6-em)]KOA, log[thin space (1/6-em)]KOW, log[thin space (1/6-em)]KAW (calculated from the Henry solubility) and log[thin space (1/6-em)]PL were obtained from OPERA. All log[thin space (1/6-em)]KHdA (L) values were extracted from the UFZ-LSER database. Experimental vapor pressure data were also considered. The literature data are compiled with the respective sources in the SI. As OPERA does not provide data for N2O, the log[thin space (1/6-em)]KOW was estimated from the Ostwald coefficients for this substance. Using the data from Makranczy et al.86 for 1-octanol and from Gabel and Schultz87 for water, log[thin space (1/6-em)]KOW = 0.55 is obtained at T = 298 K. This is in fair agreement with the values of log[thin space (1/6-em)]KOW = 0.36 and 0.43 from Hansch et al.82,88 The comparisons of COSMO-RS results with the other quantum mechanical methods and the literature data for partition coefficients at 298 K are shown in Fig. 3.
Table 2 Partition coefficients and vapor pressures of the subcooled liquid at 298 K for the drug molecules given in Table 1, quantum mechanically calculated using COSMO-RS
Drug log[thin space (1/6-em)]KOW log[thin space (1/6-em)]KOA log[thin space (1/6-em)]KAW log[thin space (1/6-em)]KHdA log[thin space (1/6-em)]PL (Pa)
BZP 2.45 7.79 −5.34 6.64 −0.42
HBA −0.47 7.38 −7.85 3.78 −0.73
AMP 2.15 6.24 −4.08 4.87 1.27
KHAT 4.96 6.93 −1.97 5.3 0.01
COC 1.06 9.37 −8.31 10.3 −3.25
MDMA 2.14 6.26 −4.12 6.22 0.54
FEN 4.47 13.41 −8.93 12 −6.79
HER 3.13 10.51 −7.38 9.36 −4.74
KET 1.63 9.67 −8.04 7.6 −3.56
LSD 3.43 13.91 −10.48 3.44 −7.82
MPD 2.47 7.28 −4.81 6.27 −0.38
MES 1.15 9.85 −8.69 7.83 −2.63
MTD 5.20 10.01 −4.81 9.59 −3.93
METH 2.94 5.73 −2.78 5.25 1.48
MEQ 2.29 10.25 −7.95 9.3 −4.26
MEP 3.94 8.88 −4.94 8.28 −2.16
MOR 2.10 12.02 −9.92 6.97 −5.87
N2O 0.40 0.06 0.34 −0.29 6.63
PET 3.75 8.72 −4.97 8.17 −2.10
PCP 5.19 8.37 −3.18 8.35 −1.96
PSY −2.97 19.53 −22.5 13.58 −14.80
THC 6.97 14.52 −7.55 12.83 −7.31
TFPP 2.85 7.7 −4.85 6.81 −1.17



image file: d5em00524h-f3.tif
Fig. 3 Scatter diagrams of the partition coefficients at 298 K calculated using different methods, plotted against COSMO-RS. For the log[thin space (1/6-em)]KHdA the UFZ-LSER data are used. The log[thin space (1/6-em)]KAW according to OPERA were calculated from Henry solubilities using eqn (9). Notable points are marked with their respective abbreviation in the figure.

In contrast to our previous work on partition coefficients32 and vapor pressures,33 the problem here is that there is hardly any reliable experimental data available for the 23 target compounds. Therefore, only computed values can be discussed to estimate their environmental behavior. Since most of the compounds are weak acids or bases, dissociation effects must also be considered.

3.1 Octanol/water partitioning

For the partition coefficients between 1-octanol and water, we generally find good agreement between all methods. openCOSMO-RS predicts slightly larger log[thin space (1/6-em)]KOW values compared to COSMO-RS, whereas SMD agrees well with the COSMO-RS data. Compared to the QM results, the OPERA data are broader distributed, with a slight tendency to underestimate COSMO-RS. The methods disagree the most for PSY, MPD and KHAT. For PSY, SMD and OPERA predict log[thin space (1/6-em)]KOW of 0.48 and 1.18, whereas the COSMO-RS variants predict −2.97 and −2.20. The negative values are plausible because the zwitterion is the thermodynamically preferred tautomer and its solubility in water is higher than in organic solvents.89 For MPD and KHAT, OPERA predicts a log[thin space (1/6-em)]KOW around 2–3[thin space (1/6-em)]log units lower than the quantum chemical methods. Notable is also N2O, where SMD predicts a log[thin space (1/6-em)]KOW of around −1.85, whereas all other methods predict values between 0.36 and 0.83.

Assessing the root cause of SMD's misprediction is challenging, since the model is highly empirical and relies on parameters fitted to reproduce experimental data. SMD predicts a higher (i.e. less favourable) solvation free energy for N2O in 1-octanol of around 3 kcal mol−1, which leads to the wrong sign in the prediction of the log[thin space (1/6-em)]KOW. Similar shortcomings of SMD have been noted elsewhere in the literature (e.g. in ref. 90–92). However, it should still be noted, that an error of a few kcal mol−1 is still reasonable and often to be expected for implicit solvation models in more difficult cases.92,93

As mentioned in a previous section, it is difficult to assess the quality of the OPERA data, because no information about the origin of the data is provided. It can only be distinguished between experimental and computed values (see Fig. 3). In this context, “experimental” means that the substance is present in the OPERA dataset. However OPERA provides information about the training set and test set, the confidence level, and the five nearest neighbors, which makes it easier to assess the validity of the calculated value. In the case of morphine (MOR), for example, it can be assumed that the log[thin space (1/6-em)]KOW = 0.89 given by OPERA is the experimentally determined value published by Avdeef et al.94 Moreover, the work of Avdeef et al.94 touches the important aspect that most of the 23 target molecules are acids or bases, but the data in Table 2 refer to the undissociated molecule. The pH dependence of log[thin space (1/6-em)]KOW is discussed in the next section. Furthermore, psilocybin can form a zwitterion through intramolecular proton transfer, and the solvation energy ΔG naturally depends strongly on the tautomeric form.89 Here it is obvious that the OPERA algorithm fails, because the five nearest neighbors are molecules without zwitterionic character and OPERA probably cannot take the zwitterionic state into account at all.

3.2 Dependence of log[thin space (1/6-em)]KOW on the pH value

When considering octanol/water partitioning, it is particularly important to note that at 298 K, a considerable amount of water dissolves in 1-octanol, with a molar fraction of 0.27 – while the solubility of 1-octanol in water is significantly lower.95 Therefore, log[thin space (1/6-em)]KOW usually refers to water-saturated 1-octanol. A further problem arises for ionizable substances.96 Hansch and Leo97 applied correction terms in their molecular fragment-based CLOGP method for calculating log[thin space (1/6-em)]KOW to take into account the properties of ionizable compounds and zwitterions. Experimental methods often use buffer solutions, so that the log[thin space (1/6-em)]KOW of the undissociated acid or base is obtained. To better understand the environmental behavior and bioavailability of ionizable compounds, the distribution coefficient DOW was introduced, which is defined as the ratio of the neutral and charged species in the lipid and aqueous phase at a given pH. In the pH range where the molecule is predominantly non-ionized, log[thin space (1/6-em)]DOW = log[thin space (1/6-em)]KOW. Based on the Henderson–Hasselbalch equation for buffer solutions, Scherrer and Howard98 developed relationships between log[thin space (1/6-em)]DOW and log[thin space (1/6-em)]KOW for acids (eqn (7)) and bases (eqn (8)) under the assumption that charged species are not present in the octanol phase.
 
image file: d5em00524h-t7.tif(7)
 
image file: d5em00524h-t8.tif(8)

However, due to the solubility of water in 1-octanol, eqn (7) and (8) can only be considered approximate. For an exact treatment, the partition coefficient of the charged species is also required.96 Due to the acidic or basic properties of the investigated compounds (see Table 1), reliable pKa values are needed for the neutral molecule and the corresponding ion to convert log[thin space (1/6-em)]KOW to log[thin space (1/6-em)]DOW for a specific pH value. For morphine, Abraham et al.99 reported a log[thin space (1/6-em)]DOW = 0.76 at a pH of 8.9. The value calculated using eqn (8) is 0.89 − 0.06 = 0.83. A commonly used reference value for environmental conditions is pH 7. The agreement between pKa obtained from OPERA and data from other sources varies considerably. For many substances, the agreement is good94,100–102 (see also SI), but deviations of approximately two orders of magnitude must be considered in some cases. Examples include tetrahydrocannabinol (7.80 compared to 10.60 (ref. 103)) and methadone (6.76 compared to 8.94 (ref. 104)). Differences are also observed in the zwitterion psilocybin. OPERA yields only one pKa value of 3.63, but the acidic phosphate group accounts for two dissociation constants, which Richter et al.105 calculate as 1.87 and 6.21. In addition, the dimethylamine group has basic character with a pKa of 9.24.105

3.3 Octanol/air partitioning

This partition coefficient can be used to describe the distribution of an airborne substance between the gas phase and the particle phase,106 or the distribution between the gas phase and settled house dust.107 At a particle concentration in the air of 20 μg m−3 molecules with log[thin space (1/6-em)]KOA ≤ 9 are completely in the gas phase, and those with log[thin space (1/6-em)]KOA ≥ 13 are completely in the particle phase. The curve is sigmoidal and shifts to a higher particle phase/gas phase ratio with higher particle concentrations. Especially in the range of the inflection point of the sigmoidal curve, small differences Δlog[thin space (1/6-em)]KOA are sufficient to significantly change the particle phase/gas phase ratio.18 Therefore, it is important to know the log[thin space (1/6-em)]KOA in this range as precisely as possible.

If log[thin space (1/6-em)]KOA is taken as a predictor for the partitioning of semi-volatile organic compounds (SVOCs) to aerosols,106 then with log[thin space (1/6-em)]KOA ≥ 12 more than 95% of the molecules should be in the particle phase, even at low particle concentrations.18 However, it is obvious from Fig. 3 that for PSY, MOR, THC, and LSD, the QM and OPERA values vary considerably (note that all OPERA values are computed), where QM tends to significantly higher log[thin space (1/6-em)]KOA values. The vapor pressure of the subcooled liquid PL can be used as a support here. According to the theory by Junge and Pankow,108 molecules with log[thin space (1/6-em)]PL (Pa) ≤ 6 are almost completely in the particle phase at a particle concentration of 20 μg m−3. Thus, it can be assumed that the four substances mentioned above are also essentially attached to particles. Especially for PSY, the difference between the QM methods and OPERA is more than 10[thin space (1/6-em)]log units. This could be attributed to the zwitterionic nature of psilocybin in 1-octanol. Unless the QSAR tool OPERA fully captures the true zwitterionic nature of PSY, the model will predict an erroneous partitioning between 1-octanol and air. This applies analogously to other binary systems. Also noteworthy is the large difference of almost 4[thin space (1/6-em)]log units for THC between COSMO-RS (14.52) and OPERA (10.82). In a previous calculation, we obtained 13.76,32 while Askari et al. calculated a value of 12.26.16 We thus believe the OPERA computed value is too small.

Generally, however, we find good agreement between all QM methods. Note that Parnis and Metcalfe109 calculated a log[thin space (1/6-em)]KOA = 9.94 for cocaine using COSMOtherm, which is very close to our value of 9.37. A notable deviation can be seen for N2O. The log[thin space (1/6-em)]KOA of N2O can be estimated from experimental data. For 1-octanol, Makranczy et al.86 report an Ostwald coefficient of 2.139 ml ml−1 at 298 K and p = 760 mmHg from which log[thin space (1/6-em)]KOA = +0.33 is obtained. The calculated values with COSMO-RS and openCOSMO-RS agree well. However, SMD predicts a too small value of −2.28. The reason is the same as stated in Section 3.1 – SMD predicts a too positive solvation free energy in 1-octanol, thus yielding a too small log[thin space (1/6-em)]KOA.

3.4 Air/water partitioning

The air/water partition coefficient is calculated from the Henry solubility HS according to eqn (9), with HS in mol (m−3 Pa−1), R = 8.314 J (mol−1 K−1) and T in K.
 
image file: d5em00524h-t9.tif(9)

The conversion unit factor from mol per (atm m3) to mol per (m3 Pa) is 9.87 × 10−6. In the case of air/water partitioning, similar prerequisites apply to the dissociation of molecules as in octanol/water. The conditions are even simpler, since completely undissociated molecules can be assumed in air. This allows the eqn (7) and (8) with DAW,acid = αKAW and DAW,base = (1 − α)KAW to be used analogously for the air/water system.

As with the other partition coefficients, hardly any experimental log[thin space (1/6-em)]KAW values are available for the molecules. For N2O, Sander85 calculated a Henry solubility of HS = 2.4 × 10−4 mol (m−3 Pa−1) from the data of Weiss and Price.110 Using eqn (9) this results in log[thin space (1/6-em)]KAW = 0.23, which is in good agreement both with the experimental value log[thin space (1/6-em)]KAW = 0.22 of Gabel and Schultz87 determined from the Ostwald coefficient and with the values calculated with COSMO-RS. Sander's85 database includes Henry solubilities for some other of the target substances, but these are not measured values. For cocaine, Parnis and Metcalfe109 calculated a log[thin space (1/6-em)]KAW = −6.59 using COSMOtherm and estimated log[thin space (1/6-em)]DAW values for pH 0 (−14.37), pH 4 (−10.38) and pH 7 (−7.44) with pKa = 7.78.

We generally observe good agreement between the QM models without any noticeable outliers. All OPERA values are computed and exhibit a significantly larger scatter, particularly for KHAT, FEN, LSD, and MEP, with deviations of about three orders of magnitude compared to COSMO-RS. Furthermore, OPERA predicts a log[thin space (1/6-em)]KAW for PSY that is ten orders of magnitude higher. Again, it is likely that OPERA does not account for the zwitterionic nature of PSY. This leads to a significantly higher PSY fraction in the gas phase compared to the aqueous phase. Most of the 23 substances in Table 1 show low water solubility. However, it should be noted that this study concerns the actual molecules. Free bases such as cocaine, methamphetamine, etc., are typically further processed into their water soluble hydrochlorides. In the illicit intravenous use of heroin, the free base is dissolved using a weak acid directly before injection.

3.5 Hexadecane/air partitioning

The hexadecane/air partition coefficient log[thin space (1/6-em)]KHdAL is useful for characterizing the nonspecific intermolecular interactions of organic chemicals in other partitioning processes and is used in its logarithmic form as a descriptor in LFERs. Due to the hydrophobic properties of n-hexadecane, molecular dissociation processes hardly need to be taken into account. In principle, log[thin space (1/6-em)]KHdA is a quantity that can be experimentally measured using gas chromatography on nonpolar capillary columns.111 However, the data available for the drug molecules listed in Table 1 are limited. An experimental value of log[thin space (1/6-em)]KHdA = 0.17 for N2O can be deduced from measurements of the Ostwald coefficient by Makranczy et al.112 In the UFZ-LSER database, values are published for 13 of the 23 compounds, and only for morphine there is a publication by Abraham et al.113 cited. The authors argue that the descriptors for morphine were generated from experimental values, but the origin of the data is not stated. For the other 12 substances, no literature source is given and it is therefore not clear whether their log[thin space (1/6-em)]KHdA values are measured or computed. The log[thin space (1/6-em)]KHdA (L) values of the other 10 substances were calculated from the SMILES structures as described above.

The COSMO-RS methods are in very good agreement, while SMD systematically underestimates the log[thin space (1/6-em)]KHdA values. Most values from the UFZ-LSER database agree well with COSMO-RS; however, outliers are observed for heroin, morphine and methadone with higher reported values compared to COSMO-RS. Particularly striking, however, is the large deviation of approximately 10[thin space (1/6-em)]log units for LSD (see Fig. 3). The observed deviation originates from the presence of a relevant tautomer that differs from the structure given by the SMILES string (shown in Fig. 4C). Our QM workflow identified multiple distinct gas-phase tautomers, three of which are shown in Fig. 4, with form A being the most stable. Form B lies about 5.8 kcal mol−1 higher than A, and form C about 12.4 kcal mol−1 higher. This tautomeric effect is effectively the nuclear relaxation contribution to the solvation free energy as defined in eqn (5), and is therefore fully captured by our QM workflow. QSPR methods, in contrast, do not account for tautomerization and the associated energetic changes, leading to substantial deviations in their computed values. This highlights the need for tautomer screening and how a robust QM workflow can yield qualitatively different, and more consistent results. Stenzel et al.111 compared experimentally determined log[thin space (1/6-em)]KHdA values for 387 environmentally relevant compounds with COSMOtherm data and found deviations of up to three orders of magnitude. Assuming that mainly computed values are compared for the drug molecules considered here, most deviations are within the expected range; only for heroin and LSD are there notable discrepancies of five and ten orders of magnitude, respectively.


image file: d5em00524h-f4.tif
Fig. 4 Relative gas-phase energies ΔE (kcal mol−1) of three LSD tautomers identified by our QM workflow. Form A is the most stable and set to 0.0 kcal mol−1, while forms B and C are 5.8 and 12.4 kcal mol−1 higher in energy, respectively. Hydrogens involved in tautomerization are highlighted in magenta. The SMILES-based tautomer is C.

3.6 Vapor pressure of the subcooled liquid

Among the vapor pressures obtained from OPERA, the CompTox database indicates that the values for amphetamine (AMP), heroin (HER), and cocaine (COC), are experimental. Verifiable experimental vapor pressure data at 298 K are available from original publications for 12 of the drug molecules considered here. There is agreement between OPERA and the results of Lawrence et al.114 for heroin and cocaine. However, Lawrence et al. extrapolated the vapor pressure of HER at 298 K from Antoine data, which were determined in the range 51 °C to 66 °C. The vapor pressure of cocaine was determined in the range between 21 °C and 41 °C.

The OPERA value for amphetamine roughly matches the value of Thornton et al.115 Since a reasonable experimental data set is available and to avoid confusion, we did not distinguish between experiment and prediction in the OPERA values. Except for PSY, the correlation of COSMO-RS with these experimental values and the OPERA values is very good (see Fig. 5) and within the range of our previous study.33 As expected, the results for COSMO-RS and openCOSMO-RS are almost identical. The highest deviations can be found for OPERA for MDMA and HER with about two orders of magnitude. In general, the quality of experimental data is difficult to assess. Meng et al.116 used an indirect gas chromatographic method to extrapolate the vapor pressures of drug molecules using di-n-butyl phthalate (DnBP) as a standard. The authors report log(PL/Pa) = −2.04 for the vapor pressure of DnBP at 298 K, whereas Gobble et al.117 give for the same temperature log(PL/Pa) = −2.35. The experimentally determined vapor pressures for HER,114 COC,114 AMP,115 FEN,118 METH,119 and PCP117 appear plausible. However, the value for LSD published by Okumus et al.120 is not directly verifiable. Uncertainty also exists for THC. COSMO-RS here yields log(PL/Pa) = −7.31 (see Table 2). A previous calculation without using explicit solution and gas phase optimized geometries (i.e., not including the nuclear relaxation) yielded −4.76. This agrees with the experimental value of Meng et al.116 Additionally, the log[thin space (1/6-em)]KOA = 12.27 calculated by Askari et al.,16 also using COSMO-RS, is also noticeably smaller than our calculated log[thin space (1/6-em)]KOA of 14.52 (see Table 2).


image file: d5em00524h-f5.tif
Fig. 5 Left: Scatter diagram of experimental and OPERA calculated vapor pressure data at 298 K plotted against COSMO-RS data. Right: Temperature-dependent vapor pressure curves of the 23 target molecules, calculated with COSMO-RS. Note that the critical temperature of nitrous oxide is 309.5 K.29

In our work, all reported vapor-pressure values include the explicit nuclear relaxation term – i.e., used the optimized geometry of each solute in both the gas and subcooled liquid phases. For comparison, values obtained without nuclear relaxation are provided in the SI, and in the cases of AMP, FEN, METH and PCP these results agree better with experimental vapor pressures. It is important to bear in mind, however, that these calculated quantum-chemical vapor pressure estimate relies on additional approximations – most notably the use of the subcooled liquid state to represent pure-liquid behavior – and that implicit solvation models (like COSMO-RS and openCOSMO-RS) are typically far more extensively parameterized for regular solution phase properties than for vapor pressures.52,121 Additionally, we approximate the continuum for the geometries of the subcooled liquid to be that one of 1-octanol, for the reasons stated in Section 2.2. Generally, we do not expect a noticeable change in the results when using geometries optimized in their respective subcooled liquid phase, as structural differences across dielectric constants in the typical range for drug-like compounds (εr ≈ 8–15) are small and the resulting effect on the calculated vapor pressures is expected to be minor compared to the intrinsic uncertainty of the solvation models themselves.33,52 The assessment of available publications has shown that some of the experimental vapor pressure determinations of drug molecules are questionable or at least have significant uncertainties. On the other hand, the results of QM calculations also depend on the model used. However, it has been demonstrated here and in previous work that COSMO-RS calculations are in most cases in good agreement with reliable experimental data. We are aware that this data set is limited. However, based on our earlier study,33 we assume that the uncertainty (standard deviation) of our vapor pressure calculations remains in the range of 0.5[thin space (1/6-em)]log units.

3.7 Temperature dependence of the partition coefficients and the vapor pressure

In general, the temperature of the environment often differs from the temperature given in the tabulated data. Therefore knowledge of the temperature dependence of the partition coefficients Kαβ and the vapor pressure PL is essential for many purposes. If both α and β are liquids, the temperature dependence of Kαβ is relatively small. Experimental findings around room-temperature indicate that as a rule of thumb10,122–127 one can use
 
image file: d5em00524h-t10.tif(10)

In the case of KOW both an increase as well as a decrease is observed in experiments.122 Our findings reveal a similar behavior with the same order of magnitude for dlog[thin space (1/6-em)]KOW/dT. In the case of METH, direct comparison can be made with the experimental findings of Brodin et al.127 (note that the entry of Brodin's result in Table 1 of Sangster's compilation122 is wrong). Brodin et al. report log[thin space (1/6-em)]KOW = 2.13 ± 0.67 at 298.15 K. Our findings of log[thin space (1/6-em)]KOW = 2.94 ± 0.50 are in fair agreement. The temperature dependence was observed to be127 dlog[thin space (1/6-em)]KOW/dT ≈ 0.01 K−1 whereas our calculations give ≈0.003 K−1. COSMO-RS is parameterized using the Gibbs free energy and does not truly account for enthalpy or entropy directly. Although the model includes some statistical mechanics, it can be limited and can underestimates the quantitative aspect of the temperature dependence.128–132 The situation is generally much more clear if one phase is a liquid (1-octanol, hexadecane, water) and the other phase is air. As the solubility of substances in a liquid generally decreases with increasing temperature,

 
image file: d5em00524h-t11.tif(11)
should result for α = liquid and β = air. This is also observed in our calculations. From a quantitative perspective, the question arises as to how partition coefficients and vapor pressures determined for a temperature T1 can be converted to a temperature T2. This is usually done using the van't Hoff equation, which assumes a constant phase transition enthalpy ΔHαβ in the temperature interval under consideration. LFER equations based on experimental data sets are often used to predict the required enthalpies.10,133,134 In contrast, QM methods allow the direct calculation of the solvation energy ΔGsolv at the desired temperature.

The calculations were performed over a temperature interval from 223 K to 333 K for all parameters. The results are fully summarized in the SI. The temperature dependence of log[thin space (1/6-em)]PL is shown in Fig. 5 for the interval from 283 K to 313 K. Apart from PSY, the largest range occurred for MOR with log[thin space (1/6-em)]PL(283 K) = −6.26 and log[thin space (1/6-em)]PL(313 K) = −5.51.

With reference to Fig. 6, we discuss the influence of the solvation energies ΔGsolv for 1-octanol and water on the partition parameters log[thin space (1/6-em)]KOW, log[thin space (1/6-em)]KOA and log[thin space (1/6-em)]KAW using COC as an example in the temperature range between 283 K and 313 K. We chose COC because, among the 23 target compounds, the relatively largest changes were observed here when neglecting PSY. As expected, ΔGsolv increases with temperature, with the effect being more pronounced for water than for 1-octanol. This then leads to the expected decrease in log[thin space (1/6-em)]KOA and increase in log[thin space (1/6-em)]KAW. The log[thin space (1/6-em)]KOW value also increases with increasing temperature. Cocaine is generally a good example to demonstrate the temperature effect on log[thin space (1/6-em)]KOW. A log coefficient ratio of ≈4 can be observed, with log[thin space (1/6-em)]KOW(283 K) = 0.38 and KOW(313 K) = 1.58. The ratio octanol/water is therefore 2.3 at 283 K and 38.0 at 313 K. At higher log[thin space (1/6-em)]KOW the absolute mass transfer of cocaine into the water phase with decreasing temperature is significantly lower.


image file: d5em00524h-f6.tif
Fig. 6 Left: Temperature dependence of the solvation energies ΔGsolv for cocaine in water and 1-octanol. Right: Temperature dependence of the cocaine partition coefficients log[thin space (1/6-em)]KOW, log[thin space (1/6-em)]KOA and log[thin space (1/6-em)]KAW. The calculations were performed using COSMO-RS.

After evaluating the available data, we consider QM calculations to be the best method to obtain reliable temperature-dependent partition coefficients and vapor pressures for the 23 substances. Phase transition enthalpies are only available for some of the experimentally determined vapor pressures. The same applies to the Abraham descriptors for LFER calculations. However, we also note that under ambient conditions, the differences in temperature-dependent partition behavior are small in most cases. Furthermore, a comprehensive analysis must take into account that pKa values are also temperature-dependent.135

3.8 Prediction of the partitioning behavior in the environment

The key question is whether the available data on partition coefficients and vapor pressure allow reasonable predictions about the fate of drug molecules in the environment. Assuming that the partition coefficients depend only on the free solvation energy ΔGsolv, eqn (12) for the relationship between KOW, KOA and KAW is obtained.
 
log[thin space (1/6-em)]KOA = log[thin space (1/6-em)]KOW − log[thin space (1/6-em)]KAW (12)

Eqn (12) is a simplified approach, which only serves to estimate the preferred accumulation in the hypothetical air/water/octanol system. Since there is a dependence on three parameters, a graphical representation of Mackay et al.136 is used. In Fig. 7, log[thin space (1/6-em)]KOW is plotted against log[thin space (1/6-em)]KAW, while log[thin space (1/6-em)]KOA is shown as diagonals.


image file: d5em00524h-f7.tif
Fig. 7 Plot of log[thin space (1/6-em)]KOW versus log[thin space (1/6-em)]KAW with log[thin space (1/6-em)]KOA as diagonals at 298 K for the 23 target molecules to estimate which substances preferentially accumulate in which medium (air, water, 1-octanol). The figure is based on space diagrams from a publication by Mackay et al.136

For all substances, the partitioning depends on the dimensions of the system, i.e., on the available amounts of air, water, and 1-octanol. Fig. 7 therefore only allows a general statement about which of the three media a substance would prefer. It is clear that the gas N2O will preferably accumulate in air. Psilocybin (PSY) will be preferentially present in the aqueous phase, which is evident from both the partition parameters and the pKa value. For the other 21 substances no educated guess can be made. Except for N2O, all substances have small log[thin space (1/6-em)]KAW values. The tendency toward accumulation in 1-octanol therefore increases with increasing log[thin space (1/6-em)]KOA and log[thin space (1/6-em)]KOW. Wania137 uses a similar diagram, but refers to environmental compartments instead of air, water, and 1-octanol. Accordingly, most of these 21 substances would accumulate in soils and sediments or distribute between the compartments as multimedia chemicals.

The Henderson–Hasselbalch eqn (7) and (8) state, that in aqueous media for pH < pKa, the molecule is predominantly present in the neutral form. Assuming pH 7, this is the case for 14 substances. Five substances have pKa values between 5 and 7, meaning that both the neutral and ionic species exist at pH 7. For MEQ and PSY with pKa = 2.23 and pKa = 3.63, the equilibrium is further shifted towards the ionic form in aqueous media. This means that for a more precise analysis of substances with pKa < 7 the calculation of log[thin space (1/6-em)]D and the partition coefficients of the ionic species would be necessary, but this is beyond the scope of this work. It must also be noted that more complex models are required to reliably estimate the bioaccumulation and partitioning of drugs in multiple environmental compartments.138

4 Conclusion

It is probably pointless to list the drugs available on the illegal market individually. The drug supply is evolving extremely rapidly, and new substances or derivatives of already known substances appear almost daily. The CIA World Factbook,26 for example, defines five categories of illicit drugs: narcotics, stimulants, depressants (sedatives), hallucinogens, and cannabis, in which parent molecules and their derivatives are grouped together. Based on drug consumption statistics, we selected 23 substances that cover the above categories and are structurally different. We included nitrous oxide in the list because it is sold in cans via vending machines and its physical properties are also of interest from a medical perspective.

Since only few experimental data on the partition coefficients of the molecules considered here are available, it is difficult to evaluate the results of the quantum mechanical calculations. However, our previous studies have already shown that quantum mechanics tends to more reliable results than QSAR methods.32,33 The agreement between log[thin space (1/6-em)]KOW, log[thin space (1/6-em)]KAW and log[thin space (1/6-em)]KOA determined using QSAR (OPERA) and QM (COSMO-RS) varies. With the exception of the zwitterion psilocybin (PSY), which always shows large deviations, various substances are conspicuous (see Fig. 3). There may be several reasons for this; one possibility could be the respective training sets of the QSAR method. In the case of log[thin space (1/6-em)]KHdA, heroin (HER), morphine (MOR) and LSD show significant discrepancies between LFER and QM. The consideration of tautomers in the calculation of the log[thin space (1/6-em)]KHdA of LSD demonstrates one of the strengths of a QM-based approach.

For the vapor pressure, a comparison with experimental data is possible, but it is obvious from Fig. 5 that COSMO-RS underestimates the experimental values at low vapor pressures. The temperature dependence of partition coefficients and vapor pressure is important from a practical point of view. As shown for cocaine (COC) in Fig. 6, QM methods allow for direct calculations for a specific temperature. However, Grimme et al.34 state that COSMO-RS can only consider the temperature dependence of the free solvation energy semiquantitatively. Alternatively, the van't Hoff equation and the Clausius–Clapeyron equation can be used if the respective enthalpy of phase transfer is known.

Many of the 23 molecules are acids or bases. This makes the log[thin space (1/6-em)]KOW values particularly uncertain, since these are determined for the free molecule, but in the aqueous phase, including the aqueous portion of the octanol phase, the pKa and pH dependent dissociation must be taken into account. Correction formulas for log[thin space (1/6-em)]D are at best approximate. log[thin space (1/6-em)]KOA, log[thin space (1/6-em)]KAW and log[thin space (1/6-em)]KHdA are less influenced by the acid/base properties.

Even in light of all experimental and theoretical uncertainties, the results communicated and discussed in this work contribute to a significantly improved understanding of the physical properties of drug molecules. It is possible to plausibly estimate whether certain substances preferentially accumulate in air, water, or an organic medium (1-octanol). Furthermore, valuable information is obtained on their distribution behavior between the gas and particle phase in aerosols139 and on accumulation in house dust.107

Conflicts of interest

The authors have no conflicts of interest to declare.

Data availability

All data supporting this article have been included as part of the supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5em00524h.

Acknowledgements

L. W. would like to express his deepest gratitude to Prof. Stefan Grimme for his exceptional support and for granting access to the computational resources at the Mulliken Center for Theoretical Chemistry. L. W. also greatly acknowledges support of the Stiftung Stipendien-Fonds des Verbandes der Chemischen Industrie e.V. through its Kekulé Fellowship program and the granted access to the Marvin cluster hosted by the University of Bonn. T. S. thanks Manuela Lingnau (WK) for her support in preparing the graphic abstract. All authors thank Professor Stefan Grimme for fruitful discussions.

References

  1. F. Carod-Artal, Hallucinogenic drugs in pre-Columbian Mesoamerican cultures, Neurologia, 2015, 30, 42–49 CrossRef PubMed.
  2. W. Sneader, Drug Discovery – A History, John Wiley & Sons Ltd, Chichester, UK, 2005 Search PubMed.
  3. R. Gerona, Designer Drugs, Elsevier Inc., Amsterdam, 2024 Search PubMed.
  4. P. Armenian, K. T. Vo, J. Barr-Walker and K. L. Lynch, Fentanyl, fentanyl analogs and novel synthetic opioids: A comprehensive review, Neuropharmacology, 2018, 134, 121–132 CrossRef PubMed.
  5. E. A. Glasper, Nitrous oxide deaths among children and young people is no laughing matter, Compr. Child Adolesc. Nurs., 2023, 46, 259–261 CrossRef PubMed.
  6. C. Ort, et al., Spatial differences and temporal changes in illicit drug use in Europe quantified by wastewater analysis, Addiction, 2014, 109, 1338–1352 CrossRef PubMed.
  7. B. Petrie, R. Barden and B. Kasprzyk-Hordern, A review on emerging contaminants in wastewaters and the environment: Current knowledge, understudied areas and recommendations for future monitoring, Water Res., 2015, 72, 3–27 CrossRef PubMed.
  8. Y.-C. Lin, T.-C. Hsiao and A. Y.-C. Lin, Urban wastewater treatment plants as a potential source of ketamine and methamphetamine emissions to air, Water Res., 2020, 172, 115495 CrossRef PubMed.
  9. C. de Dios-Perez, A. Ballesteros-Gomez, J. Ballesteros and S. Rubio, Supramolecular solvent extraction of drugs and pharmaceuticals in indoor dust from public buildings, Sci. Total Environ., 2025, 968, 178859 CrossRef PubMed.
  10. R. P. Schwarzenbach, P. M. Gschwend and D. M. Imboden, Environmental Organic Chemistry, John Wiley & Sons, Hoboken, 2017 Search PubMed.
  11. N. Mastroianni, C. Postigo, M. Lopez de Alda, M. Viana, A. Rodriguez, A. Alastuey, X. Querol and D. Barcelo, Comprehensive monitoring of the occurrence of 22 drugs of abuse and transformation products in airborne particulate matter in the city of Barcelona, Sci. Total Environ., 2015, 532, 344–352 CrossRef PubMed.
  12. M. Viana, X. Querol, A. Alastuey, C. Postigo, M. Lopez de Alda, D. Barcelo and B. Artinano, Drugs of abuse in airborne particulates in urban environments, Environ. Int., 2010, 36, 527–534 CrossRef PubMed.
  13. G. Morrison, N. V. Shakila and K. Parker, Accumulation of gas-phase methamphetamine on clothing, toy fabrics, and skin oil, Indoor Air, 2015, 25, 405–414 CrossRef PubMed.
  14. A. Cecinato, C. Balducci and M. Perilli, Illicit psychotropic substances in the air: The state-of-art, Sci. Total Environ., 2016, 539, 1–6 CrossRef PubMed.
  15. A. Cecinato, P. Romagnoli, M. Perilli and C. Balducci, Psychotropic substances in house dusts: a preliminary assessment, Environ. Sci. Pollut. Res., 2017, 24, 21256–21261 CrossRef PubMed.
  16. A. Askari, F. Wania and A. W. H. Chan, Modeling the fate and involuntary exposure to tetrahydrocannabinol emitted from indoor cannabis smoking, Environ. Sci.: Atmos., 2023, 3, 760–772 CAS.
  17. T. Salthammer, The legalization of cannabis may result in increased indoor exposure to Δ9-tetrahydrocannabinol, J. Hazard. Mater., 2024, 464, 132949 CrossRef CAS PubMed.
  18. T. Salthammer, Assessment of methods for predicting physical and chemical properties of organic compounds, Indoor Environ., 2024, 1, 100031 CrossRef.
  19. W. Sippl and D. Robaa, QSAR/QSPR, in Applied Chemoinformatics: Achievements and Future Opportunities, ed. T. Engel and J. Gasteiger, WILEY-VCH, 2018, pp. 9–52 Search PubMed.
  20. I. V. Tetko, A. Yan and J. Gasteiger, Prediction of Physicochemical Properties of Compounds, in Applied Chemoinformatics: Achievements and Future Opportunities, ed. T. Engel and J. Gasteiger, Wiley-VCH, Weinheim, 2018, pp. 53–81 Search PubMed.
  21. Q. Zang, K. Mansouri, A. J. Williams, R. S. Judson, D. G. Allen, W. M. Casey and N. C. Kleinstreuer, In silico prediction of physicochemical properties of environmental chemicals using molecular fingerprints and machine learning, J. Chem. Inf. Model., 2017, 57, 36–49 CrossRef CAS PubMed.
  22. K. Mansouri, N. F. Cariello, A. Korotcov, V. Tkachenko, C. M. Grulke, C. S. Sprankle, D. Allen, W. M. Casey, N. C. Kleinstreuer and A. J. Williams, Open-source QSAR models for pKa prediction using multiple machine learning approaches, J. Cheminf., 2019, 11, 60 Search PubMed.
  23. X. Kang, B. Hu, M. C. Perdana, Y. Zhao and Z. Chen, Extreme learning machine models for predicting the n-octanol/water partition coefficient (Kow) data of organic compounds, J. Environ. Chem. Eng., 2022, 10, 108552 CrossRef CAS.
  24. P. Schossler, T. Schripp, T. Salthammer and M. Bahadir, Beyond phthalates: Gas phase concentrations and modeled gas/particle distribution of modern plasticizers, Sci. Total Environ., 2011, 409, 4031–4038 CAS.
  25. Bundeskriminalamt, Rauschgiftkriminalität – Bundeslagebild 2023, Bundeskriminalamt, Wiesbaden, 2024 Search PubMed.
  26. U.S Central Intelligence Agency, The World Factbook – Illicit Drugs, US Central Intelligence Agency – Office of Public Affairs, Washington D.C., 2025 Search PubMed.
  27. National Institute on Drug Abuse, Commonly Used Drugs Charts, National Institutes of Health, Bethesda, MD, 2023 Search PubMed.
  28. K. Mansouri, C. M. Grulke, R. S. Judson and A. J. Williams, OPERA models for predicting physicochemical properties and environmental fate endpoints, J. Cheminf., 2018, 10, 10 Search PubMed.
  29. J. R. Rumble, T. J. Bruno and M. J. Doa, Handbook of Chemistry and Physics – 102st Edition, CRC Press, Boca Raton, 2021 Search PubMed.
  30. A. Ben-Naim, Solvation Thermodynamics, Springer Science – Business Media, New York, 1987 Search PubMed.
  31. M. F. Vitha and P. W. Carr, The chemical meaning of the standard free energy of transfer: Use of van der Waals’ equation of state to unravel the interplay between free volume, volume entropy, and the role of standard states, J. Phys. Chem. B, 2000, 104, 5343–5349 CrossRef CAS.
  32. T. Salthammer, S. Grimme, M. Stahn, U. Hohm and W.-U. Palm, Quantum chemical calculation and evaluation of partition coefficients for classical and emerging environmentally relevant organic compounds, Environ. Sci. Technol., 2022, 56, 379–391 CrossRef CAS PubMed.
  33. M. Stahn, S. Grimme, T. Salthammer, U. Hohm and W.-U. Palm, Quantum chemical calculation of the vapor pressure of volatile and semi volatile organic compounds, Environ. Sci.:Processes Impacts, 2022, 24, 2153–2166 RSC.
  34. S. Grimme, F. Bohle, A. Hansen, P. Pracht, S. Spicher and M. Stahn, Efficient quantum chemical calculation of structure ensembles and free energies for nonrigid molecules, J. Phys. Chem. A, 2021, 125, 4039–4054 CrossRef CAS PubMed.
  35. A. A. Maryott, Circular of the Bureau of Standards No. 514: Table of Dielectric Constants of Pure Liquids, National Bureau of Standards, Gaithersburg, MD, 1951 Search PubMed.
  36. Z. Tang, C. Chang, F. Bao, L. Tian, H. Liu, M. Wang, C. Zhu and J. Xu, Feasibility of predicting static dielectric constants of polymer materials: A density functional theory method, Polymers, 2021, 13, 284 CrossRef CAS PubMed.
  37. M. Jorge and L. Lue, The dielectric constant: Reconciling simulation and experiment, J. Chem. Phys., 2019, 150, 084108 CrossRef PubMed.
  38. R. M. Redheffer, R. C. Wildman and V. O'Gorman, The computation of dielectric constants, J. Appl. Phys., 1952, 23, 505–508 CrossRef CAS.
  39. S. Spicher and S. Grimme, Efficient computation of free energy contributions for association reactions of large molecules, J. Phys. Chem. Lett., 2020, 11, 6606–6611 CrossRef CAS PubMed.
  40. M. Bursch, J. Mewes, A. Hansen and S. Grimme, Best-practice DFT protocols for basic molecular computational chemistry, Angew. Chem., Int. Ed., 2022, 61, e202205735 CrossRef CAS PubMed.
  41. J. Gorges, S. Grimme, A. Hansen and P. Pracht, Towards understanding solvation effects on the conformational entropy of non-rigid molecules, Phys. Chem. Chem. Phys., 2022, 24, 12249–12259 RSC.
  42. A. J. Garza, Solvation entropy made simple, J. Chem. Theory Comput., 2019, 15, 3204–3214 CrossRef CAS PubMed.
  43. IUPAC, Solvation Energy (ST07102), https://goldbook.iupac.org/terms/view/ST07102 Search PubMed.
  44. A. V. Marenich, C. J. Cramer and D. G. Truhlar, Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions, J. Phys. Chem. B, 2009, 113, 6378–6396 CrossRef CAS PubMed.
  45. J. M. Herbert, Dielectric continuum methods for quantum chemistry, WIREs Comput. Mol. Sci., 2021, 11, e1519 CrossRef CAS.
  46. A. V. Marenich, C. J. Cramer and D. G. Truhlar, Perspective on foundations of solvation modeling: The electrostatic contribution to the free energy of solvation, J. Chem. Theory Comput., 2008, 4, 877–887 CrossRef CAS PubMed.
  47. V. Radtke, et al., A unified pH scale for all solvents: part I – intention and reasoning (IUPAC Technical Report), Pure Appl. Chem., 2021, 93, 1049–1060 CrossRef CAS.
  48. S. Ehlert, M. Stahn, S. Spicher and S. Grimme, Robust and efficient implicit solvation model for fast semiempirical methods, J. Chem. Theory Comput., 2021, 17, 4250–4261 CrossRef CAS PubMed.
  49. A. Klamt and G. Schüürmann, COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient, J. Chem. Soc., Perkin Trans. 2, 1993, 799–805 RSC.
  50. V. Barone and M. Cossi, Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model, J. Phys. Chem. A, 1998, 102, 1995–2001 CrossRef CAS.
  51. A. Klamt, Conductor-like screening model for real solvents: A new approach to the quantitative calculation of solvation phenomena, J. Phys. Chem., 1995, 99, 2224–2235 CrossRef CAS.
  52. T. Gerlach, S. Müller, A. G. De Castilla and I. Smirnova, An open source COSMO-RS implementation and parameterization supporting the efficient implementation of multiple segment descriptors, Fluid Phase Equilib., 2022, 560, 113472 CrossRef CAS.
  53. P. Pracht, C. A. Bauer and S. Grimme, Automated and efficient quantum chemical determination and energetic ranking of molecular protonation sites, J. Comput. Chem., 2017, 38, 2618–2631 CrossRef CAS PubMed.
  54. P. Pracht, S. Grimme, C. Bannwarth, F. Bohle, S. Ehlert, G. Feldmann, J. Gorges, M. Müller, T. Neudecker, C. Plett, S. Spicher, P. Steinbach, P. A. Wesołowski and F. Zeller, CREST—A program for the exploration of low-energy molecular chemical space, J. Chem. Phys., 2024, 160, 114110 CrossRef CAS PubMed.
  55. F. Bohle, J. Seibert and S. Grimme, Automated quantum chemistry-based calculation of optical rotation for large flexible molecules, J. Org. Chem., 2021, 86, 15522–15531 CrossRef CAS PubMed.
  56. A. Katbashev, M. Stahn, T. Rose, V. Alizadeh, M. Friede, C. Plett, P. Steinbach and S. Ehlert, Overview on building blocks and applications of efficient and robust extended tight binding, J. Phys. Chem. A, 2025, 129, 2667–2682 CrossRef CAS PubMed.
  57. F. Neese, Software Update: The ORCA Program System—Version 6.0, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2025, 15, e70019 Search PubMed.
  58. N. M. O'Boyle, M. Banck, C. A. James, C. Morley, T. Vandermeersch and G. R. Hutchison, Open Babel: An open chemical toolbox, J. Cheminf., 2011, 3, 33 Search PubMed.
  59. P. Pracht, F. Bohle and S. Grimme, Automated exploration of the low-energy chemical space with fast quantum chemical methods, Phys. Chem. Chem. Phys., 2020, 22, 7169–7192 RSC.
  60. N. van Staalduinen and C. Bannwarth, MolBar: a molecular identifier for inorganic and organic molecules with full support of stereoisomerism, Digital Discovery, 2024, 3, 2298–2319 RSC.
  61. C. Bannwarth, S. Ehlert and S. Grimme, GFN2-xTB - An accurate and broadly parametrized self-consistent tight-binding quantum chemical method with multipole electrostatics and density-dependent dispersion contributions, J. Chem. Theory Comput., 2019, 15, 1652–1671 CrossRef CAS PubMed.
  62. S. Grimme, A. Hansen, S. Ehlert and J.-M. Mewes, r2SCAN-3c: A “Swiss army knife” composite electronic-structure method, J. Chem. Phys., 2021, 154, 064103 CrossRef CAS PubMed.
  63. J. W. Furness, A. D. Kaplan, J. Ning, J. P. Perdew and J. Sun, Accurate and numerically efficient r2SCAN meta-generalized gradient approximation, J. Phys. Chem. Lett., 2020, 11, 8208–8215 CrossRef CAS PubMed.
  64. J. W. Furness, A. D. Kaplan, J. Ning, J. P. Perdew and J. Sun, Correction to: Accurate and numerically efficient r2SCAN meta-generalized gradient approximation, J. Phys. Chem. Lett., 2020, 2020, 9248 CrossRef PubMed.
  65. L. Wittmann, H. Neugebauer, S. Grimme and M. Bursch, Dispersion-corrected r2SCAN based double-hybrid functionals, J. Chem. Phys., 2023, 159, 224103 CrossRef CAS PubMed.
  66. E. Caldeweyher, S. Ehlert, A. Hansen, H. Neugebauer, S. Spicher, C. Bannwarth and S. Grimme, A generally applicable atomic-charge dependent London dispersion correction, J. Chem. Phys., 2019, 150, 154122 CrossRef PubMed.
  67. D. Rappoport and F. Furche, Property-optimized Gaussian basis sets for molecular response calculations, J. Chem. Phys., 2010, 133, 134105 CrossRef PubMed.
  68. K. A. Peterson, D. Figgen, E. Goll, H. Stoll and M. Dolg, Systematically convergent basis sets with relativistic pseudopotentials. II. Small-core pseudopotentials and correlation consistent basis sets for the post-d group 16–18 elements, J. Chem. Phys., 2003, 119, 11113–11123 CrossRef CAS.
  69. D. Andrae, U. Häußermann, M. Dolg, H. Stoll and H. Preuß, Energy-adjusted ab initio pseudopotentials for the second and third row transition elements, Theor. Chim. Acta, 1990, 77, 123–141 CrossRef CAS.
  70. G. L. Stoychev, A. A. Auer and F. Neese, Automatic generation of auxiliary basis sets, J. Chem. Theory Comput., 2017, 13, 554–562 CrossRef CAS PubMed.
  71. F. Neese, F. Wennmohs, A. Hansen and U. Becker, Efficient, approximate and parallel Hartree–Fock and hybrid DFT calculations. A ‘chain-of-spheres’ algorithm for the Hartree–Fock exchange, Chem. Phys., 2009, 356, 98–109 CrossRef CAS.
  72. R. Izsák and F. Neese, An overlap fitted chain of spheres exchange method, J. Chem. Phys., 2011, 135, 144105 CrossRef PubMed.
  73. B. Helmich-Paris, B. De Souza, F. Neese and R. Izsák, An improved chain of spheres for exchange algorithm, J. Chem. Phys., 2021, 155, 104109 CrossRef CAS PubMed.
  74. A. Klamt, The COSMO and COSMO-RS solvation models, WIREs Comput. Mol. Sci., 2018, 8, e1338 CrossRef.
  75. S. G. Balasubramani, et al., TURBOMOLE: Modular program suite for ab initio quantum-chemical and condensed-matter simulations, J. Chem. Phys., 2020, 152, 184107 CrossRef PubMed.
  76. J. P. Perdew and W. Yue, Accurate and simple density functional for the electronic exchange energy: Generalized gradient approximation, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8800–8802 CrossRef PubMed.
  77. F. Weigend and R. Ahlrichs, Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  78. M. Garcia-Ratés and F. Neese, Effect of the solute cavity on the solvation energy and its derivatives within the framework of the Gaussian charge scheme, J. Comput. Chem., 2020, 41, 922–939 CrossRef PubMed.
  79. K. Mansouri, C. M. Grulke, A. M. Richard, R. S. Judson and A. J. Williams, An automated curation procedure for addressing chemical errors and inconsistencies in public datasets used in QSAR modelling, SAR QSAR Environ. Res., 2016, 27, 911–937 CrossRef PubMed.
  80. K. Mansouri, J. T. Moreira-Filho, C. N. Lowe, N. Charest, T. Martin, V. Tkachenko, R. Judson, M. Conway, N. C. Kleinstreuer and A. J. Williams, Free and open-source QSAR-ready workflow for automated standardization of chemical structures in support of QSAR modeling, J. Cheminf., 2024, 16, 19 Search PubMed.
  81. A. J. Williams, C. M. Grulke, J. Edwards, A. D. McEachran, K. Mansouri, N. C. Baker, G. Patlewicz, I. Shah, J. F. Wambaugh, R. S. Judson and A. M. Richard, The CompTox Chemistry Dashboard: a community data resource for environmental chemistry, J. Cheminf., 2017, 9, 61 Search PubMed.
  82. C. Hansch, A. Leo and D. Hoeckman, Exploring QSAR – Hydrophobic, Electronic and Steric Constants, American Chemical Society, Washington D.C., 1995 Search PubMed.
  83. N. Ulrich, S. Endo, T. Brown, N. Watanabe, G. Bronner, M. H. Abraham and K.-U. Goss, UFZ-LSER Database V 3.2.1, Helmholtz Centre for Environmental Research, Leipzig, 2017 Search PubMed.
  84. T. Brown, Predicting hexadecane–air equilibrium partition coefficients (L) using a group contribution approach constructed from high quality data, SAR QSAR Environ. Res., 2014, 25, 51–71 CrossRef PubMed.
  85. R. Sander, Compilation of Henry's law constants (version 5.0.0) for water as solvent, Atmos. Chem. Phys., 2023, 23, 10901–12440 CrossRef.
  86. J. Makranczy, L. Rusz and K. Balog-Megyery, Solubility of gases in normal alcohols, Hung. J. Ind. Chem., 1979, 7, 41–46 Search PubMed.
  87. R. A. Gabel and B. Schultz, Solubility of nitrous oxide in water, 20-80 C, Anesthesiology, 1973, 38, 75–81 CrossRef PubMed.
  88. C. Hansch, A. Vittoria, C. Silipo and P. Y. C. Jow, Partition coefficients and the structure-activity relationship of the anesthetic gases, J. Med. Chem., 1975, 18, 546–548 CrossRef CAS PubMed.
  89. L. Paul, C. T. Namba-Nzanguim, A. Telesphory, J. Oppong Mensah, D. Mteremko, R. Costa, S. M. Katundu, L. P. Kwiyukwa, N. D. Kambaine, J. Juvenary, S. Mlowe, G. Deogratias, D. M. Shadrack and A. S. Paluch, Investigation of the structure, stability, and relative solubility of psilocybin in water and pure organic solvents: A molecular simulation study, J. Mol. Liq., 2023, 392, 123479 CrossRef CAS.
  90. M. Amezcua, L. El Khoury and D. Mobley, SAMPL7 Host-Guest Challenge Overview: Assessing the reliability of polarizable and non-polarizable methods for binding free energy calculations (Version 2), ChemRxiv, 2020, preprint,  DOI:10.26434/chemrxiv.12768353.v2.
  91. C. C. Zanith and J. R. Pliego, Performance of the SMD and SM8 models for predicting solvation free energy of neutral solutes in methanol, dimethyl sulfoxide and acetonitrile, J. Comput.-Aided Mol. Des., 2015, 29, 217–224 CrossRef CAS PubMed.
  92. R. Sure and S. Grimme, Comprehensive benchmark of association (free) energies of realistic host-guest complexes, J. Chem. Theory Comput., 2015, 11, 3785–3801 CrossRef CAS PubMed.
  93. A. Savoy, E. Paenurk, R. Pollice, P. H. Hünenberger and P. Chen, Solvation free energies of ion dissociations in dichloromethane: En route to accurate computations, J. Phys. Chem. B, 2025, 129, 6276–6288 CrossRef CAS PubMed.
  94. A. Avdeef, D. A. Barrett, P. N. Shaw, R. D. Knaggs and S. S. Davis, Octanol–, chloroform–, and propylene glycol dipelargonat–water partitioning of morphine–6–glucuronide and other related opiates, J. Med. Chem., 1996, 39, 4377–4381 CrossRef CAS PubMed.
  95. J. Sangster, Octanol-water Partition Coefficients: Fundamentals and Physical Chemistry, John Wiley & Sons Ltd, Chichester, 1997 Search PubMed.
  96. M. Kah and C. D. Brown, Log D: Lipophilicity for ionisable compounds, Chemosphere, 2008, 72, 1401–1408 CrossRef CAS PubMed.
  97. C. Hansch and A. Leo, Exploring QSAR – Fundamentals and Applications in Chemistry and Biology, American Chemical Society, Washington D.C., 1995 Search PubMed.
  98. R. A. Scherrer and S. M. Howard, Use of distribution coefficients in quantitative structure-activity relations, J. Med. Chem., 1977, 20, 53–58 CrossRef CAS PubMed.
  99. M. H. Abraham, K. Takacs-Novak and R. C. Mitchell, On the partition of ampholytes: application to blood–brain distribution, J. Pharm. Sci., 1997, 86, 310–315 CrossRef CAS PubMed.
  100. E. B. Leffler, H. M. Spencer and A. Burger, Dissociation constants of adrenergic amines, J. Am. Chem. Soc., 1951, 73, 2611–2613 CrossRef.
  101. M. Rosa Boleda, M. Huerta-Fontela, F. Ventura and M. T. Galceran, Evaluation of the presence of drugs of abuse in tap waters, Chemosphere, 2011, 84, 1601–1607 CrossRef CAS PubMed.
  102. L. Settimo, K. Bellman and R. M. A. Knegtel, Comparison of the accuracy of experimental and predicted pKa values of basic and acidic compounds, Pharm. Res., 2014, 31, 1082–1095 CrossRef CAS PubMed.
  103. E. R. Garrett and C. Hunt, Physicochemical properties, solubility, and protein binding of Δ9 -tetrahydrocannabinol, J. Pharm. Sci., 1974, 63, 1056–1064 CrossRef CAS PubMed.
  104. A. H. Beckett, Analgesics and their antagonists: Some steric and chemical considerations, J. Pharm. Pharmacol., 1956, 8, 848–859 CrossRef CAS PubMed.
  105. L. M. Richter, J. Al-Gousous, G. L. B. de Araujo, N. M. Davies and R. Löbenberg, Assessing the utility of in silico tools in early drug development: The case of a pharmaceutically relevant formulation of the prodrug psilocybin, J. Drug Delivery Sci. Technol., 2024, 92, 105305 CrossRef CAS.
  106. A. Finizio, D. Mackay, T. Bidleman and T. Harner, Octanol-air partition coefficient as a predictor of partitioning of semi-volatile organic chemicals to aerosols, Atmos. Environ., 1997, 31, 2289–2296 CrossRef CAS.
  107. C. J. Weschler and W. W. Nazaroff, SVOC partitioning between the gas phase and settled dust indoors, Atmos. Environ., 2010, 44, 3609–3620 CrossRef CAS.
  108. J. F. Pankow, Review and comparative analysis of the theories on partitioning between the gas and aerosol particulate phases in the atmosphere, Atmos. Environ., 1987, 21, 2275–2283 CrossRef CAS.
  109. J. M. Parnis and C. D. Metcalfe, The potential for atmospheric transport of contaminants to remote lakes: An assessment using COSMO-RS solvation theory, Chemosphere, 2025, 385, 144555 CrossRef CAS PubMed.
  110. R. Weiss and B. Price, Nitrous oxide solubility in water and seawater, Mar. Chem., 1980, 8, 347–359 CrossRef CAS.
  111. A. Stenzel, S. Endo and K.-U. Goss, Measurements and predictions of hexadecane/air partition coefficients for 387 environmentally relevant compounds, J. Chromatogr. A, 2012, 1220, 132–142 CrossRef CAS PubMed.
  112. J. Makranczy, K. Megyery-Balog, L. Rusz and L. Patyi, Solubility of gases in normal-alkanes, Hung. J. Ind. Chem., 1976, 4, 269–280 CAS.
  113. M. H. Abraham, R. Sanchez-Moreno, J. Gil-Lostes, W. E. Acree, J. Enrique Cometto-Muniz and W. S. Cain, The biological and toxicological activity of gases and vapors, Toxicol. in Vitro, 2010, 24, 357–362 CrossRef CAS PubMed.
  114. A. H. Lawrence, L. Elias and M. Authier-Martin, Determination of amphetamine, cocaine, and heroin vapour pressures using a dynamic gas blending system and gas chromatographic analysis, Can. J. Chem., 1984, 62, 1886–1888 CrossRef CAS.
  115. M. Thornton, J. Chickos, I. V. Garist, M. A. Varfolomeev, A. A. Svetlov and S. P. Verevkin, The vaporization enthalpy and vapor pressure of (d)-amphetamine and of several primary amines used as standards at T/K = 298 as evaluated by correlation gas chromatography and transpiration, J. Chem. Eng. Data, 2013, 58, 2018–2027 CrossRef CAS.
  116. Y. Meng, A. H. Lichtmann, D. T. Bridgen and B. R. Martin, Inhalation studies with drugs of abuse, NIDA Res. Monogr., 1997, 173, 201–224 CAS.
  117. C. Gobble, B. Walker and J. S. Chickos, The vaporization enthalpy and vapor pressure of fenpropidin and phencyclidine (PCP) at T/K = 298.15 by correlation gas chromatography, J. Chem. Eng. Data, 2016, 61, 896–902 CrossRef CAS.
  118. P. K. Gupta, K. Ganesan, P. K. Gutch, L. Manral and D. K. Dubey, Vapor pressure and enthalpy of vaporization of fentanyl, J. Chem. Eng. Data, 2008, 53, 841–845 CrossRef CAS.
  119. M. Thornton, C. Gobble and J. Chickos, The vaporization enthalpy and vapor pressure of S(+)-methamphetamine at T = 298.15 K by correlation gas chromatography, J. Chem. Thermodyn., 2014, 73, 51–56 CrossRef CAS.
  120. B. Okumus, A. Metin and I. A. Kariper, Understanding the effects of lysergic acid diethylamide and the importance of its prevention, Health Sci. Rev., 2023, 8, 100107 Search PubMed.
  121. A. Klamt, V. Jonas, T. Bürger and J. C. W. Lohrenz, Refinement and parametrization of COSMO-RS, J. Phys. Chem. A, 1998, 102, 5074–5085 CrossRef CAS.
  122. J. Sangster, Octanol-water partition coefficients of simple organic compounds, J. Phys. Chem. Ref. Data, 1989, 18, 1111–1229 CrossRef CAS.
  123. N. P. Bahadur, W.-Y. Shiu, D. G. B. Boocock and D. Mackay, Temperature dependence of octanol water partition coefficient for selected chlorobenzenes, J. Chem. Eng. Data, 1997, 42, 685–688 CrossRef CAS.
  124. A. Leo, C. Hansch and D. Elkins, Partition coefficients and their uses, Chem. Rev., 1971, 71, 525–616 CrossRef CAS.
  125. Y. D. Lei, S. Baskaran and F. Wania, Measuring the octan-1-ol air partition coefficient of volatile organic chemicals with the variable phase ratio headspace technique, J. Chem. Eng. Data, 2019, 64, 4793–4800 CrossRef CAS.
  126. S. Sarraute, H. Delepine, M. F. Costa Gomez and V. Majer, Aqueous solubility, Henry's law constants and air/water partition coefficients of n-octane and two halogenated octanes, Chemosphere, 2004, 57, 1543–1551 CrossRef CAS PubMed.
  127. A. Brodin, B. Sandin and B. Failerson, Rates of transfer of organic protolytic solutes between an aqueous and an organic phase V. The thermodynamics of mass transfer, Acta Pharm. Suec., 1976, 13, 331–352 CAS.
  128. J. M. Parnis, D. Mackay and T. Harner, Temperature dependence of Henry’s law constants and KOA for simple and heteroatom-substituted PAHs by COSMO-RS, Atmos. Environ., 2015, 110, 27–35 CrossRef CAS.
  129. Y. S. Sistla and A. Khanna, Validation and prediction of the temperature-dependent Henry’s constant for CO2–ionic liquid systems using the Conductor-like Screening Model for Realistic Solvation (COSMO-RS), J. Chem. Eng. Data, 2011, 56, 4045–4060 CrossRef CAS.
  130. A. Klamt, Prediction of the mutual solubilities of hydrocarbons and water with COSMO-RS, Fluid Phase Equilib., 2003, 206, 223–235 CrossRef CAS.
  131. M. Diedenhofen and A. Klamt, COSMO-RS as a tool for property prediction of IL mixtures - A review, Fluid Phase Equilib., 2010, 294, 31–38 CrossRef CAS.
  132. D. Zaitsau, M. Trawny, J. Arndt, L. Wittmann, C. Held and R. Siewert, Critical evaluation of amino acids enthalpies of formation and phase transitions with a new approach based on solution calorimetry, quantum cluster growth, COSMO-RS and PC-SAFT, J. Chem. Eng. Data, 2025 Search PubMed , submitted.
  133. S. Baskaran and F. Wania, Applications of the octanol–air partitioning ratio: a critical review, Environ. Sci.: Atmos., 2023, 3, 1045–1065 CAS.
  134. S. Baskaran, A. Podagatlapalli, A. Sangion and F. Wania, Predicting the temperature dependence of the octanol–air partition ratio: a new model for estimating ΔUoOA, J. Solution Chem., 2023, 52, 51–69 CrossRef CAS.
  135. R. L. Thurlkill, D. A. Cross, J. M. Scholtz and C. N. Pace, pKa of fentanyl varies with temperature: implications for acid-base management during extremes of body temperature, J. Cardiothorac. Vasc. Anesth., 2005, 19, 759–762 CrossRef CAS PubMed.
  136. D. Mackay, A. K. Celsie and J. M. Parnis, The evolution and future of environmental partition coefficients, Environ. Rev., 2016, 24, 101–113 CrossRef.
  137. F. Wania, Assessing the potential of persistent organic chemicals for long-range transport and accumulation in polar regions, Environ. Sci. Technol., 2003, 37, 1344–1351 CrossRef.
  138. J. M. Parnis and D. Mackay, Multimedia Environmental Models – the Fugacity Approach, CRC Press, Boca Raton, FL, 2021 Search PubMed.
  139. B. J. Finlayson-Pitts and J. N. Pitts, Chemistry of the Upper and Lower Atmosphere, Academic Press, San Diego, 2000 Search PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.