Next Article in Journal
Highly Sensitive Ethanol Sensing Using NiO Hollow Spheres Synthesized via Hydrothermal Method
Next Article in Special Issue
Simultaneous Analysis of MLH1, MSH2, MSH6, PMS2 and KRAS in Patients with Gastric and Colon Cancer Using Stochastic Sensors
Previous Article in Journal
MIP-Based Screen-Printed Potentiometric Cell for Atrazine Sensing
Previous Article in Special Issue
How Meaningful Are Minor Details in the Generation of Nanomodified Electrochemical Enzyme Biosensors? Exploring the Scenario with Sinusoidal Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Liquid Chromatography on the Different Methods for the Determination of Lipophilicity: An Essential Analytical Tool in Medicinal Chemistry

by
José X. Soares
1,2,3,†,
Álvaro Santos
1,†,
Carla Fernandes
1,2,* and
Madalena M. M. Pinto
1,2
1
Laboratório de Química Orgânica e Farmacêutica, Departamento de Ciências Químicas, Faculdade de Farmácia, Universidade do Porto, Rua Jorge Viterbo Ferreira, 228, 4050-313 Porto, Portugal
2
Centro Interdisciplinar de Investigação Marinha e Ambiental (CIIMAR), Universidade do Porto, Terminal de Cruzeiros do Porto de Leixões, Avenida General Norton de Matos, s/n, 4450-208 Matosinhos, Portugal
3
LAQV-REQUIMTE, Laboratory of Applied Chemistry, Department of Chemical Sciences, Faculty of Pharmacy, University of Porto, Rua de Jorge Viterbo Ferreira, 228, 4050-313 Porto, Portugal
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Chemosensors 2022, 10(8), 340; https://doi.org/10.3390/chemosensors10080340
Submission received: 15 July 2022 / Revised: 14 August 2022 / Accepted: 14 August 2022 / Published: 18 August 2022
(This article belongs to the Collection Women Special Issue in Chemosensors and Analytical Chemistry)

Abstract

:
Lipophilicity is one of many parameters involved in the biological activity of drugs, as it affects their pharmacokinetic and pharmacodynamic behavior. Generally, lipophilicity is assessed by the partition coefficient of a compound between a nonpolar phase (n-octanol) and an aqueous phase (water), expressed as P (partition coefficient) or as its decimal logarithm (Log P). The gold standard method for the experimental determination of Log P is the shake-flask method. In this context, chromatographic methods enable the direct and simple quantification of the partitioned compound between the two phases. This review discusses the use of liquid chromatography (LC) for direct and indirect determination of lipophilicity. Beyond the classical isotropic log P determination, methods for assessing anisotropic lipophilicity are also reviewed. Several examples are discussed that highlight the versatility of LC technique and current trends. The last section of this review focuses on a case study describing an experience of our group and emphasizing the dual role of LC in determining Log P.

Graphical Abstract

1. Lipophilicity and Its Importance in Drug Discovery and Design

Lipophilicity, as defined by the International Union of Pure and Applied Chemistry (IUPAC), is a “physicochemical property which describes a partitioning equilibrium of solute molecules between water and an immiscible organic solvent, favouring the latter” [1]. Commonly, lipophilicity is expressed as the logarithm to the base 10 of the partition coefficient (log P) of a solute [2]:
log P = log [ s u b s t a n c e ] nonaqueous   [ s u b s t a n c e ] aqueous
where [ s u b s t a n c e ] nonaqueous is the solute equilibrium concentration in a nonaqueous (nonpolar phase) and [ s u b s t a n c e ] aqueous is its equilibrium concentration in an aqueous phase (polar phase). Classically, the aqueous phase is water, but the partition coefficient can also be determined using buffer solutions. In this case, the partitioning involving co-existing electrical species of a solute at a given pH is expressed as the distribution coefficient log D:
log D pH = log i n ( f i × P i )
where P i is the partition coefficient of the electrical species i and f i , its molar fraction [3].
Classically, the nonaqueous phase is an isotropic organic solvent, such as n-octanol. However, artificial and natural membranes, such as liposomes and micelles (Figure 1a), have been recently used as alternative nonaqueous phases [4]. These systems are anisotropic because they establish different topographical relations between the solute and the nonaqueous phase involving different interaction forces [3]. The choice of the nonaqueous phase impacts the type of intermolecular forces encoded in the partition coefficient value. Lipophilicity is the product of all intermolecular forces involved in the partition of a solute between the two phases, assembling not only the contributions of hydrophobicity, the tendency of apolar groups or molecules to associate in an aqueous environment, but also the contributions of polarity [5]. The different intermolecular forces encoded in lipophilicity are shown in Figure 1, and they are different depending on the nonaqueous phase. While isotropic lipophilicity results from the net sum of hydrophobicity minus polarity, anisotropic lipophilicity also considers the ionic bonds [3]. Therefore, isotropic and anisotropic partition coefficients express lipophilicity on different scales.
Lipophilicity is a key parameter on drug discovery and design. Every single pharmacokinetic component, namely the absorption, distribution, metabolism, excretion, and toxicity (ADMET), is modulated by lipophilicity. Absorption is related to the lipophilicity because drugs typically must cross biomembrane portals to reach the sites of action. Passive permeation into or across biomembranes is modulated by the partition of the drug into the lipid membrane and from the membrane into the aqueous intracellular environment [6]. The rate-limiting step for hydrophilic compounds is the partitioning into the hydrophobic membrane layer, and for hydrophobic compounds is partitioning into the hydrophilic intracellular medium [7]. Solubility is another parameter that is modulated by lipophilicity in such a way that solubility can be calculated by an empirical formula that correlates melting point and log P with solubility [2]. In general, compounds with lower log P tend to be water-soluble. In addition to solubility, other physical parameters are related to lipophilicity, such as the flexibility of a molecule based on the presence of rotatable bonds and on the ratio of sp3 carbons, the presence of polar groups, and the presence of intramolecular hydrogen bonding [8].
Drug distribution is also affected by lipophilicity. Lipophilic compounds have a large volume of distribution due to affinity to lipophilic tissues such as adipose tissue [2]. Nevertheless, lipophilicity also indirectly influences the distribution of drugs through plasma protein binding, as lipophilic compounds tend to have high plasma protein binding [9,10]. The facility to cross the compact and protected blood–brain barrier (BBB) is also dependent on lipophilicity [6]. In general, small and highly lipophilic compounds can penetrate easier across the BBB, and increasing lipophilicity is likely to enhance BBB permeation [11]. However, increasing lipophilicity can also be associated with higher binding affinity to efflux pumps, such as P-glycoprotein [2]. Lipophilicity is also related to metabolic rate as in vivo metabolism often converts lipophilic drugs into more polar metabolites to facilitate their clearance. In addition, lipophilic compounds are more likely to be toxic [12]. One example is the interaction with the human ether-a-go-go related gene (hERG), especially for positively charged lipophilic compounds, which can cause increases in the cardiac QT interval [13].
Despite being mostly recognized for its role in pharmacokinetics, lipophilicity also plays a crucial role in the pharmacodynamic behaviour of a drug. Lipophilicity contributes to understanding ligand–target interactions, considering the different forces that govern the intermolecular interactions involved in biochemical and pharmacological processes. In fact, lipophilicity is a parameter considered in numerous quantitative structure–activity relationship studies [5,14]. The presence of lipophilic moiety offers lead compounds with a higher probability of having anchors on protein targets with large and hydrophobic binding sites [15]. However, increasing lipophilicity to the pursuit of potency through increasing lipophilicity across the drug development comes with costs. One is the trend, known as ‘molecular obesity’, for seeking huge and lipophilic molecules that, despite being highly potent, do not have a suitable pharmacokinetic profile [7]. Another risk is target promiscuity. Compounds that achieve their potency through the increase of lipophilicity are more likely to be less selective, increasing the risk of off-target interactions and undesirable toxicity [16].
In a nutshell, lipophilicity is probably the single most informative physicochemical parameter in modern drug discovery. Therefore, drug design seeks to control this property within a defined optimal range [15]. To fulfil this aim, suitable methods for the determination of lipophilicity are required mainly in the early steps of the drug discovery pipeline.

2. Methods for Determination of Lipophilicity

The determination of lipophilicity has become a routine process for medicinal chemists [17]. Lipophilicity can be measured or can be calculated [2]. The experimental methods to determine lipophilicity can be divided into two categories: direct and indirect methods [18,19]. In direct experimental methods, the partition coefficient is obtained directly from the concentration ratio in the equilibrium of a compound partitioned between the nonaqueous and aqueous phases. In indirect experimental methods, the partition coefficient is estimated through a correlation, usually the correlation between a compound retention factor in a reverse-phase chromatographic system and its partition coefficient. Liquid chromatography (LC) plays a relevant role both in the direct and indirect determination of lipophilicity.

2.1. In Silico Determination of Lipophilicity

Since the pivotal work of Hansch et al., which found out that the partition coefficients have an additive constitutive character, a myriad of in silico methodologies have been developed to predict lipophilicity from the molecular structure of a molecule [20]. If some of these methods are very simple [21], others exist that are more sophisticated and based on deep neural networks [22]. Excellent reviews on the in silico estimation of isotropic [23] and anisotropic lipophilicity [24,25] are available in the literature.
Independent of the computational method, the calculation is always based on a set of experimentally obtained values [26]. Therefore, the calculated values should be regarded as approximations. Often, calculated log P values are inaccurate, and the reliability of calculation methods is low for highly complex compounds [27]. In addition, in silico methods can also be imprecise as computed values for a class of compounds could vary up to two log P units depending on the approach used [28,29]. For the above-mentioned reasons, calculated values should be considered when [30]:
  • Choosing the experimental method;
  • Selecting conditions for a LC analysis;
  • Examining the plausibility of experimentally obtained values;
  • Providing an estimate when experimental methods are not applicable.
In the early discovery process, in silico methods are very useful for filtering drug-like compounds [31]. However, as soon as possible, the predicted values should be substituted by more accurate measured values [28].

2.2. Direct Experimental Determination of Lipophilicity

Direct determination of lipophilicity requires the quantification of the compound concentration present in the nonaqueous and in the aqueous phase. The direct method known as the shake-flask method is the gold standard for log P determination. Direct methods are usually accurate when measuring log P values in the range −2 to 4, but they are labor-consuming and usually require relatively large amounts of pure compounds [32].
In direct methods, LC is the method typically used to quantify the amount of substance in each phase of the biphasic system. When compared to other analytical methods, LC provides a wider range of applicability than gas chromatography and provides a lower detection limit than UV/Vis spectroscopy. The detection limit is particularly relevant for highly lipophilic compounds because compounds with log P values larger than 4 are often limited by the analyte minimum detection limit in the aqueous phase [19]. The use of LC is similar in the different direct methods. Therefore, in the next subchapters the different direct methods that can be coupled with LC will be discussed, providing more focus on the biphasic system rather than on the LC conditions.

2.2.1. Shake-Flask Method

The shake-flask method uses n-octanol (hydrophobic) and water (hydrophilic) as the biphasic liquid system, and it is the standard method recommended by Organization for Economic Co-operation and Development (OECD) [32]. This classical procedure consists of dissolving the sample, shaking until equilibrium has been reached, and measuring the compound concentration in each phase of the biphasic system [32].
The shake-flask method is a direct measurement of the partition coefficient, and that is its main advantage [33]. However, this procedure has several drawbacks. It requires control over a large number of experimental parameters (different experimental conditions usually produce different values of log P for the same analyte) [34]. It is a highly time-consuming method as the partition needs to reach equilibrium. The time needed to reach equilibrium concentrations varies hugely (from 1 to 24 h), and partitioning rates may be fast or slow depending on the log P of the solute and the degree of shaking [35]. It consumes large amounts of solvent per analyte [35], and it is not suitable for degradable compounds and surface-active materials [27]. Finally, it is not appropriate for poorly soluble compounds [35].
Modification of the shake-flask method has been proposed that enables the determination using different alternatives to the n-octanol/water system aimed to imitate different physiological cell barriers (e.g.: chloroform/water, alkane/water, or propylene glycol dipelargonate/water) [36].

2.2.2. Slow-Stirring Method

In the slow-stirring method, n-octanol and aqueous phases are mixed under slow stirring (instead of shaking) [37]. By proceeding this way, the formation of emulsions can be prevented. The obtained values with the shake-flask and slow-stirring method are very similar for compounds exhibiting log P < 4.5, but they differ for compounds with higher log P, which can be attributed to the formation of octanol emulsions in the shake-flask procedure [38]. As a drawback, this method requires a long period of stirring, up to 2–3 days, to reach equilibrium [37].
The slow-stirring method can also be applicable for the determination of anisotropic lipophilicity using a liposome/buffer system. The difference between the procedures is in the quantification step, as liposomes are separated from the buffer phase by inducing liposome aggregation followed by filtration [39].

2.2.3. Water-Plug Aspiration/Injection Method

The water-plug aspiration/injection method is another shake-flask modification particularly suitable for highly lipophilic compounds. In the shake-flask procedure, it is difficult to manually separate the water phase from the n-octanol phase without contaminating the water phase because of the high viscosity of n-octanol. Improper phase separation can impact the determination with a higher impact for highly lipophilicity compounds [40]. To avoid this problem, a small volume of water is aspirated into the needle before collecting the aqueous phase in order to decrease the potential contamination, as water in the needle should repel contaminating n-octanol phase [40]. The sampling of the water phase can be automated, but the obtained values must be manually inspected to identify instrument problems or solubility issues. The need for manual inspection can be suppressed by the use of an injection marker [41].

2.2.4. Flow-Based Method

Flow injection analysis can be exploited to set a variant of the shake-flask method that allows standardization of the measurement [42,43]. In this method, the test substance is dissolved in either n-octanol or water phase. The substance is then injected into the flow of a capillary tube with the corresponding phase, and the two immiscible phases are continuously pumped to a mixing point in the system. The substance partitions between the two phases, and after equilibration, a fraction of the aqueous flow is separated, and the concentration of the analyte is determined. The large surface area and short distance result in the fast transport of the analyte from one phase to the other, allowing it to rapidly reach equilibrium [42].

2.2.5. Miniaturization of Shake-Flask Method

To increase the throughput of log P measurements and to decrease the amounts of compound needed, the traditional shake-flask method was transposed to a 96-well format. In this technique, the partition coefficient of a solute is measured between a polymer phase and an aqueous phase [44]. The polymer phase is prepared in 96-well microplates. The tested compound is added to a polymer phase prepared in a 96-well microplate, and after incubation, in a shaker, the amount of compound is determined in the supernatant.

2.2.6. Vortex Liquid–Liquid Microextraction (VALLME) Method

VALLME is a microextraction method that, coupled with LC, aims to increase the throughput of log P measurements. In VALLME method, microvolumes of n-octanol are dispersed in the aqueous phase using vortex agitation [35]. The mechanical stress promoted by vortex agitation bursts the n-octanol phase into several smaller droplets forming an emulsion (Figure 2). The fine microdroplets formed to increase the interfacial contact area and reduce the thickness of the stagnant aqueous film, which is usually present at the n-octanol/water interface [35]. Consequently, the time required to reach equilibrium is dramatically shortened (equilibrium conditions are achieved in 2 min of vortex agitation) [35]. As the formed dispersion is unstable, the target analyte can be easily separated using centrifugation [45]. The microdroplet is then collected with a microsyringe, and the solute concentration is determined by LC [35].
The log P values are calculated as the ratio of equilibrium concentrations in the n-octanol and water phases. The equilibrium concentration of the analyte in the n-octanol phase is calculated each time using calibration curves obtained with the octanolic individual stock standard solutions. The log P values were calculated for each concentration level using the mass balance Equation (3) [46]:
log P = log   ( vol . water vol . water × [ octanol eq . ] ×   vol . octanol [ water i ] × vol . water [ octanol eq . ] ×   vol . octanol )
where vol . octanol is the volume of n-octanol phase, vol . water is the volume of the water phase, [octanoleq.] is the concentration of the compound in the n-octanol phase calculated externally by the calibration curve, and [ water i ] is the initial concentration in the aqueous phase. The main advantages of VALLME are its reliability, simplicity, low cost, minimal solvent, and solute consumption [35]. The main disadvantage of VALLME is the requirement of a low-density organic solvent as a nonaqueous phase. From a practical point of view, only n-octanol is suitable to be used in VALLME, as it is the only one that can restore to its initial single-drop shape after centrifugation [47].

2.2.7. Nano-Absorbent Based Method

Nanoparticles with absorbent properties can be used for the determination of partition coefficient. One of these methods is based on the use of porous silica-encapsulated magnetic nanoparticles, which are preloaded with a known amount of n-octanol [48,49]. These nanoparticles are then dispersed into an aqueous phase containing the tested compound. The small size of the n-octanol droplets and the efficient mixing provided by the nanoparticles increase the interfacial contact area between the two phases, which significantly reduces the time required to achieve equilibrium. The magnetic properties of the nanoparticles allow the easy separation of phases. The partition coefficient is determined by measuring the concentration of the targeted analyte in the aqueous phase before and after partitioning.
Despite being used for isotropic lipophilicity determination, nano-absorbent-based methods acquire particular relevance for the determination of anisotropic lipophilicity. Nano-sized artificial membranes, like liposomes [50] or micelles [51], act as absorbents enabling the direct determination of the lipophilicity in these biomimetic nonaqueous phase. The partitioning occurs just by mixing the artificial membranes with an aqueous solution of the compound to be tested. The major technical difficulty is the phase separation after reaching equilibrium. While n-octanol and water are two immiscible solvents, liposomes and micelles form a colloidal suspension in water. The most used techniques for phase separation are ultrafiltration [50] and ultracentrifugation [51]. The partition coefficient is determined by quantifying the compound present in the filtrate or in the supernatant.
An alternative method for the determination of anisotropic lipophilicity is based on solid-supported lipid membranes (SSLMs). SSLMs are silica nanobeads coated with a liposomal membrane that was noncovalently bound to the bead [52]. SSLMs beads are added to an aqueous solution of the compound to be tested, and the mixture is incubated to allow the partition between the lipid and aqueous phase. Phase separation is performed by centrifugation with a regular bench centrifuge [52] or by a short filtration step [53]. SSLM silica beads are commercially available and offer a high throughput as the time-consuming ultracentrifugation or ultrafiltration steps are not required.

2.2.8. Dialysis-Based Method

Dialysis bags can be used for the determination of isotropic lipophilicity [54]. In this method, the dialysis bag is filled with a solution of the compound under investigation dissolved in n-octanol saturated with water or buffer. The dialysis bag is then immersed into the aqueous phase saturated with n-octanol, and sonication is applied to shorten the equilibration time. After sonication, the phases are easily separated and analyzed by LC.
Similar to the nano-absorbent-based methods, dialysis-based methods are more important for the determination of anisotropic lipophilicity than for the determination of isotropic lipophilicity. In fact, the “gold standard” to measure lipid-water partitioning is the equilibrium dialysis method [52]. The dialysis cell consisted of two glass chambers separated by a dialysis membrane [55]. For the experiment, two dialysis cells are required: a reference cell and a measurement cell. One chamber of each cell is filled with a solution of the compound under is added, but the other chamber is filled with buffer and with the membrane suspension in the reference and measurement cell, respectively. The dialysis membrane allows the diffusion across the two chambers of the free compound but impedes the diffusion of the compound bound to the liposomes. The partition coefficient is determined by LC quantification of the compound concentration in the chambers without membrane suspension in the reference and in the measurement cells [56].

2.3. Indirect Experimental Determination of Lipophilicity

Due to the disadvantages associated with direct experiments, there is an increasing demand for developing methods that are able to estimate lipophilicity without quantification requirements [34]. These methods provide a greener and high-throughput measurements than can direct methods while simultaneously can reach good accuracy and versatility [19].
Since the 1980s, chromatographic procedures have been used to obtain physicochemical parameters associated with structural characteristics, namely log P [57]. Exploiting the relationship between the compound retention in a hydrophobic stationary phase and the compound lipophilicity, LC methods really shine in the indirect measurement of lipophilicity. In the following subchapters, the different LC methods used for indirect experimental determination of lipophilicity will be discussed.

2.3.1. Reversed-Phase Thin-Layer Chromatography

Reversed-phase thin-layer chromatography (RP-TLC) is the easiest LC method that can be used for the indirect determination of isotropic lipophilicity [58]. In RP-TLC [59] and reversed-phase-high performance thin layer chromatography (RP-HPTLC) [60], the stationary phase is a commercially available C8 or C18 silica gel plate (Figure 3). Alternatively, it is possible to prepare a non-polar stationary phase from the commonly available silica gel plates after its pulverization with n-octanol 5% (v/v) in diethyl ether. The mobile phases consist of binary mixtures of water (or buffer) and an organic solvent (modifier), usually methanol, which is the preferred solvent due to its compatibility with water, or alternatively tetrahydrofuran or acetonitrile [27].
The retention in reversed-phase systems is mainly governed by partitioning between the stationary and mobile phases [61,62]. Therefore, the lipophilicity index measured in RP-TLC can be derived from the RM value [61,62]:
RM = log [ ( 1 Rf ) 1 ]
where Rf is the retention factor which is obtained by dividing the distance covered by the sample by the distance covered by the mobile phase.
RM values are determined in the presence of different organic solvent concentrations in the mobile phase. The linear relationship between the RM values and different mobile phase proportions is established, and the partition coefficient is calculated by extrapolating to a pure water mobile phase [63].

2.3.2. Reversed-Phase High-Performance Liquid Chromatography

Reversed-phase high-performance liquid chromatography (RP-HPLC) is the most widely used indirect method to experimentally determine the lipophilicity [58,64] and the standard procedure for measuring log P is officially recommended by OECD [30]. The wide use of RP-HPLC is justified by the advances in the elucidation of solute interactions, the availability of well-characterized stationary phases, and the ubiquitous presence of automated HPLC systems [2]. Table 1 provides a list of reported RP-HPLC methods to illustrate the versatility and large applicability of this technique, including for each example the aim of the study, the compounds under analysis, and the chromatographic method and conditions used. In addition, Table 1 also describes other indirect methods coupled with LC, which will be discussed in the following subchapters. The bibliographic research of assembling Table 1 was conducted using Web of Science and Scopus without any temporal restriction. The keywords used were “lipophilicity AND chromatography”, “lipophilicity AND chromatography”, “partition AND reversed-phase high-pressure liquid chromatography, “micellar liquid chromatography”, and “immobilized artificial membrane”.
As an indirect measurement of lipophilicity, RP-HPLC does not directly provide a partition coefficient but rather a lipophilicity index which is correlated with the partition coefficient. The lipophilicity index is inferred by the assumption that compounds are retained in proportion to their nonaqueous-aqueous partition coefficient, with hydrophilic compounds eluted first and lipophilic compounds last, and it is expressed by the retention factor k [30]. The retention factor k is easily calculated using the retention time ( t r ), according to Equation (5) [27,30]:
log k = log ( t r t 0 t 0 )
where t 0 is the dead retention time (the retention time of an unretained solute) and t r is the retention time of the solute. The major factors that influence the retention time are the nature, length, and diameter of the stationary phase and the composition and flow rate of the mobile phase [57]. Compounds that are either unretained or too retained may have to be assessed using different column lengths or eluent compositions [77].
Under suitable isocratic conditions, the log P of the compound under study is obtained by measuring its retention factor and then inputting the k values into the Collander equation:
log P = a   × log k + b
where a and b are regression constants for the slope and intercept obtained by linear regression of the log P of reference substances against the log k of the reference substances. In these conditions, the k-value of a given solute and stationary phase depends on the composition of the mobile phase [77]. To address this problem, pure aqueous eluent (log kw) should be used as it is not affected by the organic modifier and reflects the partitioning behavior that occurred in a shake-flask measurement [77,85]. As under most conditions pure water cannot be used as mobile phase, log k w   is obtained by extrapolation to 100% water [19,78,114]. The Snyder–Soczewinski equation (Equation (7)) is the most common approach to estimate log kw [19]:
log k = log k w S φ
where φ is the volume fraction of the organic modifier in the mobile phase and S is a constant derived by linear regression analysis. The extrapolation for log kw is performed by determining the log k values using different percentages of organic solvent in the mobile phases (at least four isocratic log k values) and then extrapolating to 100% of pure water eluent [19,81,204]. Triplicates should be made [30] and they should fall within a range of ±0.1 log units [30].
The classical method for the determination of log P by RP-HPLC employs isocratic conditions. For compounds with high retention factors, the determination may be challenging due to band broadening. Although it requires a more complex mathematical treatment, elution in gradient mode can be used for the determination of log P and offers faster analysis time and resolution with constant peak width [166]. In gradient methods, the lipophilicity index is expressed by a different chromatographic parameter—the chromatographic hydrophobicity index (CHI). CHI corresponds to the percentage of organic modifier required to achieve equal distribution of the analyte between the stationary and the mobile phase. CHI values are obtained by measuring the retention times in gradient mode (tg) and then inputting the tg into the equation [19]:
CHI = a   ×   t g + b
where a and b are regression constants for the slope and intercept obtained by linear regression of t g of reference substances against the CHI of the reference substances.
For measuring log P with RP-HPLC, the following features need to be considered:
  • Precision: triplicates should be performed, and the obtained values of the log P values should be within a range of ±0.1 log units [26];
  • Sensitivity: HPLC enables log P to be determined over a range of about 0–6 [30];
  • Specificity: the presence of impurities may difficult the peak assignment and the interpretation of the obtained values [26];
  • Accuracy: the obtained log P values can be within ±1 log unit of the shake-flask value [26].
To control the above-mentioned features, the choice of suitable stationary and mobile phases plays a pivotal role.
Various stationary phases have been applied to evaluate indirectly isotropic lipophilicity [34]. Most of the chromatographic methods for measuring lipophilicity use RP silica-based stationary phases containing long-chain hydrocarbon, like octyl (C8) and octadecyl (C18), bound onto silica (Figure 3) [19,26,30,81].
Silica is the most preferred packing material for RP columns, but it has the inconvenience of possible silanophilic interactions between the analyte and the remaining free silanol sites of the stationary phase. These interactions, more prominent in positively charged compounds, can produce distortions in peak shape and a considerable increase in the compound retention, which affects their log P determination [19,27]. In addition, these stationary phases are not stable at pH above 7.5 and can even break down [115]. OECD recommends that the stationary phases should have a minimal percentage of polar groups as they may impair the performance of the RP-HPLC column [30]. However, to overcome this problem, the best recommendation is to use highly end-capped columns for the RP-HPLC method, especially for basic compounds under high pH conditions [77,166]. One of the strategies for end-capping is to use base-deactivated silica stationary phases where the end-capping of the silanol residues is performed by the treatment with trimethylchlorosilane [86]. Another strategy to suppress silanophilic effects is the use of polar embedded stationary phases. Here, electrostatic shielding is achieved by the incorporation of functional groups, such as amide or carbamate, at the alky chains [86]. LC-ABZ+ and Discovery-RP-Amide-C16 are two examples of polar embedded stationary phases which have been used for lipophilicity determination [19,86].
However, other secondary interactions between these polar groups and certain analytes may occur. Many of these end-capped columns are still limited to pH below 8 because of their instability in an alkaline medium. Masking the silica units with hydrophobic ethylene groups prevents them from dissolution, and the structural ethylene bridges increase the column stability at a wide pH range (1–12) [19,166]. Acquity BEH Shield RP18 and XBridge Shield RP18 column are two examples of ethyl bridged columns that can be used to measure the lipophilicity not only for neutral and acidic compounds but also for basic ones [19,129].
Polymer-based stationary phases, composed of a rigid and porous cross-linked polymer, are another type of phases aimed at suppressing silanophilic effects. Polystyrene-divinylbenzene (PS-DVB) and octadecyl-poly(vinyl alcohol) (ODP) are two examples of these types of stationary phases that have been successfully used for log P determination. Silanophilic effects are suppressed because they do not have any polar site, and they can be used at a wide range of pH values (pH 1–13) [19,95,205]. However, these columns often have a longer retention time when compared to the silica-based ones [115] and the retention mechanism may be controlled by different structural properties of the polymer [19].
Regarding the mobile phase, the most extensively used are binary mixtures of water or buffer with an organic modifier [27]. HPLC-grade modifier and distilled water should be used to prepare the eluting solvent, which is degassed prior to use [30]. The use of binary mixtures is a requirement as the use of only water in RP silica-based columns is not feasible [61]. With high percentages of water, more than 98%, C18 and C8 columns exhibit poorly reproducible retention due to ligand collapse and the decreased contact between the mobile phase and the stationary phase [206]. For the lipophilicity assessment, methanol is considered the most suitable organic modifier. Being a protic solvent, methanol can interact with free silanol groups suppressing this way the silanophilic effects [166]. In addition, methanol forms a monolayer that mimics the n-octanol/water interface of the shake-flask method [27]. For lipophilicity determination, the volume fraction of methanol should be above 25–30% [30]. At a high-volume fraction of methanol, the eluent mixture is so unlike water; the phase aimed to be mimicked by the mobile phase, that the correlation between retention times and lipophilicity will be poor [77]. Acetonitrile is another organic modifier used for lipophilicity determination. When acetonitrile is used as an organic modifier instead of methanol, the correlation between retention times and organic modifier percentage in the mobile phase is generally better described by a quadratic model instead of a linear model [71,166]. In addition, the correlation between log P and log kW was better when methanol was used instead of acetonitrile [71].
Additives to the organic modifier–water mixture can be used, particularly when dealing with dissociable compounds [19]. Hydrophobic amines, such as triethylamine, n-decylamine, and N,N-dimethyloctylamine, are the traditional examples of additives added to the mobile phase. These masking agents can reduce silanophilic interactions and are very useful when basic compounds are analyzed [19]. Ionic liquids, such as 1-butyl-3-methylimidazolium tetrafluoroborate ([BMIM][BF4]), are greener alternatives to the traditional additives [19,72]. These additives are usually composed of relatively large asymmetric organic cations and inorganic/organic anions, meaning that both anion and cation can participate in the retention mechanism [19]. Therefore, the inclusion of a masking agent can selectively affect the compound retention times, and it should only be used on justifiable occasions [77].
In a similar way, ion suppressors can be added to the mobile phase to suppress the dissociation of ionizable compounds. Ion suppression is achieved by using buffer solutions based on phosphate, phosphate-buffered saline (PBS), or morphilinepropanesulfonic acid (MOPS) [19]. n-Octanol itself can be used as an additive to provide a better correlation with lipophilicity. The idea behind it is to render octanol-like characteristics to the column by saturating the column with n-octanol and then use of n-octanol saturated water as eluent. However, the inclusion of n-octanol as an additive should be tested. Some reports show better correlations between log P and log kW if better if n-octanol is supplementarily added [72,77] while others show the opposite trend [174].
Regarding the detection, most of the reported LC methods use UV-Vis or diode-array detectors (DAD), while mass spectrometry (MS) detection is only scarcely used (Table 1). This is attributed to the easier use and ubiquitous presence of UV-Vis detection and to the fact the majority of the compounds under study (drugs or drug candidates) tend to absorb in this range.
Strictly related to RP-HPLC, solvatochromic analysis is an approach based on linear solvation–energy relationships (LSER) that allow for the understanding of the contribution of each intermolecular interaction involved in the partition [166]. The model considers both solute properties, such as Van der Waals volume or dipolarity/polarizability or hydrogen-bond–donor acidity, as well as the properties of the chromatographic system properties, which are calculated by multiple linear regression analysis for a diverse group of compounds [107,167,205]. The establishment of a robust correlation with the nonaqueous/aqueous partition coefficients involves not only a correlation between the RP-HPLC retention factor and the log P values but also LSER equations that are in agreement with the structural parameters [115].
When compared to the “gold standard” shake-flask method, RP-HPLC offers several advantages: it requires a smaller amount of sample and solvent, it has higher reproducibility, and it is faster as only the determination of retention times is required, and retention times are independent of the injected compound concentration/amount [19,30,81]. When compared to the RP-HPTLC, RP-HPLC tends to provide similar results [60,165].

2.3.3. Ultra-Performance Liquid Chromatography

Ultra-performance liquid chromatography (UPLC) works with columns packed with sub-2 μm porous particles. Compared to the conventional LC, the drastic reduction in the column length allows a faster determination of lipophilicity with a significant improvement in chromatographic performance and efficiency [166]. Few reports of lipophilicity determination using UPLC have been reported [166,167,168,169,170] because this technique requires specific equipment.
The different stationary phases will yield different partition coefficient values. If the RP-phases are similar, the analyte chromatographic behavior might be different, but the relationship between log P and retention factors tends to be similar [71]. However, if the distinct RP phases are used, the results are not comparable. To demonstrate this, Giaginis et al. [174] evaluated the lipophilicity of a set of thiazolidinediones derivatives with different RP phases, including one that is endcapped, polar-embedded, and polar-endcapped. For all compounds under study, the one polar endcapped phase provided the lowest retention factors, and the endcapped provided the highest retention factors. When compared with the partition coefficient values obtained by the shake-flask method, the partition coefficient values obtained with the endcapped phase provided the best correlation, namely in the pH range 5–8.

2.3.4. Counter-Current Chromatography

Counter-current chromatography (CCC) uses two immiscible solvents: one will act as a mobile phase while the other will act as a liquid stationary phase [204]. The liquid stationary phase is retained by applying a gravitational or a centrifugal field [43]. The partition coefficient is determined by taking into account the retention volume of the substance and the volume of the stationary and mobile phases [205]. The advantages of this technique are the use of small sample sizes and insensitivity to impurities; however, it is only applicable for log D < 4.3, and it is time- and labor-consuming [43].

2.3.5. Immobilized Artificial Membrane Chromatography

Immobilized artificial membrane (IAM) chromatography is the most widely used method for the determination of anisotropic lipophilicity [207]. Table 1 describes numerous examples of IAM chromatography applied to lipophilicity determination. The motivation behind IAM stationary phase is to better mimic specific interactions that solutes or drugs can establish with the membrane phospholipidic bilayer. To fulfil this goal, IAM stationary phases are prepared by binding covalently phospholipids to a propylamino functionalized silica particles (Figure 4a) [27,208]. The incorporated phospholipid can have a single or a double chain, but a double chain can better simulate biomembranes [27]. The remaining free propylamine residues on the surface of the silica backbone should be treated to suppress the undesired silanophilic interactions discussed above [27]. The most recent generation of IAM columns contains double-chain phosphatidylcholine, they are end-capped with decanoic, and propionic anhydride and are commercially available (IAM.PC.DD2 from Regis Technologies) [209]. A simple and economical alternative is the preparation of the stationary phase by coating a conventional RP column with phospholipids [34].
From a practical point of view, one of the main advantages of IAM columns is that they do not need the use of an organic modifier [27]. In fact, methanol as an organic modifier should be avoided as it disturbs the stability of the column by causing methanolysis of the phospholipids.
By simulating the phospholipid bilayer of biological membranes, the retention factor obtained with IAM chromatography correlates better with biological partition as the retention mechanism considers not only hydrophobic but also polar/ionic interactions [27,208]. For neutral solutes, the correlation between log P and the IAM partition coefficient (log kW-IAM) is strong. For ionized compounds, the correlation is poor due to the contribution of electrostatic interactions, which are not recognized in the octanol–water system [189]. When compared with partitioning with liposomes, log kIAM are well correlated for hydrophobic analytes [190]. Due to their resemblance to biological barriers, IAM chromatography has been successfully applied to calculate complex pharmacokinetic properties, such as cell permeation [210], intestinal absorption [183], unbound volume of distribution [184], and drug-induced phospholipidosis [186].

2.3.6. Immobilized Liposome Chromatography

Immobilized liposome chromatography (ILC) is another tool that has been used in anisotropic lipophilicity measurement. ILC stationary phases are prepared by immobilizing liposomes into the support silica particles by either covalent, hydrophobic, or electrostatic interactions (Figure 4b) [34,211].
When compared to IAM chromatography, ILC are better surrogates of the biological membranes because their liposomes have a bilayer structure and present a curvature effect [212]. In addition, the lipidic composition can be tailored to mimic specific biological barriers. Liu et al. studied different ILC column compositions aimed to mimic the small intestine membrane [192]. The authors suggested that liposomes composed of egg PC–PS–PE–cholesterol (5:1:2:2) were a suitable model for the prediction of the fraction absorption in humans. Despite its potential, the applications of the ILC method are very limited (Table 1) because of the low stability of liposomes [211].

2.3.7. Micellar Liquid Chromatography (MLC)

Micellar liquid chromatography (MLC) is a variant of RP-LC in which a surfactant at a concentration above the critical micellar concentration is added to the mobile phase, leading to the formation of micelles within the aqueous mobile phase (Figure 4c) [194,213]. MLC can be coupled to TLC, HPLC, and over-pressured-layer chromatography using micellar mobile phases [213].
In MLC, the micellar aggregates act as partition sites in the mobile phase, while the surfactant monomers present in the mobile phase modify the stationary phase either by both hydrophobic and/or silanophilic adsorption [203]. Due to the amphiphilic character of surfactants, the micellar mobile phase mimics the phospholipidic bilayer structure of biological membranes, while the modified stationary phase mimics not only the ordered array of the hydrophobic chains of phospholipids, mainly through the C18 carbon chains but also the polar heads of phospholipids, mainly through the adsorbed surfactant monomers [214]. During the elution, the solute can establish interactions with the micellar core, the micellar surface, the polar head of the surfactant bound to the stationary phase, and the alkyl chains bound to the stationary [194]. Therefore, solute retention involves the establishment of two different equilibriums: one between the bulk aqueous phase and the surfactant-coated stationary phase and another between the aqueous phase and the micelles [203]. The outcome index of MLC (log kMLC) is a metric of anisotropic lipophilicity as it depends on hydrophobic, electronic, and steric interactions [215]. Due to its similarity with biological membranes and extracellular fluids, MLC is also known as biopartitioning micellar chromatography [203].
Since the first reports of MLCs in the early 1980s, modifications of MLC have been reported [194]. These modifications are aimed to solve the decreased column efficiency in MLC due to slow mass transfer. The slow mass transfer is attributed to the poor wettability of a purely aqueous mobile phase and modification of the stationary phases by the adsorption of surfactant monomers [216]. To improve column efficiency, the main approaches are the addition of organic modifiers, the increase of column temperature, and decreased flow rate.
Exploiting the ability to mimic the phospholipidic barriers, log kMLC values have been used to predict several pharmacokinetic parameters, such as human oral absorption (HOA) [196,197,198,199], blood–brain barrier (BBB) permeability [196], plasma protein binding (PPB) [200,201], the volume of distribution [200,201], and therapeutic parameters [200,201]. When compared to other chromatographic techniques, MLC appears to be in the mid-ground between IAM and RP-HPLC. As an example, the lipophilicity of a set of fused azaisocytosine analogues was evaluated using RP-HPLC, IAM, and MLC [193]. For all compounds under study, the log kMLC values were higher than log kW-IAM but lower than log kW. In another example, the lipophilicity of a set of 88 compounds was evaluated using IAM and MLC [203]. Here, the obtained log kMLC values covered a narrower range than the obtained log kW-IAM values (log kW-IAM range: from −1.15 to 3.98, log kMLC range: from 1.84 to 3.98). These differences are attributed to the strong electrostatic interactions between positively charged drugs and the IAM phase and to the improved distribution of lipophilic compounds in the micellar MLC mobile phase. The chromatographic retention factors were used to predict cell permeability, HOA, and PPB. MLC retention factors demonstrated a better predictor capacity than IAM for cell permeability, while similar results were found for the remaining parameters. The advantage of the MLC is that it uses a conventional and easily available RP column, which is usually available in any analytical laboratory, while IAM requires a specific and expensive column.

3. Case Study: LC in Lipophilicity Assessment of Xanthone Derivatives

To emphasize the dual role of LC in the determination of lipophilicity and its importance as a medicinal chemistry tool, we provide a case study describing the experience of our group with xanthone derivatives. Xanthone derivatives comprise an important class of oxygenated heterocyclic compounds in medicinal chemistry concerning not only its large range of biological and/or pharmacological activities [217,218,219], but also other applications, including as fluorescence probes [220,221], and chiral stationary phases for LC [222,223]. For the last several years, the search for new xanthone derivatives with potential bioactivity has remained one of the main areas of interest for our group [224].
Recently, the determination of lipophilicity of a small library of bioactive xanthone derivatives was performed in our group [225]. The chosen xanthone derivatives were previously synthesized and exhibited interesting biological activities, namely growth inhibitory effects on different tumour cell lines [226], nerve conduction blockade acting in a similar manner to local anaesthetics [227], and the inhibition of cyclooxygenases, the enzymes involved in inflammatory processes [228].
Among the different methods used for lipophilicity assessment, LC was applied for both indirect and direct determination of log P, specifically with RP-HPLC and with VALLME coupled with HPLC [225].
Regarding the RP-HPLC method, two hydrophobic RP silica-based columns with a C8 and a C18 containing hydrocarbon long-chain (same length, diameter, and particle size) were used to evaluate their influence on the experimental lipophilicity. Different percentages of methanol in the aqueous mobile phase were chosen with increments of 5% or 10%, ranging between 40 to 75%, in accordance with the OECD Guidelines. For each compound, the log k was determined (Equation (5)) at different proportions of methanol. At least five different percentages were used to obtain a linear relationship, and log kw was obtained by extrapolation to 100% of pure water. The different percentages of methanol in the mobile phase were chosen based on the theoretical log P values for each xanthone derivative [225].
As an example of the xanthone derivative XD-10 (Figure 5), the log kw obtained in both columns (C8 and C18) was 3.71.
Figure 6a shows the chromatograms of XD-10 on the C18 column and with different proportions of methanol in the mobile phase. As expected, the retention time increased with the decrease of the percentage of methanol in the mobile phase.
Similar to other xanthone derivatives, for XD-10, a linear correlation between the chromatographic parameter log k and the proportion of methanol in the mobile phase was obtained, with R2 of 0.9967 and 0.9982 for C18 and C8 columns, respectively.
LC was also applied as an analytical method for direct quantification of the xanthone derivatives in the n-octanol phase after VALLME procedure [225]. For the analysis, a C18 hydrophobic silica-based column and a mixture of water/methanol (70:30 v/v) as a mobile phase were used. Once again, the theoretical log P values were the basis for the selection of the appropriate concentrations of the octanolic solutions for each xanthone derivative to fit the predictive concentrations in the n-octanol droplet into the linear zone of the calibration curve.
The external calibration curves for xanthone derivatives demonstrated good linear correlations. As an example, for XD-10 (Figure 5), a R2 of 0.9933 was achieved. Figure 6b shows the chromatograms for XD-10 present in n-octanol phase after VALLME procedure with different concentrations. As expected, the concentration of XD-10 found in the n-octanol phase increases with the augment of the concentration. The log P values for all the compounds were calculated for each concentration using Equation (3) and fell within a maximum range of ±0.24 units, which is in accordance with the OECD acceptance criterion [32]. Moreover, the methodologies and data assembled in this study contributed to enlarging the knowledge of the lipophilicity of xanthone derivatives.

4. Conclusions

Lipophilicity exerts a huge influence in pharmacokinetic and pharmacodynamic behavior of drugs, thus motivating the intensive search for methods to properly assess this property. As discussed in this work, several methods have been developed to predict, measure, or estimate the partition coefficient between a nonaqueous and an aqueous phase. In silico approaches are quite useful to rapidly predict log P values, but for decision-making purposes, the predicted values should always be complemented with experimentally obtained values. Direct methods allow for directly measuring the log P values. However, the n-octanol/water partition determined by the traditional shake-flask method is labor-consuming, and it is not always the most appropriate method. A modification of the shake-flask method has been explored, but the absence of comparative studies concerning them makes it very difficult to identify the best approach. Indirect methods, namely RP-HPLC, have been proposed to circumvent the drawbacks of direct methods. The advent of UPLC enables an even faster determination of lipophilicity due to the significant improvement of chromatographic performance and efficiency. In addition, indirect measurements of lipophilicity provide a partial but incomplete model of the partition in the n-octanol/water system because the intermolecular interactions occurring in the retention in the reversed-phase column are similar but not identical. The case study described in this work neatly emphasizes the dual role of LC in the determination of Log P.
Besides the variability associated with the method itself, the variability of model used as a nonaqueous phase also hampers the comparation between partition coefficients values. In fact, isotropic and anisotropic lipophilicity are expressed in different and hard to compare scales. This problem could be solved soon by high-throughput measurements of anisotropic lipophilicity, such as IAM and MLC, which could provide the large datasets required to settle the proper correlations between the isotropic and anisotropic partition coefficient values.

Author Contributions

Conceptualization, C.F. and M.M.M.P.; writing—original draft preparation, J.X.S. and Á.S.; writing—review and editing, J.X.S., M.M.M.P. and C.F.; project administration, C.F. and M.M.M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially supported by national funds through FCT (Foundation for Science and Technology) within the scope of Base Funding UIDB/04423/2020 and UIDP/04423/2020 (Group of Natural Products and Medicinal Chemistry-CIIMAR), UID/QUI/5000612019, the project PTDC/CTA-AMB/0853/2021, and the project PTDC/CTA-AMB/6686/2020. Also, CHIRALSINTESE-APSFCT-IINFACTS_2021. JXS thanks for the FCT PhD Programmes, specifically by the Biotech-Health Programme (PD/00016/2012), and for the grants (SFRH/BD/98105/2013 and SFRH/BD/116167/2016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Duffus, J.H.; Nordberg, M.; Templeton, D.M. Glossary of terms used in toxicology, 2nd edition (IUPAC recommendations 2007). Pure Appl. Chem. 2007, 79, 1153–1344. [Google Scholar] [CrossRef]
  2. Arnott, J.A.; Planey, S.L. The influence of lipophilicity in drug discovery and design. Expert Opin. Drug Discov. 2012, 7, 863–875. [Google Scholar] [CrossRef] [PubMed]
  3. van Balen, G.P.; Martinet, C.; Caron, G.; Bouchard, G.; Reist, M.; Carrupt, P.A.; Fruttero, R.; Gasco, A.; Testa, B. Liposome/water lipophilicity: Methods, information content, and pharmaceutical applications. Med. Res. Rev. 2004, 24, 299–324. [Google Scholar] [CrossRef] [PubMed]
  4. Loureiro, D.R.P.; Soares, J.X.; Lopes, D.; Macedo, T.; Yordanova, D.; Jakobtorweihen, S.; Nunes, C.; Reis, S.; Pinto, M.M.M.; Afonso, C.M.M. Accessing lipophilicity of drugs with biomimetic models: A comparative study using liposomes and micelles. Eur. J. Pharm. Sci. 2018, 115, 369–380. [Google Scholar] [CrossRef] [PubMed]
  5. Rutkowska, E.; Pajak, K.; Jozwiak, K. Lipophilicity—methods of determination and its role in medicinal chemistry. Acta Pol. Pharm. 2013, 70, 3–18. [Google Scholar]
  6. Liu, X.; Testa, B.; Fahr, A. Lipophilicity and its relationship with passive drug permeation. Pharm. Res. 2011, 28, 962–977. [Google Scholar] [CrossRef]
  7. Young, R.J.; Green, D.V.; Luscombe, C.N.; Hill, A.P. Getting physical in drug discovery II: The impact of chromatographic hydrophobicity measurements and aromaticity. Drug Discov. Today 2011, 16, 822–830. [Google Scholar] [CrossRef]
  8. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2012, 64, 4–17. [Google Scholar] [CrossRef]
  9. Laznicek, M.; Laznickova, A. The effect of lipophilicity on the protein binding and blood cell uptake of some acidic drugs. J. Pharm. Biomed. Anal. 1995, 13, 823–828. [Google Scholar] [CrossRef]
  10. Cardoso, T.; Almeida, A.S.; Remião, F.; Fernandes, C. Enantioresolution and Binding Affinity Studies on Human Serum Albumin: Recent Applications and Trends. Chemosensors 2021, 9, 304. [Google Scholar] [CrossRef]
  11. Gleeson, M.P. Generation of a set of simple, interpretable ADMET rules of thumb. J. Med. Chem. 2008, 51, 817–834. [Google Scholar] [CrossRef]
  12. Hughes, J.D.; Blagg, J.; Price, D.A.; Bailey, S.; DeCrescenzo, G.A.; Devraj, R.V.; Ellsworth, E.; Fobian, Y.M.; Gibbs, M.E.; Gilles, R.W.; et al. Physiochemical drug properties associated with in vivo toxicological outcomes. Bioorg. Med. Chem. Lett. 2008, 18, 4872–4875. [Google Scholar] [CrossRef] [PubMed]
  13. Waring, M.J.; Johnstone, C. A quantitative assessment of hERG liability as a function of lipophilicity. Bioorg. Med. Chem. Lett. 2007, 17, 1759–1764. [Google Scholar] [CrossRef] [PubMed]
  14. Giaginis, C.; Tsantili-Kakoulidou, A. Alternative measures of lipophilicity: From octanol-water partitioning to IAM retention. J. Pharm. Sci. 2008, 97, 2984–3004. [Google Scholar] [CrossRef]
  15. Hann, M.M.; Keseru, G.M. Finding the sweet spot: The role of nature and nurture in medicinal chemistry. Nat. Rev. Drug Discov. 2012, 11, 355–365. [Google Scholar] [CrossRef] [PubMed]
  16. Leeson, P.D.; Springthorpe, B. The influence of drug-like concepts on decision-making in medicinal chemistry. Nat. Rev. Drug Discov. 2007, 6, 881–890. [Google Scholar] [CrossRef]
  17. Hitzel, L.; Watt, A.; Locker, K. An Increased Throughput Method for the Determination of Partition Coefficients. Pharm. Res. 2000, 17, 1389–1395. [Google Scholar] [CrossRef]
  18. Sangster, J. Octanol-Water Partition Coefficients: Fundamentals and Physical Chemistry; John Wiley & Sons: New York, NY, USA, 1997. [Google Scholar]
  19. Liang, C.; Lian, H.-Z. Recent advances in lipophilicity measurement by reversed-phase high-performance liquid chromatography. TrAC Trends Anal. Chem. 2015, 68, 28–36. [Google Scholar] [CrossRef]
  20. Bergstrom, C.A.; Norinder, U.; Luthman, K.; Artursson, P. Experimental and computational screening models for prediction of aqueous drug solubility. Pharm. Res. 2002, 19, 182–188. [Google Scholar] [CrossRef]
  21. Petrauskas, A.A.; Kolovanov, E.A. ACD/Log P method description. Perspect. Drug Discov. Des. 2000, 19, 99–116. [Google Scholar] [CrossRef]
  22. Ulrich, N.; Goss, K.-U.; Ebert, A. Exploring the octanol–water partition coefficient dataset using deep learning techniques and data augmentation. Commun. Chem. 2021, 4, 90. [Google Scholar] [CrossRef]
  23. Mannhold, R.; Poda, G.I.; Ostermann, C.; Tetko, I.V. Calculation of molecular lipophilicity: State-of-the-art and comparison of log P methods on more than 96,000 compounds. J. Pharm. Sci. 2009, 98, 861–893. [Google Scholar] [CrossRef] [PubMed]
  24. Lin, S.Y.; Yang, X.H.; Liu, H.H. Development of liposome/water partition coefficients predictive models for neutral and ionogenic organic chemicals. Ecotoxicol. Environ. Saf. 2019, 179, 40–49. [Google Scholar] [CrossRef] [PubMed]
  25. Endo, S.; Escher, B.I.; Goss, K.U. Capacities of Membrane Lipids to Accumulate Neutral Organic Chemicals. Environ. Sci. Technol. 2011, 45, 5912–5921. [Google Scholar] [CrossRef]
  26. Eadsforth, C.V.; Moser, P. Assessment of reverse-phase chromatographic methods for determining partition coefficients. Chemosphere 1983, 12, 1459–1475. [Google Scholar] [CrossRef]
  27. Giaginis, C.; Tsantili-Kakoulidou, A. Current State of the Art in HPLC Methodology for Lipophilicity Assessment of Basic Drugs. A Review. J. Liq. Chromatogr. Relat. Technol. 2007, 31, 79–96. [Google Scholar] [CrossRef]
  28. Lombardo, F.; Shalaeva, M.Y.; Tupper, K.A.; Gao, F.; Abraham, M.H. ElogPoct: A tool for lipophilicity determination in drug discovery. J. Med. Chem. 2000, 43, 2922–2928. [Google Scholar] [CrossRef]
  29. Pinto, P.; Machado, C.M.; Moreira, J.; Almeida, J.D.P.; Silva, P.M.A.; Henriques, A.C.; Soares, J.X.; Salvador, J.A.R.; Afonso, C.; Pinto, M.; et al. Chalcone derivatives targeting mitosis: Synthesis, evaluation of antitumor activity and lipophilicity. Eur. J. Med. Chem. 2019, 184, 111752. [Google Scholar] [CrossRef]
  30. OECD. Test No. 117: Partition Coefficient (n-Octanol/Water), HPLC Method; OECD Publishing: Paris, France, 2022. [Google Scholar]
  31. Lipinski, C.A. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods 2000, 44, 235–249. [Google Scholar] [CrossRef]
  32. OECD. Test No. 107: Partition Coefficient (n-Octanol/Water): Shake Flask Method; OECD Publishing: Paris, France, 2022. [Google Scholar]
  33. Berthod, A.; Carda-Broch, S. Determination of liquid–liquid partition coefficients by separation methods. J. Chromatogr. A 2004, 1037, 3–14. [Google Scholar] [CrossRef]
  34. Godard, T.; Grushka, E. The use of phospholipid modified column for the determination of lipophilic properties in high performance liquid chromatography. J. Chromatogr. A 2011, 1218, 1211–1218. [Google Scholar] [CrossRef] [PubMed]
  35. Roman, I.P.; Mastromichali, A.; Tyrovola, K.; Canals, A.; Psillakis, E. Rapid determination of octanol-water partition coefficient using vortex-assisted liquid-liquid microextraction. J. Chromatogr. A 2014, 1330, 1–5. [Google Scholar] [CrossRef]
  36. Hartmann, T.; Schmitt, J. Lipophilicity—Beyond octanol/water: A short comparison of modern technologies. Drug Discov. Today Technol. 2004, 1, 431–439. [Google Scholar] [CrossRef] [PubMed]
  37. Brooke, D.N.; Dobbs, A.J.; Williams, N. Octanol:water partition coefficients (P): Measurement, estimation, and interpretation, particularly for chemicals with P greater than 10(5). Ecotoxicol. Environ. Saf. 1986, 11, 251–260. [Google Scholar] [CrossRef]
  38. De Bruijn, J.; Busser, F.; Seinen, W.; Hermens, J. Determination of octanol/water partition coefficients for hydrophobic organic chemicals with the “slow-stirring” method. Environ. Toxicol. Chem. 1989, 8, 499–512. [Google Scholar] [CrossRef]
  39. Jabusch, T.W.; Swackhamer, D.L. Partitioning of polychlorinated biphenyls in octanol/water, triolein/water, and membrane/water systems. Chemosphere 2005, 60, 1270–1278. [Google Scholar] [CrossRef]
  40. Dohta, Y.; Yamashita, T.; Horiike, S.; Nakamura, T.; Fukami, T. A system for LogD screening of 96-well plates using a water-plug aspiration/injection method combined with high-performance liquid chromatography-mass spectrometry. Anal. Chem. 2007, 79, 8312–8315. [Google Scholar] [CrossRef]
  41. Nishimura, I.; Hirano, A.; Yamashita, T.; Fukami, T. Improvement of the high-speed logD assay using an injection marker for the water plug aspiration/injection method. J. Chromatogr. A 2009, 1216, 2984–2988. [Google Scholar] [CrossRef]
  42. Marine, N.A.; Klein, S.A.; Posner, J.D. Partition Coefficient Measurements in Picoliter Drops Using a Segmented Flow Microfluidic Device. Anal. Chem. 2009, 81, 1471–1476. [Google Scholar] [CrossRef]
  43. Danielsson, L.G.; Zhang, Y.H. Methods for determining n-octanol-water partition constants. TrAC Trends Anal. Chem. 1996, 15, 188–196. [Google Scholar] [CrossRef]
  44. Chen, Z.; Weber, S.G. High-throughput method for lipophilicity measurement. Anal. Chem. 2007, 79, 1043–1049. [Google Scholar] [CrossRef] [PubMed]
  45. Psillakis, E. Vortex-assisted liquid-liquid microextraction revisited. TrAC Trends Anal. Chem. 2019, 113, 332–339. [Google Scholar] [CrossRef]
  46. Dearden, J.C.; Bresnen, G.M. The Measurement of Partition Coefficients. Quant. Struct.-Act. Relat. 1988, 7, 133–144. [Google Scholar] [CrossRef]
  47. Bosch Ojeda, C.; Sánchez Rojas, F. Vortex-Assisted Liquid–Liquid Microextraction (VALLME): Applications. Chromatographia 2014, 77, 745–754. [Google Scholar] [CrossRef]
  48. Gao, X.; Yu, C.H.; Tam, K.Y.; Tsang, S.C. New magnetic nano-absorbent for the determination of n-octanol/water partition coefficients. J. Pharm. Biomed. Anal. 2005, 38, 197–203. [Google Scholar] [CrossRef]
  49. Yu, C.H.; Tam, K.; Tsang, S.C. A new high-throughput method utilizing porous silica-based nano-composites for the determination of partition coefficients of drug candidates. J. Sep. Sci. 2011, 34, 2505–2512. [Google Scholar] [CrossRef]
  50. Austin, R.P.; Davis, A.M.; Manners, C.N. Partitioning of Ionizing Molecules between Aqueous Buffers and Phospholipid Vesicles. J. Pharm. Sci. 1995, 84, 1180–1183. [Google Scholar] [CrossRef]
  51. Park, J.Y.; Rippie, E.G. Micellar distribution equilibria: Ultracentrifugal study of apparent partition coefficients. J. Pharm. Sci. 1977, 66, 858–861. [Google Scholar] [CrossRef]
  52. Loidl-Stahlhofen, A.; Hartmann, T.; Schöttner, M.; Rhöring, C.; Brodowsky, H.; Schmitt, J.; Keldenich, J. Multilamellar liposomes and solid-supported lipid membranes (TRANSIL): Screening of lipid-water partitioning toward a high-throughput scale. Pharm. Res. 2001, 18, 1782–1788. [Google Scholar] [CrossRef]
  53. Dolzonek, J.; Cho, C.W.; Stepnowski, P.; Markiewicz, M.; Thoming, J.; Stolte, S. Membrane partitioning of ionic liquid cations, anions and ion pairs—Estimating the bioconcentration potential of organic ions. Environ. Pollut. 2017, 228, 378–389. [Google Scholar] [CrossRef]
  54. Schrader, W.; Andersson, J.T. Fast and direct method for measuring 1-octanol-water partition coefficients exemplified for six local anesthetics. J. Pharm. Sci. 2001, 90, 1948–1954. [Google Scholar] [CrossRef] [PubMed]
  55. Pauletti, G.M.; Wunderli-Allenspach, H. Partition coefficients in vitro: Artificial membranes as a standardized distribution model. Eur. J. Pharm. Sci. 1994, 1, 273–282. [Google Scholar] [CrossRef]
  56. Escher, B.I.; Schwarzenbach, R.P. Partitioning of Substituted Phenols in Liposome–Water, Biomembrane–Water, and Octanol–Water Systems. Environ. Sci. Technol. 1996, 30, 260–270. [Google Scholar] [CrossRef]
  57. Terada, H. Determination of Log Poct by High-Performance Liquid Chromatography, and its Application in the Study Quantitative Structure-Activity Relationships. Quant. Struct.-Act. Relat. 1986, 5, 81–88. [Google Scholar] [CrossRef]
  58. Abbas, N.S.; Mohamed, Y.A.S.; Derayea, S.M.; Omar, M.A.; Saleh, G.A. Simple TLC-spectrodensitometric method for studying lipophilicity and quantitative analysis of hypoglycemic drugs in their binary mixture. Biomed. Chromatogr. 2021, 35, e5154. [Google Scholar] [CrossRef] [PubMed]
  59. De Mello, H.; Echevarria, A. Hydrophobicity study for some pyrazolo-pyridine derivatives by RP-TLC and RP-HPLC. J. Liq. Chromatogr. Relat. Technol. 2006, 29, 1317–1330. [Google Scholar] [CrossRef]
  60. Cozma, A.; Vlase, L.; Ignat, A.; Zaharia, V.; Gocan, S.; Grinberg, N. Prediction of the lipophilicity of eight new p-toluenesulfonyl-hydrazinothiazole and hydrazine-bis-thiazole derivatives: A comparison between rp-hptlc and rp-hplc. J. Liq. Chromatogr. Relat. Technol. 2012, 35, 590–601. [Google Scholar] [CrossRef]
  61. Dross, K.; Rekker, R.F.; de Vries, G.; Mannhold, R. The lipophilic behaviour of organic compounds: 3. The search for interconnections between reversed-phase chromatographic data and log Pfoct values. Quant. Struct.-Act. Relat. 1998, 17, 549–557. [Google Scholar] [CrossRef]
  62. Komsta, Ł.; Skibiński, R.; Berecka, A.; Gumieniczek, A.; Radkiewicz, B.; Radoń, M. Revisiting thin-layer chromatography as a lipophilicity determination tool—A comparative study on several techniques with a model solute set. J. Pharm. Biomed. Anal. 2010, 53, 911–918. [Google Scholar] [CrossRef]
  63. Cozma, A.; Zaharia, V.; Ignat, A.; Gocan, S.; Grinberg, N. Prediction of the lipophilicity of nine new synthesized selenazoly and three aroyl-hydrazinoselenazoles derivatives by reversed-phase high performance thin-layer chromatography. J. Chromatogr. Sci. 2012, 50, 157–161. [Google Scholar] [CrossRef]
  64. Liang, C.; Han, S.-y.; Qiao, J.-q.; Lian, H.-z.; Ge, X. Determination of the n-octanol/water partition coefficients of weakly ionizable basic compounds by reversed-phase high-performance liquid chromatography with neutral model compounds. J. Sep. Sci. 2014, 37, 3226–3234. [Google Scholar] [CrossRef] [PubMed]
  65. Dacić, M.; Uzunović, A.; Pilipović, S.; Alagić-Džambić, L.; Durić, K. RP-HPLC Determination of Lipophilicity in Series of Corticosteroids. In Proceedings of the CMBEBIH 2021, Cham, Switzerland, 21–24 April 2021; pp. 535–540. [Google Scholar]
  66. Dobričić, V.; Bošković, J.; Vukadinović, D.; Vladimirov, S.; Čudina, O. Estimation of lipophilicity and design of new 17β-carboxamide glucocorticoids using RP-HPLC and quantitative structure-retention relationships analysis. Acta Chromatogr. 2021, 34, 130–137. [Google Scholar] [CrossRef]
  67. Krmar, J.; Protic, A.; Dajic, N.; Zecevic, M.; Otasevic, B. Chromatographic and computational lipophilicity assessment of novel antibiofilm agents. J. Liq. Chromatogr. Relat. Technol. 2020, 43, 615–623. [Google Scholar] [CrossRef]
  68. Jevrić, L.R.; Karadžić, M.Ž.; Mandić, A.I.; Podunavac Kuzmanović, S.O.; Kovačević, S.Z.; Nikolić, A.R.; Oklješa, A.M.; Sakač, M.N.; Penov Gaši, K.M.; Stojanović, S.Z. Lipophilicity estimation and characterization of selected steroid derivatives of biomedical importance applying RP HPLC. J. Pharm. Biomed. Anal. 2017, 134, 27–35. [Google Scholar] [CrossRef]
  69. Sima, I.A.; Kot-Wasik, A.; Wasik, A.; Namieśnik, J.; Sârbu, C. Assessment of Lipophilicity Indices Derived from Retention Behavior of Antioxidant Compounds in RP-HPLC. Molecules 2017, 22, 550. [Google Scholar] [CrossRef]
  70. Łączkowski, K.Z.; Biernasiuk, A.; Baranowska-Łączkowska, A.; Zavyalova, O.; Redka, M.; Malm, A. Synthesis, lipophilicity determination, DFT calculation, antifungal and DPPH radical scavenging activities of tetrahydrothiophen-3-one based thiazoles. J. Mol. Struct. 2018, 1171, 717–725. [Google Scholar] [CrossRef]
  71. Cimpan, G.; Irimie, F.; Gocan, S.; Claessens, H.A. Role of stationary phase and eluent composition on the determination of log P values of N-hydroxyethylamide of aryloxyalkylen and pyridine carboxylic acids by reversed-phase high-performance liquid chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 1998, 714, 247–261. [Google Scholar] [CrossRef]
  72. Giaginis, C.; Tsantili-Kakoulidou, A. The performance of 1-ethyl-3-methylimidazolium tetrafluoroborate ionic liquid as mobile phase additive in HPLC-based lipophilicity assessment. Biomed. Chromatogr. 2011, 25, 606–612. [Google Scholar] [CrossRef]
  73. El Tayar, N.; Van De Waterbeemd, H.; Gryllaki, M.; Testa, B.; Trager, W.F. The lipophilicity of deuterium atoms a comparison of shake-flask and high performance liquid chromatographic methods. Int. J. Pharm. 1984, 19, 271–282. [Google Scholar] [CrossRef]
  74. Welerowicz, T.; Buszewski, B. The effect of stationary phase on lipophilicity determination of beta-blockers using reverse-phase chromatographic systems. Biomed. Chromatogr. 2005, 19, 725–736. [Google Scholar] [CrossRef]
  75. Zagórska, A.; Czopek, A.; Pełka, K.; Bajda, M.; Stanisz-Wallis, K.; Pawłowski, M. Reversed-phase high-performance liquid chromatography study of lipophilicity of imidazo [2,1-f]theophylline derivatives. Acta Pol. Pharm. Drug Res. 2015, 72, 663–669. [Google Scholar]
  76. Altomare, C.; Carotti, A.; Cellamare, S.; Ferappi, M. Lipophilicity measurements of benzenesulfonamide inhibitors of carbonic-anhydrase by reversed-phase hplc. Int. J. Pharm. 1989, 56, 273–281. [Google Scholar] [CrossRef]
  77. Griffin, S.; Wyllie, S.G.; Markham, J. Determination of octanol-water partition coefficient for terpenoids using reversed-phase high-performance liquid chromatography. J. Chromatogr. A 1999, 864, 221–228. [Google Scholar] [CrossRef]
  78. Mrkvickova, Z.; Kovarikova, P.; Balikova, S.; Klimes, J. Determination of lipophilicity of novel potential antituberculotic agents using HPLC on monolithic stationary phase and theoretical calculations. J. Pharm. Biomed. Anal. 2008, 48, 310–314. [Google Scholar] [CrossRef] [PubMed]
  79. Brychtova, K.; Jampilek, J.; Opatrilova, R.; Raich, I.; Farsa, O.; Csollei, J. Synthesis, physico-chemical properties and penetration activity of alkyl-6-(2,5-dioxopyrrolidin-1-yl)-2-(2-oxopyrrolidin-1-yl)hexanoates as potential transdermal penetration enhancers. Biorg. Med. Chem. 2010, 18, 73–79. [Google Scholar] [CrossRef]
  80. Ayouni, L.; Cazorla, G.; Chaillou, D.; Herbreteau, B.; Rudaz, S.; Lanteri, P.; Carrupt, P.A. Fast determination of lipophilicity by HPLC. Chromatographia 2005, 62, 251–255. [Google Scholar] [CrossRef]
  81. Zhang, L.; Li, Z.G.; Huang, R.Q.; Wang, Q.S. Lipophilicity determination of N-(benzothiazol-2-yl)-alpha-amino alkyl phosphonic diesters by RP-HPLC and RP-HPTLC. Chin. J. Chem. 2000, 18, 872–876. [Google Scholar] [CrossRef]
  82. Lesniewska, M.A.; Gdaniec, Z.; Muszalska, I. Calculation procedures and HPLC method for analysis of the lipophilicity of acyclovir esters. Drug Dev. Ind. Pharm. 2015, 41, 663–669. [Google Scholar] [CrossRef]
  83. Du, C.M.; Valko, K.; Bevan, C.; Reynolds, D.; Abraham, M.H. Rapid gradient RP HPLC method for lipophilicity determination: A solvation equation based comparison with isocratic methods. Anal. Chem. 1998, 70, 4228–4234. [Google Scholar] [CrossRef]
  84. Dolezel, J.; Hirsova, P.; Opletalova, V.; Dohnal, J.; Marcela, V.; Kunes, J.; Jampilek, J. Rhodanineacetic Acid Derivatives as Potential Drugs: Preparation, Hydrophobic Properties and Antifungal Activity of (5-Arylalkylidene-4-oxo-2-thioxo-1,3-thiazolidin-3-yl)acetic Acids. Molecules 2009, 14, 4197–4212. [Google Scholar] [CrossRef]
  85. Hong, H.; Wang, L.; Zou, G. Retention in RP-HPLC: Lipophilicity Determination of Substituted Biphenyls by Reversed-Phase High Performance Liquid Chromatography. J. Liq. Chromatogr. Relat. Technol. 1997, 20, 3029–3037. [Google Scholar] [CrossRef]
  86. Giaginis, C.; Theocharis, S.; Tsantili-Kakoulidou, A. Octanol/water partitioning simulation by RP-HPLC for structurally diverse acidic drugs: Comparison of three columns in the presence and absence of n-octanol as the mobile phase additive. J. Sep. Sci. 2013, 36, 3830–3836. [Google Scholar] [CrossRef] [PubMed]
  87. Musiol, R.; Jampilek, J.; Podeszwa, B.; Finster, J.; Tabak, D.; Dohnal, J.; Polanski, J. RP-HPLC determination of lipophilicity in series of quinoline derivatives. Cent. Eur. J. Chem. 2009, 7, 586–597. [Google Scholar] [CrossRef]
  88. Vrakas, D.; Panderi, I.; Hadjipavlou-Litina, D.; Tsantili-Kakoulidou, A. Investigation of the relationships between logP and various chromatographic indices for a series of substituted coumarins. Evaluation of their similarity/dissimilarity using multivariate statistics. QSAR Comb. Sci. 2005, 24, 254–260. [Google Scholar] [CrossRef]
  89. Kostecka, M. Synthesis of a New Group of Aliphatic Hydrazide Derivatives and the Correlations between Their Molecular Structure and Biological Activity. Molecules 2012, 17, 3560–3573. [Google Scholar] [CrossRef]
  90. Pallicer, J.M.; Sales, J.; Roses, M.; Rafols, C.; Bosch, E. Lipophilicity assessment of basic drugs (log Po/w determination) by a chromatographic method. J. Chromatogr. A 2011, 1218, 6356–6368. [Google Scholar] [CrossRef]
  91. Cozma, A.; Vlase, L.; Ignap, A.; Zaharia, V.; Gocan, S.; Marutoiu, C.; Fodor, A. Lipophilicity study of new selenazole derivatives by RP-HPLC. Rev. Chim. 2012, 63, 651–655. [Google Scholar]
  92. Dołowy, M.; Pyka, A. Lipophilicity assessment of spironolactone by means of reversed phase liquid chromatography and by newly developed calculation procedures. Acta Pol. Pharm. Drug Res. 2015, 72, 235–244. [Google Scholar]
  93. Brychtova, K.; Dvorakova, L.; Opatrilova, R.; Raich, I.; Kacerova, S.; Placek, L.; Kalinowski, D.S.; Richardson, D.R.; Jampilek, J. Investigation of substituted 6-aminohexanoates as skin penetration enhancers. Biorg. Med. Chem. 2012, 20, 86–95. [Google Scholar] [CrossRef]
  94. Zheng, B.; West, L.M. Estimating the lipophilicity of natural products using a polymeric reversed phase hplc method. J. Liq. Chromatogr. Relat. Technol. 2010, 33, 118–132. [Google Scholar] [CrossRef]
  95. Zhao, Y.N.; Jona, J.; Chow, D.T.; Rong, H.J.; Semin, D.; Xia, X.Y.; Zanon, R.; Spancake, C.; Maliski, E. High-throughput logP measurement using parallel liquid chromatography/ultraviolet/mass spectrometry and sample-pooling. Rapid Commun. Mass Spectrom. 2002, 16, 1548–1555. [Google Scholar] [CrossRef] [PubMed]
  96. Hawrył, A.; Kuśmierz, E.; Hawrył, M.; świeboda, R.; Wujec, M. Determination of lipophilicity of new thiosemicarbazide and 1,2,4-triazole-3-thione derivatives using reversed-phase HPLC method and theoretical calculations. J. Liq. Chromatogr. Relat. Technol. 2015, 38, 430–437. [Google Scholar] [CrossRef]
  97. Lozowicka, B.; Kaczynski, P.; Nawrot, J. Study of lipophilicity of alpha-asarone derivatives and their deterrent activity against the Colorado potato beetle. Cent. Eur. J. Chem. 2013, 11, 2120–2133. [Google Scholar] [CrossRef]
  98. Britto, M.M.; Montanari, C.A.; Donnici, C.L.; Cass, Q.B. On the partitioning of some newly synthesized mesoionic 1,3,4-thiadiazolium-2-aminide and precursors evaluated by RP-HPLC. J. Liq. Chromatogr. Relat. Technol. 1999, 22, 357–366. [Google Scholar] [CrossRef]
  99. Abdou, W.M.; Barghash, R.F.; Sediek, A.A. Design of new arylamino-2-ethane-1,1-diyl- and benzoxazole-2-methylene-bisphosphonates vs. cytotoxicity and chronic inflammation diseases. From hydrophobicity prediction to synthesis and biological evaluation. Eur. J. Med. Chem. 2012, 57, 362–372. [Google Scholar] [CrossRef]
  100. Teijeiro, S.A.; Moroni, G.N.; Motura, M.I.; Brinon, M.C. Lipophilic character of pyrimidinic nucleoside derivatives: Correlation between shake flask, chromatographic (RP-TLC and RP-HPLC) and theoretical methods. J. Liq. Chromatogr. Relat. Technol. 2000, 23, 855–872. [Google Scholar] [CrossRef]
  101. Reymond, D.; Chung, G.N.; Mayer, J.M.; Testa, B. Lipophilicity measurement of nicotinates by reversed-phase high-performance liquid-chromatography—Differences in retention behavior, but similarities of log kw values, in methanol-water and acetonitrile-water eluents. J. Chromatogr. A 1987, 391, 97–109. [Google Scholar] [CrossRef]
  102. Jensen, D.A.; Gary, R.K. Estimation of alkane-water logP for neutral, acidic, and basic compounds using an alkylated polystyrene-divinylbenzene high-performance liquid chromatography column. J. Chromatogr. A 2015, 1417, 21–29. [Google Scholar] [CrossRef]
  103. Durcekova, T.; Boronova, K.; Mocak, J.; Lehotay, J.; Cizmarik, J. QSRR models for potential local anaesthetic drugs using high performance liquid chromatography. J. Pharm. Biomed. Anal. 2012, 59, 209–216. [Google Scholar] [CrossRef]
  104. Dimitriou-Christidis, P.; Harris, B.C.; McDonald, T.J.; Reese, E.; Autenrieth, R.L. Estimation of selected physicochemical properties for methylated naphthalene compounds. Chemosphere 2003, 52, 869–881. [Google Scholar] [CrossRef]
  105. Fikri, K.; Debord, J.; Bollinger, J.C.; Cledat, D.; Penicaut, B.; Robert, J.M.H. RP-HPLC lipophilicity studies for some (hetero)arylamides derived from 2-amino 4,6-dimethyl pyridine: Introduction of an hydrogen-bond descriptor. J. Liq. Chromatogr. Relat. Technol. 2011, 34, 1356–1366. [Google Scholar] [CrossRef]
  106. Dias, N.C.; Nawas, M.I.; Poole, C.F. Evaluation of a reversed-phase column (Supelcosil LC-ABZ) under isocratic and gradient elution conditions for estimating octanol-water partition coefficients. Analyst 2003, 128, 427–433. [Google Scholar] [CrossRef] [PubMed]
  107. Paschke, A.; Neitzel, P.L.; Walther, W.; Schuurmann, G. Octanol/water partition coefficient of selected herbicides: Determination using shake-flask method and reversed-phase high-performance liquid chromatography. J. Chem. Eng. Data 2004, 49, 1639–1642. [Google Scholar] [CrossRef]
  108. Strzemecka, L.; Hawryl, A.; Swieboda, R.; Hawryl, M.; Jagiello-Wojtowicz, E.; Piatkowska-Chmiel, I.; Herbet, M.; Chodkowska, A. Determination of Lipophilicity of Allyl Thiosemicarbazide, N-1-Thiocarbamylamidrazone Derivatives, and their Cyclic Products by RP-HPLC, RP-TLC, and Theoretical Methods: Effects of Selected Compounds on the CNS of Mice. J. Liq. Chromatogr. Relat. Technol. 2015, 38, 1452–1465. [Google Scholar] [CrossRef]
  109. Takacs-Novak, K.; Szasz, G.; Budvari-Barany, Z.; Jozan, M.; Lore, A. Relationship study between reversed phase HPLC retention and octanol/water partition among amphoteric compounds. J. Liq. Chromatogr. 1995, 18, 807–825. [Google Scholar] [CrossRef]
  110. Perisic-Janjic, N.; Kaliszan, R.; Wiczling, P.; Milosevic, N.; Uscumlic, G.; Banjac, N. Reversed-Phase TLC and HPLC Retention Data in Correlation Studies with in Silico Molecular Descriptors and Druglikeness Properties of Newly Synthesized Anticonvulsant Succinimide Derivatives. Mol. Pharmaceut. 2011, 8, 555–563. [Google Scholar] [CrossRef]
  111. Zhao, M.; Li, Z.; Wu, Y.; Tang, Y.-R.; Wang, C.; Zhang, Z.; Peng, S. Studies on log P, retention time and QSAR of 2-substituted phenylnitronyl nitroxides as free radical scavengers. Eur. J. Med. Chem. 2007, 42, 955–965. [Google Scholar] [CrossRef]
  112. Trezzi, M.M.; Vidal, R.A.; Dick, D.P.; Peralba, M.C.R.; Kruse, N.D. Sorptive behavior of sorgoleone in ultisol in two solvent systems and determination of its lipophilicity. J. Environ. Sci. Health B Pestic. Food Contam. Agric. Wastes 2006, 41, 345–356. [Google Scholar] [CrossRef]
  113. Liu, X.; Tanaka, H.; Yamauchi, A.; Testa, B.; Chuman, H. Lipophilicity Measurement by Reversed-Phase High-Performance Liquid Chromatography (RP-HPLC): A Comparison of Two Stationary Phases Based on Retention Mechanisms. Helv. Chim. Acta 2004, 87, 2866–2876. [Google Scholar] [CrossRef]
  114. Fei, X.; Zheng, Q.H. Lipophilicity coefficients of [11C]Me-Halo-CGS 27023A analogs determined by HPLC. J. Liq. Chromatogr. Relat. Technol. 2005, 28, 939–945. [Google Scholar] [CrossRef]
  115. Pagliara, A.; Khamis, E.; Trinh, A.; Carrupt, P.A.; Tsai, R.S.; Testa, B. Structural Properties Governing Retention Mechanisms on RP-HPLC Stationary Phases Used for Lipophilicity Measurements. J. Liq. Chromatogr. 1995, 18, 1721–1745. [Google Scholar] [CrossRef]
  116. Kubik, Ł.; Struck-Lewicka, W.; Kaliszan, R.; Wiczling, P. Simultaneous determination of hydrophobicity and dissociation constant for a large set of compounds by gradient reverse phase high performance liquid chromatography-mass spectrometry technique. J. Chromatogr. A 2015, 1416, 31–37. [Google Scholar] [CrossRef] [PubMed]
  117. Djaković-Sekulić, T.; Ačanski, M.; Perišić-Janjić, N. Evaluation of the predictive power of calculation procedure for molecular hydrophobicity of some estradiol derivates. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2002, 766, 67–75. [Google Scholar] [CrossRef]
  118. Sarbu, C.; Nascu-Briciu, R.D.; Casoni, D.; Kot-Wasik, A.; Wasik, A.; Namiesnik, J. Chromatographic lipophilicity determination using large volume injections of the solvents non-miscible with the mobile phase. J. Chromatogr. A 2012, 1266, 53–60. [Google Scholar] [CrossRef] [PubMed]
  119. Chrysanthakopoulos, M.; Nicolaou, I.; Demopoulos, V.J.; Tsantili-Kakoulidou, A. HPLC-based lipophilicity of pyrrolyl-acetic acid ARIs: Relationships with biological activity. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2010, 878, 61–67. [Google Scholar] [CrossRef]
  120. Eltayar, N.; Vandewaterbeemd, H.; Testa, B. Lipophilicity measurements of protonated basic compounds by reversed-phase high-performance liquid chromatography: II. Procedure for the determination of a lipophilic index measured by reversed-phase high-performance liquid chromatography. J. Chromatogr. A 1985, 320, 305–312. [Google Scholar] [CrossRef]
  121. Liu, X.; Tanaka, H.; Yamauchi, A.; Testa, B.; Chuman, H. Determination of lipophilicity by reversed-phase high-performance liquid chromatography. Influence of 1-octanol in the mobile phase. J. Chromatogr. A 2005, 1091, 51–59. [Google Scholar] [CrossRef]
  122. Caiali, E.; Casoni, D.; Ionita, P.; David, V.; Sarbu, C. Parabens lipophilicity determination with mobile phases containing low and medium hydrophobic alcohols. J. Liq. Chromatogr. Relat. Technol. 2014, 37, 2287–2301. [Google Scholar] [CrossRef]
  123. Han, S.-y.; Qiao, J.-q.; Zhang, Y.-y.; Yang, L.-l.; Lian, H.-z.; Ge, X.; Chen, H.-y. Determination of n-octanol/water partition coefficient for DDT-related compounds by RP-HPLC with a novel dual-point retention time correction. Chemosphere 2011, 83, 131–136. [Google Scholar] [CrossRef]
  124. Brychtova, K.; Opatrilova, R.; Raich, I.; Kalinowski, D.S.; Dvorakova, L.; Placek, L.; Csollei, J.; Richardson, D.R.; Jampilek, J. Investigating the activity of 2-substituted alkyl-6-(2,5-dioxopyrrolidin-1-yl)hexanoates as skin penetration enhancers. Biorg. Med. Chem. 2010, 18, 8556–8565. [Google Scholar] [CrossRef]
  125. Yamana, T.; Tsuji, A.; Miyamoto, E.; Kubo, O. Novel method for determination of partition coefficients of penicillins and cephalosporins by high-pressure liquid chromatography. J. Pharm. Sci. 1977, 66, 747–749. [Google Scholar] [CrossRef] [PubMed]
  126. Nowotnik, D.P.; Feld, T.; Nunn, A.D. Examination of some reversed-phase high-performance liquid chromatography systems for the determination of lipophilicity. J. Chromatogr. A 1993, 630, 105–115. [Google Scholar] [CrossRef]
  127. Koba, M.; Belka, M.; Ciesielski, T.; Baczek, T. Determination of lipophilicity for antitumor acridinone derivatives supported by gradient high-performance liquid chromatography method. Cent. Eur. J. Chem. 2012, 10, 216–223. [Google Scholar] [CrossRef]
  128. Valko, K.; Bevan, C.; Reynolds, D. Chromatographic Hydrophobicity Index by Fast-Gradient RP-HPLC: A High-Throughput Alternative to log P/log D. Anal. Chem. 1997, 69, 2022–2029. [Google Scholar] [CrossRef] [PubMed]
  129. Liu, X.; Hefesha, H.; Tanaka, H.; Scriba, G.; Fahr, A. Lipophilicity measurement of drugs by reversed phase HPLC over wide pH range using an alkaline-resistant silica-based stationary phase, XBridge(TM) Shield RP18. Chem. Pharm. Bull. 2008, 56, 1417–1422. [Google Scholar] [CrossRef] [PubMed]
  130. Huszar, M.; Varga, A.; Horvath, A.; Lorand, T.; Agocs, A.; Idei, M.; Mandl, J.; Vantus, T.; Keri, G. Comparative Characterization of Experimental and Calculated Lipophilicity and Anti-Tumour Activity of Isochromanone Derivatives. Curr. Med. Chem. 2010, 17, 321–333. [Google Scholar] [CrossRef]
  131. Chen, F.-y.; Cao, X.-w.; Han, S.-y.; Lian, H.-z.; Mao, L. Relationship between hydrophobicity and rplc retention behavior of amphoteric compounds. J. Liq. Chromatogr. Relat. Technol. 2014, 37, 2711–2724. [Google Scholar] [CrossRef]
  132. Bajda, M.; Bucki, A.; Szlek, J.; Szwaczkiewicz, M.; Swierczek, M.; Malawska, B. Determination of lipophilicity of alpha-(4-phenylpiperazine) derivatives of N-benzylamides using chromatographic and computational methods. Biomed. Chromatogr. 2008, 22, 428–432. [Google Scholar] [CrossRef]
  133. Harnisch, M.; Möckel, H.J.; Schulze, G. VII. International symposium on column liquid chromatographyRelationship between log Pow, shake-flask values and capacity factors derived from reversed-phase high-performance liquid chromatography for n-alkylbenzenes and some oecd reference substances. J. Chromatogr. A 1983, 282, 315–332. [Google Scholar] [CrossRef]
  134. Epifanio, I.; Genovese, S.; Carlucci, G.; Epifano, F.; Locatelli, M. Secondary plant metabolites LogP determination: The case of boropinic and geraniloxyferulic acids. Curr. Bioact. Compd. 2015, 11, 131–141. [Google Scholar] [CrossRef]
  135. Pachuta-Stec, A.; Hawrył, A.M.; Wróbel, A.; Hawrył, M.A.; Pitucha, M. Chromatographic Evaluation of the Lipophilic Properties of Some 1,2,4-Triazole with Potential Antitumour Activity. J. Liq. Chromatogr. Relat. Technol. 2015, 38, 1199–1206. [Google Scholar] [CrossRef]
  136. Bartalis, J.; Halaweish, F.T. Relationship between cucurbitacins reversed-phase high-performance liquid chromatography hydrophobicity index and basal cytotoxicity on HepG2 cells. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2005, 818, 159–166. [Google Scholar] [CrossRef] [PubMed]
  137. Podunavac-Kuzmanovic, S.O.; Jevric, L.R.; Tepic, A.N.; Sumic, Z. Reversed-phase HPLC retention data in correlation studies with lipophilicity molecular descriptors of carotenoids. Hem. Ind. 2013, 67, 933–940. [Google Scholar] [CrossRef]
  138. Barghash, R.F.; Ganoub, N.A.; Abdou, W.M. Development of a general and efficient route to highly lipophilic benzoxazole-2ylphosphonates and their antineoplastic properties. Monatsh. Chem. 2014, 145, 1621–1630. [Google Scholar] [CrossRef]
  139. Musilek, K.; Jampilek, J.; Dohnal, J.; Jun, D.; Gunn-Moore, F.; Dolezal, M.; Kuca, K. RP-HPLC determination of the lipophilicity of bispyridinium reactivators of acetylcholinesterase bearing a but-2-ene connecting linker. Anal. Bioanal. Chem. 2008, 391, 367–372. [Google Scholar] [CrossRef]
  140. Flieger, J.; Czajkowska-Zelazko, A.; Rzadkowska, M.; Szacon, E.; Matosiuk, D. Usefulness of reversed-phase HPLC enriched with room temperature imidazolium based ionic liquids for lipophilicity determination of the newly synthesized analgesic active urea derivatives. J. Pharm. Biomed. Anal. 2012, 66, 58–67. [Google Scholar] [CrossRef]
  141. Mornar, A.; Damic, M.; Nigovic, B. Pharmacokinetic Parameters of Statin Drugs Characterized by Reversed Phase High-Performance Liquid Chromatography. Anal. Lett. 2011, 44, 1009–1020. [Google Scholar] [CrossRef]
  142. Wiczling, P.; Kawczak, P.; Nasal, A.; Kaliszan, R. Simultaneous determination of pKa and lipophilicity by gradient RP HPLC. Anal. Chem. 2006, 78, 239–249. [Google Scholar] [CrossRef]
  143. Kobarfard, F.; Khalaj, A.; Daryaee, F.; Ardeshir, L.Z.; Rezaee, S. Correlation between different lipophilicity parameters and antimycobacterial activities of 2-hydroxyacetamides. DARU J. Pharm. Sci. 2008, 16, 55–59. [Google Scholar]
  144. Sztanke, M.; Tuzimski, T.; Janicka, M.; Sztanke, K. Structure-retention behaviour of biologically active fused 1,2,4-triazinones—Correlation with in silico molecular properties. Eur. J. Pharm. Sci. 2015, 68, 114–126. [Google Scholar] [CrossRef]
  145. Wiczling, P.; Waszczuk-Jankowska, M.; Markuszewski, M.J.; Kaliszan, R. The application of gradient reversed-phase high-performance liquid chromatography to the pK(a) and log k(w) determination of polyprotic analytes. J. Chromatogr. A 2008, 1214, 109–114. [Google Scholar] [CrossRef] [PubMed]
  146. Yuksek, H.; Akyildirim, O.; Yola, M.L.; Gursoy-Kol, O.; Celebier, M.; Kart, D. Synthesis, In Vitro Antimicrobial and Antioxidant Activities of Some New 4,5-Dihydro-1H-1,2,4-triazol-5-one Derivatives. Arch. Pharm. 2013, 346, 470–480. [Google Scholar] [CrossRef] [PubMed]
  147. Rezaee, S.; Khalaj, A.; Adibpour, N.; Saffary, M. Correlation between lipophilicity and antimicrobial activity of some 2-(4-substituted phenyl)-3(2H)-isothiazolones. DARU J. Pharm. Sci. 2009, 17, 256–263. [Google Scholar]
  148. Posa, M.; Pilipovic, A.; Lalic, M.; Popovic, J. Determination and importance of temperature dependence of retention coefficient (RPHPLC) in QSAR model of nitrazepams’ partition coefficient in bile acid micelles. Talanta 2011, 83, 1634–1642. [Google Scholar] [CrossRef] [PubMed]
  149. Pizarro, A.M.; McQuitty, R.J.; Mackay, F.S.; Zhao, Y.; Woods, J.A.; Sadler, P.J. Cellular Accumulation, Lipophilicity and Photocytotoxicity of Diazido Platinum(IV) Anticancer Complexes. ChemMedChem 2014, 9, 1169–1175. [Google Scholar] [CrossRef] [PubMed]
  150. Pallicer, J.M.; Pous-Torres, S.; Sales, J.; Roses, M.; Rafols, C.; Bosch, E. Determination of the hydrophobicity of organic compounds measured as log P(o/w) through a new chromatographic method. J. Chromatogr. A 2010, 1217, 3026–3037. [Google Scholar] [CrossRef]
  151. Flieger, J. Application of perfluorinated acids as ion-pairing reagents for reversed-phase chromatography and retention-hydrophobicity relationships studies of selected beta-blockers. J. Chromatogr. A 2010, 1217, 540–549. [Google Scholar] [CrossRef]
  152. Rudraraju, A.V.; Amoyaw, P.N.A.; Hubin, T.J.; Khan, M.O.F. Determination of log P values of new cyclen based antimalarial drug leads using RP-HPLC. Pharmazie 2014, 69, 655–662. [Google Scholar]
  153. Roda, A.; Minutello, A.; Angellotti, M.A.; Fini, A. Bile acid structure-activity relationship: Evaluation of bile acid lipophilicity using 1-octanol/water partition coefficient and reverse phase HPLC. J. Lipid Res. 1990, 31, 1433–1443. [Google Scholar] [CrossRef]
  154. Flieger, J.; Tatarczak-Michalewska, M.; Wujec, M.; Pitucha, M.; Swieboda, R. RP-HPLC analysis and in vitro identification of antimycobacterial activity of novel thiosemicarbazides and 1,2,4-triazole derivatives. J. Pharm. Biomed. Anal. 2015, 107, 501–511. [Google Scholar] [CrossRef]
  155. Komsta, L.; Pietras, R.; Bartuzi, E.; Skibinski, R.; Gumieniczek, A. Chemometric Comparison of Thin-Layer Chromatography, Gradient High-Performance Liquid Chromatography, and Computational Methods for Lipophilicity Assessment of Model Compounds. JPC J. Planar Chromatogr. Mod. TLC 2015, 28, 115–118. [Google Scholar] [CrossRef]
  156. Pietrogrande, M.C.; Borea, P.A.; Biagi, G.L. Lipophilicity measurement of benzodiazepine-receptor ligands by reversed-phase liquid chromatography. Comparison between high-performance liquid and thin-layer chromatography. J. Chromatogr. A 1988, 447, 404–409. [Google Scholar] [CrossRef]
  157. Han, S.Y.; Liang, C.; Qiao, J.Q.; Lian, H.Z.; Ge, X.; Chen, H.Y. A novel evaluation method for extrapolated retention factor in determination of n-octanol/water partition coefficient of halogenated organic pollutants by reversed-phase high performance liquid chromatography. Anal. Chim. Acta 2012, 713, 130–135. [Google Scholar] [CrossRef] [PubMed]
  158. Hawryl, A.M.; Popiolek, L.P.; Hawryl, M.A.; Swieboda, R.S.; Niejedli, M.A. Chromatographic and Calculation Methods for Analysis of the Lipophilicity of Newly Synthesized Thiosemicarbazides and their Cyclic Analogues 1,2,4-Triazol-3-thiones. J. Braz. Chem. Soc. 2015, 26, 1617–1624. [Google Scholar] [CrossRef]
  159. Reavill, C.; Walther, B.; Stolerman, I.P.; Testa, B. Behavioral and pharmacokinetic studies on nicotine, cytisine and lobeline. Neuropharmacology 1990, 29, 619–624. [Google Scholar] [CrossRef]
  160. Casoni, D.; Kot-Wasik, A.; Namiesnik, J.; Sarbu, C. Lipophilicity data for some preservatives estimated by reversed-phase liquid chromatography and different computation methods. J. Chromatogr. A 2009, 1216, 2456–2465. [Google Scholar] [CrossRef]
  161. Sztanke, K.; Markowski, W.; Swieboda, R.; Polak, B. Lipophilicity of novel antitumour and analgesic active 8-aryl-2,6,7,8-tetrahydroimidazo 2,1-c 1,2,4 triazine-3,4-dione derivatives determined by reversed-phase HPLC and computational methods. Eur. J. Med. Chem. 2010, 45, 2644–2649. [Google Scholar] [CrossRef]
  162. Almási, J.; Takács-Novák, K.; Kökösi, J.; Vámos, J. Characterization of potential NMDA and cholecystokinin antagonists II. Lipophilicity studies on 2-methyl-4-oxo-3H-quinazoline-3-alkyl-carboxylic acid derivatives. Int. J. Pharm. 1999, 180, 13–22. [Google Scholar] [CrossRef]
  163. Giaginis, C.; Theocharis, S.; Tsantili-Kakoulidou, A. Contribution to the standardization of the chromatographic conditions for the lipophilicity assessment of neutral and basic drugs. Anal. Chim. Acta 2006, 573, 311–318. [Google Scholar] [CrossRef]
  164. Benhaim, D.; Grushka, E. Effect of n-octanol in the mobile phase on lipophilicity determination by reversed-phase high-performance liquid chromatography on a modified silica column. J. Chromatogr. A 2008, 1209, 111–119. [Google Scholar] [CrossRef]
  165. Kovačević, S.; Banjac, M.K.; Podunavac-Kuzmanović, S.; Milošević, N.; Ćurčić, J.; Vulić, J.; Šeregelj, V.; Banjac, N.; Ušćumlić, G. Chromatographic and computational screening of anisotropic lipophilicity and pharmacokinetics of newly synthesized 1-aryl-3-ethyl-3-methylsuccinimides. Comput. Biol. Chem. 2020, 84, 107161. [Google Scholar] [CrossRef] [PubMed]
  166. Henchoz, Y.; Guillarme, D.; Martel, S.; Rudaz, S.; Veuthey, J.-L.; Carrupt, P.-A. Fast log P determination by ultra-high-pressure liquid chromatography coupled with UV and mass spectrometry detections. Anal. Bioanal. Chem. 2009, 394, 1919–1930. [Google Scholar] [CrossRef] [PubMed]
  167. Pallicer, J.M.; Calvet, C.; Port, A.; Rosés, M.; Ràfols, C.; Bosch, E. Extension of the liquid chromatography/quantitative structure-property relationship method to assess the lipophilicity of neutral, acidic, basic and amphotheric drugs. J. Chromatogr. A 2012, 1240, 113–122. [Google Scholar] [CrossRef] [PubMed]
  168. Henchoz, Y.; Guillarme, D.; Rudaz, S.; Veuthey, J.L.; Carrupt, P.A. High-throughput log P determination by ultraperformance liquid chromatography: A convenient tool for medicinal chemists. J. Med. Chem. 2008, 51, 396–399. [Google Scholar] [CrossRef]
  169. Kovačević, S.; Karadžić Banjac, M.; Anojčić, J.; Podunavac-Kuzmanović, S.; Jevrić, L.; Nikolić, A.; Savić, M.; Kuzminac, I. Chemometrics of anisotropic lipophilicity of anticancer androstane derivatives determined by reversed-phase ultra high performance liquid chromatography with polar aprotic and protic modifiers. J. Chromatogr. A 2022, 1673, 463197. [Google Scholar] [CrossRef]
  170. Rybka, S.; Obniska, J.; Żmudzki, P.; Koczurkiewicz, P.; Wójcik-Pszczoła, K.; Pękala, E.; Bryła, A.; Rapacz, A. Synthesis and Determination of Lipophilicity, Anticonvulsant Activity, and Preliminary Safety of 3-Substituted and 3-Unsubstituted N-[(4-Arylpiperazin-1-yl)alkyl]pyrrolidine-2,5-dione Derivatives. ChemMedChem 2017, 12, 1848–1856. [Google Scholar] [CrossRef]
  171. Guillot, A.; Henchoz, Y.; Moccand, C.; Guillarme, D.; Veuthey, J.-L.; Carrupt, P.-A.; Martel, S. Lipophilicity Determination of Highly Lipophilic Compounds by Liquid Chromatography. Chem. Biodivers. 2009, 6, 1828–1836. [Google Scholar] [CrossRef]
  172. Darrouzain, F.; Dallet, P.; Dubost, J.P.; Ismaili, L.; Pehourcq, F.; Bannwarth, B.; Matoga, M.; Guillaume, Y.C. Molecular lipophilicity determination of a huperzine series by HPLC: Comparison of C18 and IAM stationary phases. J. Pharm. Biomed. Anal. 2006, 41, 228–232. [Google Scholar] [CrossRef]
  173. Bajda, M.; Boryczka, S.; Wietrzyk, J.; Malawska, B. Investigation of lipophilicity of anticancer-active thioquinoline derivatives. Biomed. Chromatogr. 2007, 21, 123–131. [Google Scholar] [CrossRef]
  174. Giaginis, C.; Theocharis, S.; Tsantili-Kakoulidou, A. Investigation of the lipophilic behaviour of some thiazolidinediones Relationships with PPAR-gamma activity. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2007, 857, 181–187. [Google Scholar] [CrossRef]
  175. Matysiak, J.; Juszczak, M.; Karpinska, M.M.; Langner, E.; Walczak, K.; Lemieszek, M.K.; Skrzypek, A.; Niewiadomy, A.; Rzeski, W. Synthesis of 2-(2,4-dihydroxyphenyl)thieno-1,3-thiazin-4-ones, their lipophilicity and anticancer activity in vitro. Mol. Divers. 2015, 19, 725–736. [Google Scholar] [CrossRef] [PubMed]
  176. Malawska, B.; Kulig, K.; Bucki, A.; Zbek, P.; Wieckowska, A. The study of the lipophilicity of alpha-(4-phenylpiperazin-1-yl)-gamma-phthalimidobutyramides using chromatographic and computational methods. Biomed. Chromatogr. 2008, 22, 688–694. [Google Scholar] [CrossRef] [PubMed]
  177. Lazaro, E.; Rafols, C.; Abraham, M.H.; Roses, M. Chromatographic estimation of drug disposition properties by means of immobilized artificial membranes (IAM) and C18 columns. J. Med. Chem. 2006, 49, 4861–4870. [Google Scholar] [CrossRef] [PubMed]
  178. Essaid, D.; Chaminade, P.; Maillard, P.; Kasselouri, A. Lipophilicity of porphyrins and their retention in IAM, C8-C18 and HILIC chromatographic systems. J. Pharm. Biomed. Anal. 2015, 114, 227–240. [Google Scholar] [CrossRef]
  179. Flieger, J.; Pizon, M.; Plech, T. Chromatographic behavior of new antiepileptic active compounds on different reversed-phase materials. J. Chromatogr. A 2014, 1338, 188–196. [Google Scholar] [CrossRef]
  180. Bocian, S.; Buszewski, B. Comparison of retention properties of stationary phases imitated cell membrane in RP HPLC. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2015, 990, 198–202. [Google Scholar] [CrossRef]
  181. Dąbrowska, M.; Starek, M.; Komsta, Ł.; Szafrański, P.; Stasiewicz-Urban, A.; Opoka, W. Assessment of the chromatographic lipophilicity of eight cephalosporins on different stationary phases. Eur. J. Pharm. Sci. 2017, 101, 115–124. [Google Scholar] [CrossRef]
  182. Vrakas, D.; Giaginis, C.; Tsantili-Kakoulidou, A. Different retention behavior of structurally diverse basic and neutral drugs in immobilized artificial membrane and reversed-phase high performance liquid chromatography: Comparison with octanol–water partitioning. J. Chromatogr. A 2006, 1116, 158–164. [Google Scholar] [CrossRef]
  183. Grumetto, L.; Russo, G.; Barbato, F. Relationships between human intestinal absorption and polar interactions drug/phospholipids estimated by IAM-HPLC. Int. J. Pharm. 2015, 489, 186–194. [Google Scholar] [CrossRef]
  184. Valkó, K.L.; Nunhuck, S.B.; Hill, A.P. Estimating Unbound Volume of Distribution and Tissue Binding by In Vitro HPLC-Based Human Serum Albumin and Immobilised Artificial Membrane-Binding Measurements. J. Pharm. Sci. 2011, 100, 849–862. [Google Scholar] [CrossRef]
  185. Barbato, F.; Cirocco, V.; Grumetto, L.; La Rotonda, M.I. Comparison between immobilized artificial membrane (IAM) HPLC data and lipophilicity in n-octanol for quinolone antibacterial agents. Eur. J. Pharm. Sci. 2007, 31, 288–297. [Google Scholar] [CrossRef] [PubMed]
  186. Wang, J.; Zhao, X.; Xu, D.; Lai, L.; Crommen, J.; Guo, J.; Jiang, Z. Development of acidic phospholipid containing immobilized artificial membrane column to predict drug-induced phospholipidosis potency. J. Chromatogr. A 2021, 1647, 462147. [Google Scholar] [CrossRef] [PubMed]
  187. Grumetto, L.; Russo, G.; Barbato, F. Indexes of polar interactions between ionizable drugs and membrane phospholipids measured by IAM-HPLC: Their relationships with data of Blood-Brain Barrier passage. Eur. J. Pharm. Sci. 2014, 65, 139–146. [Google Scholar] [CrossRef] [PubMed]
  188. Yu, Z.; Ijzerman, A.P.; Heitman, L.H. K(v)11.1 (hERG)-induced cardiotoxicity: A molecular insight from a binding kinetics study of prototypical K(v)11.1 (hERG) inhibitors. Br. J. Pharmacol. 2015, 172, 940–955. [Google Scholar] [CrossRef]
  189. Vrakas, D.; Giaginis, C.; Tsantili-Kakoulidou, A. Electrostatic interactions and ionization effect in immobilized artificial membrane retention A comparative study with octanol-water partitioning. J. Chromatogr. A 2008, 1187, 67–78. [Google Scholar] [CrossRef]
  190. Taillardat-Bertschinger, A.; Martinet, C.A.M.; Carrupt, P.-A.; Reist, M.; Caron, G.; Fruttero, R.; Testa, B. Molecular Factors Influencing Retention on Immobilized Artificial Membranes (IAM) Compared to Partitioning in Liposomes and n-Octanol. Pharm. Res. 2002, 19, 729–737. [Google Scholar] [CrossRef]
  191. Liu, X.-Y.; Yang, Q.; Hara, M.; Nakamura, C.; Miyake, J. A novel chromatographic solid support with immobilized unilamellar liposomes for model analysis of solute–membrane interaction: Comparison with analysis using immobilized artificial membranes and free liposomal membranes. Mater. Sci. Eng. C 2001, 17, 119–126. [Google Scholar] [CrossRef]
  192. Liu, X.-Y.; Nakamura, C.; Yang, Q.; Kamo, N.; Miyake, J. Immobilized liposome chromatography to study drug–membrane interactions: Correlation with drug absorption in humans. J. Chromatogr. A 2002, 961, 113–118. [Google Scholar] [CrossRef]
  193. Janicka, M.; Mycka, A.; Sztanke, M.; Sztanke, K. Predicting Pharmacokinetic Properties of Potential Anticancer Agents via Their Chromatographic Behavior on Different Reversed Phase Materials. Int. J. Mol. Sci. 2021, 22, 4257. [Google Scholar] [CrossRef]
  194. Waters, L.J.; Shahzad, Y.; Mitchell, J.C. pH effects in micellar liquid chromatographic analysis for determining partition coefficients for a series of pharmaceutically related compounds. Curr. Pharm. Anal. 2012, 8, 272–277. [Google Scholar] [CrossRef]
  195. Koneva, A.S.; Ritter, E.; Anufrikov, Y.A.; Lezov, A.A.; Klestova, A.O.; Smirnova, N.A.; Safonova, E.A.; Smirnova, I. Mixed aqueous solutions of nonionic surfactants Brij 35/Triton X-100: Micellar properties, solutes’ partitioning from micellar liquid chromatography and modelling with COSMOmic. Colloids Surf. A 2018, 538, 45–55. [Google Scholar] [CrossRef]
  196. De Vrieze, M.; Janssens, P.; Szucs, R.; Van der Eycken, J.; Lynen, F. In vitro prediction of human intestinal absorption and blood–brain barrier partitioning: Development of a lipid analog for micellar liquid chromatography. Anal. Bioanal. Chem. 2015, 407, 7453–7466. [Google Scholar] [CrossRef] [PubMed]
  197. Molero-Monfort, M.; Escuder-Gilabert, L.; Villanueva-Camañas, R.M.; Sagrado, S.; Medina-Hernández, M.J. Biopartitioning micellar chromatography: An in vitro technique for predicting human drug absorption. J. Chromatogr. B Biomed. Sci. Appl. 2001, 753, 225–236. [Google Scholar] [CrossRef]
  198. Waters, L.J.; Shokry, D.S.; Parkes, G.M.B. Predicting human intestinal absorption in the presence of bile salt with micellar liquid chromatography. Biomed. Chromatogr. 2016, 30, 1618–1624. [Google Scholar] [CrossRef] [PubMed]
  199. Molero-Monfort, M.; Martín-Biosca, Y.; Sagrado, S.; Villanueva-Camañas, R.M.; Medina-Hernández, M.J. Micellar liquid chromatography for prediction of drug transport. J. Chromatogr. A 2000, 870, 1–11. [Google Scholar] [CrossRef]
  200. Martínez-Pla, J.J.; Sagrado, S.; Villanueva-Camañas, R.M.; Medina-Hernández, M.J. Retention–property relationships of anticonvulsant drugs by biopartitioning micellar chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2001, 757, 89–99. [Google Scholar] [CrossRef]
  201. Wu, L.-P.; Chen, Y.; Wang, S.-R.; Chen, C.; Ye, L.-M. Quantitative retention–activity relationship models for quinolones using biopartitioning micellar chromatography. Biomed. Chromatogr. 2008, 22, 106–114. [Google Scholar] [CrossRef]
  202. Dobričić, V.; Savić, J.; Nikolic, K.; Vladimirov, S.; Vujić, Z.; Brborić, J. Application of biopartitioning micellar chromatography and QSRR modeling for prediction of gastrointestinal absorption and design of novel β-hydroxy-β-arylalkanoic acids. Eur. J. Pharm. Sci. 2017, 100, 280–284. [Google Scholar] [CrossRef]
  203. Tsopelas, F.; Danias, P.; Pappa, A.; Tsantili-Kakoulidou, A. Biopartitioning micellar chromatography under different conditions: Insight into the retention mechanism and the potential to model biological processes. J. Chromatogr. A 2020, 1621, 461027. [Google Scholar] [CrossRef]
  204. Enami, T.; Nagae, N.; Doshi, S. The retention behaviour of reversed-phase HPLC columns when used under 100% aqueous conditions. LC-GC Europe 2003, 16, 418–425. [Google Scholar]
  205. Khan, B.M.; Liu, Y. High speed counter current chromatography: Overview of solvent-system and elution-mode. J. Liq. Chromatogr. Relat. Technol. 2018, 41, 629–636. [Google Scholar] [CrossRef]
  206. Valko, K. Application of high-performance liquid chromatography based measurements of lipophilicity to model biological distribution. J. Chromatogr. A 2004, 1037, 299–310. [Google Scholar] [CrossRef] [PubMed]
  207. Gluck, S.J.; Martin, E.J. Extended Octanol-Water Partition Coefficient Determination by Dual-Mode Centrifugal Partition Chromatography. J. Liq. Chromatogr. 1990, 13, 3559–3570. [Google Scholar] [CrossRef]
  208. Tsopelas, F.; Vallianatou, T.; Tsantili-Kakoulidou, A. Advances in immobilized artificial membrane (IAM) chromatography for novel drug discovery. Expert Opin. Drug Discov. 2016, 11, 473–488. [Google Scholar] [CrossRef]
  209. Technologies, R. Available online: https://www.registech.com/immobilized-artificial-membrane-iam-chromatography/iam-pc-dd2/iam-pc-dd2/ (accessed on 12 July 2022).
  210. Stepanić, V.; Žiher, D.; Gabelica-Marković, V.; Jelić, D.; Nunhuck, S.; Valko, K.; Koštrun, S. Physicochemical profile of macrolides and their comparison with small molecules. Eur. J. Med. Chem. 2012, 47, 462–472. [Google Scholar] [CrossRef]
  211. Lundahl, P.; Yang, Q. Liposome chromatography: Liposomes immobilized in gel beads as a stationary phase for aqueous column chromatography. J. Chromatogr. A 1991, 544, 283–304. [Google Scholar] [CrossRef]
  212. Eeman, M.; Deleu, M. From biological membranes to biomimetic model membranes. Biotechnol. Agron. Soc. Environ. 2010, 14, 719–736. [Google Scholar]
  213. Janicka, M.; Stępnik, K.; Pachuta-Stec, A. Quantification of Lipophilicity of 1,2,4-Triazoles Using Micellar Chromatography. Chromatographia 2012, 75, 449–456. [Google Scholar] [CrossRef]
  214. Escuder-Gilabert, L.; Martinez-Pla, J.J.; Sagrado, S.; Villanueva-Camanas, R.M.; Medina-Hernandez, M.J. Biopartitioning micellar separation methods: Modelling drug absorption. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2003, 797, 21–35. [Google Scholar] [CrossRef]
  215. Quiñones-Torrelo, C.; Sagrado, S.; Villanueva-Camañas, R.M.; Medina-Hernández, M.J. Development of Predictive Retention−Activity Relationship Models of Tricyclic Antidepressants by Micellar Liquid Chromatography. J. Med. Chem. 1999, 42, 3154–3162. [Google Scholar] [CrossRef]
  216. Khaledi, M.G.; Strasters, J.K.; Rodgers, A.H.; Breyer, E.D. Simultaneous enhancement of separation selectivity and solvent strength in reversed-phase liquid chromatography using micelles in hydro-organic solvents. Anal. Chem. 1990, 62, 130–136. [Google Scholar] [CrossRef] [PubMed]
  217. Fernandes, C.; Carraro, M.L.; Ribeiro, J.; Araújo, J.; Tiritan, M.E.; Pinto, M.M.M. Synthetic Chiral Derivatives of Xanthones: Biological Activities and Enantioselectivity Studies. Molecules 2019, 24, 791. [Google Scholar] [CrossRef]
  218. Soares, J.X.; Loureiro, D.R.P.; Dias, A.L.; Reis, S.; Pinto, M.M.M.; Afonso, C.M.M. Bioactive Marine Xanthones: A Review. Marine Drugs 2022, 20, 58. [Google Scholar] [CrossRef] [PubMed]
  219. Pinto, M.M.M.; Sousa, E.M.; Nascimento, S.J.M. Xanthone Derivatives: New Insights in Biological Activities. Curr. Med. Chem. 2005, 12, 2517–2538. [Google Scholar] [CrossRef]
  220. Takashima, I.; Kawagoe, R.; Hamachi, I.; Ojida, A. Development of an AND Logic-Gate-Type Fluorescent Probe for Ratiometric Imaging of Autolysosome in Cell Autophagy. Chem. Eur. J. 2015, 21, 2038–2044. [Google Scholar] [CrossRef] [PubMed]
  221. Sathyadevi, P.; Chen, Y.-J.; Wu, S.-C.; Chen, Y.-H.; Wang, Y.-M. Reaction-based epoxide fluorescent probe for in vivo visualization of hydrogen sulfide. Biosens. Bioelectron. 2015, 68, 681–687. [Google Scholar] [CrossRef] [PubMed]
  222. Phyo, Y.Z.; Teixeira, J.; Tiritan, M.E.; Cravo, S.; Palmeira, A.; Gales, L.; Silva, A.M.S.; Pinto, M.M.M.; Kijjoa, A.; Fernandes, C. New chiral stationary phases for liquid chromatography based on small molecules: Development, enantioresolution evaluation and chiral recognition mechanisms. Chirality 2020, 32, 81–97. [Google Scholar] [CrossRef] [PubMed]
  223. Fernandes, C.; Tiritan, M.E.; Cravo, S.; Phyo, Y.Z.; Kijjoa, A.; Silva, A.M.S.; Cass, Q.B.; Pinto, M.M.M. New chiral stationary phases based on xanthone derivatives for liquid chromatography. Chirality 2017, 29, 430–442. [Google Scholar] [CrossRef]
  224. Pinto, M.M.M.; Palmeira, A.; Fernandes, C.; Resende, D.I.S.P.; Sousa, E.; Cidade, H.; Tiritan, M.E.; Correia-da-Silva, M.; Cravo, S. From Natural Products to New Synthetic Small Molecules: A Journey through the World of Xanthones. Molecules 2021, 26, 431. [Google Scholar] [CrossRef]
  225. Santos, Á.; Soares, J.X.; Cravo, S.; Tiritan, M.E.; Reis, S.; Afonso, C.; Fernandes, C.; Pinto, M.M.M. Lipophilicity assessement in drug discovery: Experimental and theoretical methods applied to xanthone derivatives. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1072, 182–192. [Google Scholar] [CrossRef]
  226. Fernandes, C.; Masawang, K.; Tiritan, M.E.; Sousa, E.; de Lima, V.; Afonso, C.; Bousbaa, H.; Sudprasert, W.; Pedro, M.; Pinto, M.M. New chiral derivatives of xanthones: Synthesis and investigation of enantioselectivity as inhibitors of growth of human tumor cell lines. Biorg. Med. Chem. 2014, 22, 1049–1062. [Google Scholar] [CrossRef]
  227. Fernandes, C.; Oliveira, L.; Tiritan, M.E.; Leitao, L.; Pozzi, A.; Noronha-Matos, J.B.; Correia-de-Sá, P.; Pinto, M.M. Synthesis of new chiral xanthone derivatives acting as nerve conduction blockers in the rat sciatic nerve. Eur. J. Med. Chem. 2012, 55, 1–11. [Google Scholar] [CrossRef]
  228. Fernandes, C.; Palmeira, A.; Ramos, I.I.; Carneiro, C.; Afonso, C.; Tiritan, M.E.; Cidade, H.; Pinto, P.C.A.G.; Saraiva, M.L.M.F.S.; Reis, S.; et al. Chiral Derivatives of Xanthones: Investigation of the Effect of Enantioselectivity on Inhibition of Cyclooxygenases (COX-1 and COX-2) and Binding Interaction with Human Serum Albumin. Pharmaceuticals 2017, 10, 50. [Google Scholar] [CrossRef]
Figure 1. (a) Models used to determine isotropic and anisotropic lipophilicity; (b) forces governing intermolecular interactions between drugs and biological systems and their relationship with isotropic and anisotropic lipophilicity.
Figure 1. (a) Models used to determine isotropic and anisotropic lipophilicity; (b) forces governing intermolecular interactions between drugs and biological systems and their relationship with isotropic and anisotropic lipophilicity.
Chemosensors 10 00340 g001
Figure 2. VALLME workflow (adapted from [45]). (Reprint with permission from [45], Copyright (2019) ELSEVIER).
Figure 2. VALLME workflow (adapted from [45]). (Reprint with permission from [45], Copyright (2019) ELSEVIER).
Chemosensors 10 00340 g002
Figure 3. Octyl (a) and octadecyl (b) silica-based stationary phases.
Figure 3. Octyl (a) and octadecyl (b) silica-based stationary phases.
Chemosensors 10 00340 g003
Figure 4. (a) IAM HPLC column; (b) immobilized liposome chromatography (ILC) HPLC column; (c) schematic representation of micellar liquid chromatography (MLC).
Figure 4. (a) IAM HPLC column; (b) immobilized liposome chromatography (ILC) HPLC column; (c) schematic representation of micellar liquid chromatography (MLC).
Chemosensors 10 00340 g004
Figure 5. Chemical structure of xanthone derivative XD-10.
Figure 5. Chemical structure of xanthone derivative XD-10.
Chemosensors 10 00340 g005
Figure 6. Chromatograms of XD-10. (a): RP-HPLC method on the C18 column with different proportions of MeOH in the mobile phase, a flow rate of 1.0 mL/min, and detection at 254 nm; (b): VALLME method coupled to HPLC, with different XD-10 concentrations present in the n-octanol phase (chromatographic conditions: water/MeOH (70:30 v/v) as mobile phase, the flow rate of 1.0 mL/min, and detection at 254 nm). (Reprint with permission from [225], Copyright (2018) ELSEVIER).
Figure 6. Chromatograms of XD-10. (a): RP-HPLC method on the C18 column with different proportions of MeOH in the mobile phase, a flow rate of 1.0 mL/min, and detection at 254 nm; (b): VALLME method coupled to HPLC, with different XD-10 concentrations present in the n-octanol phase (chromatographic conditions: water/MeOH (70:30 v/v) as mobile phase, the flow rate of 1.0 mL/min, and detection at 254 nm). (Reprint with permission from [225], Copyright (2018) ELSEVIER).
Chemosensors 10 00340 g006
Table 1. Indirect methods for determining lipophilicity coupled with LC (full list of abbreviations is provided at the end as table footnote).
Table 1. Indirect methods for determining lipophilicity coupled with LC (full list of abbreviations is provided at the end as table footnote).
AnalytesAimChromatographic MethodChromatographic ConditionsRef.
Series of corticosteroidsEstimate lipophilicityRP-HPLC
(254 nm)
Columns: Zorbax Eclipse XDB-C18 (150 × 4.6 mm, 3.5 µm) and Zorbax Phenyl (250 × 4.6 mm, 5 µm); MP: mixtures of ACN (5–70%) and water with HCOOH 0.1%; FR: 1.0 mL/min[65]
17β-Carboxamide glucocorticoidsEstimate lipophilicityRP-HPLCColumns: Zorbax Eclipse XDB-C8 (150 × 4.6 mm, 5 µm), Zorbax Eclipse XDB-phenyl (150 × 4.6 mm, 5 µm), and Zorbax SB-CN (150 × 4.6 mm, 5 µm); MP: mixtures of MeOH or ACN (50–90%) and water;
FR: not specified
[66]
Antibiofilm agentsEstimate lipophilicityRP-HPLCColumns: end-capped RP BDS Hypersil C18 (30 × 4.0 mm, 3 µm) and end-capped Kinetex C8 (150 × 4.6 mm, 5 µm); MP: mixtures of ACN (45–55%) and water with ammonium acetate; FR: 1.0 mL/min[67]
Steroid derivativesEstimate lipophilicityRP-HPLC
(210 nm)
Column: Zorbax SB-C18 (250 × 3.0 mm, 5 µm); MP: mixtures of MeOH or ACN (70–80%) and water;
FR: 0.6 mL/min
[68]
Antioxidant compoundsDetermine the effects of stationary phase on the retention in terms of hydrophobicityRP-HPLC
(254 nm)
Columns: Purosphere RP-18e (125 × 3.0 mm, 5 µm), Zorbax Eclipse XDBC8 (150 × 4.6 mm, 5 µm), Discovery RP-Amide C16 (150 × 4.6 mm, 5 µm), CN100 Lichrosphere (250 × 4.0 mm, 5 µm), and pentafluorophenyl Kinetex (150 × 2.1 mm, 2.6 µm); MP: mixtures of MeOH (55–55%) and water with 0.1% formic acid; FR: 0.7 or 0.2 mL/min[69]
Tetrahydrothiophen-3-one based thiazolesEstimate lipophilicityRP-HPLC
(254 nm)
Column: Superspher 100 RP-18; MP: mixtures of MeOH (60–95%) and water; FR: 0.7 mL/min[70]
N-Hydroxyethylamides of aryloxyalkyllen and pyridine carboxylic
acids
Study the role of the stationary and mobile phases on the determination of log PRP-HPLC/
UV-Vis
(254 nm)
Columns: LiChrosorb RP-18 (125 × 4.0 mm, 5.0 μm), LiChrospher 60 RP-Select B (125 × 4.0 mm, 5.0 μm), Zorbax (5 μm, L = 150 × 4.6 mm; 5.0 μm), Zorbax-Eclipse XDB-C18 (150 × 4.6 mm; 5.0 μm); MP: mixtures of MeOH (0–60%) or ACN (0–50%) and water or 20 mM phosphate buffer at pH 7.0 or 20 mM tricine buffer at pH 7.0; FR: 1.0 mL/min[71]
Basic and neutral drugsEvaluate the performance of [EMIM][BF4] in producing extrapolated log kw indicesRP-HPLC/
UV-Vis
(220–254–268 nm)
Columns: Hypersil end-capped BDS C18 column (250 × 4.6 mm, 5.0 μm) and ABZ+ (150 × 4.6 mm, 5.0 μm); MP: mixtures of MeOH (20–80%) and 20 mM MOPS at pH 7.4 with [EMIM][BF4] (1%) and with octan-1-ol
(0–0.2%); FR: not specified
[72]
Deuterated benzene, toluene, p-xylene, pyridine, and aniline, and their isotopomers,Compare partition coefficients of deuterium atoms by shake-flask and HPLC methodsRP-HPLC/
UV-Vis
(254 nm)
Column: LiChrosorb RP-18 (300 × 4 mm, 10 μm); MP: mixtures of MeOH and water; FR: not specified[73]
β-Blockers and anti-arrhythmic drugsDetermine the effects of stationary phase on the retention in terms of hydrophobicityRP-HPLC/DAD
(220, 230, 254 and
280 nm)
Columns: SG-CHOL (250 × 4.6 mm) cholesterolic; SG-AP (250 × 4.6 mm); End-capped Supelcosil C18-DB (250 × 4.6 mm, 5.0 μm); Merck Chromolith Performance RP-18e (100 × 4.6 mm, 5.0 μm); MP: mixtures of ACN or MeOH (20–100%) (gradient mode);
FR: 1.0 mL/min
[74]
Arylpiperazines and tetrahydroisoquinoline derivatives with imidazo [2,1-f]purine-2,4-dione moietyEstimate lipophilicityRP-HPLC/DAD
(230 nm)
Column: LiChrospher 100 R C18 (100 × 4.6 mm, 5.0 μm); MP: mixtures of ACN (35–65%) and water with 0.01% TFA; FR: 1.0 mL/min[75]
meta and para Substituted benzenesulfonamidesEstimate lipophilicityRP-HPLC/
UV-Vis
(254 nm)
Column: μ-bondapak C18 (150 × 3.9 mm,10 μm); MP: mixtures of MeOH or ACN and water;
FR: 1.0 mL/min
[76]
TerpenoidsDetermine lipophilicity by RP-HPLC and compare with shake-flask and predicted valuesRP-HPLC/DAD (215 nm); RP-HPLC/RIColumn: Alltech Altima C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (30–75%) and 0.02 M MOPS buffer at pH 7.2; FR: 1.0 mL/min[77]
Anilides of pyrazine- 2-carboxylic acid and
4-benzylsulfanylpyridine derivatives
Determine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/
UV-Vis
(254 nm)
Column: Chromolith RP18e; (100 × 4.6 mm); MP: mixture of MeOH (37–42%) and 0.05 M phosphate buffer at pH 7.4; FR: 4.0 mL/min[78]
Esters of substituted 6-aminohexanoic acidDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/DAD
(204 nm)
Column: Zorbax Eclipse XDB C18 (150 mm × 4.6 mm, 5 μm); MP: mixture of MeOH (85%) and water;
FR: 0.4 mL/min
[79]
Set of compounds representing various functional groups such as ketones, acids, or alcoholsDetermine lipophilicity by RP-HPLC and compare it with reported valuesRP-HPLC/
UV-Vis
(220 and 254 nm)
Columns: Discovery RP Amide C16 (150 × 4.0 mm, 5.0 μm) and Discovery RP Amide C16 (20 × 4.0 mm, 5.0 μm); MP: mixtures of MeOH and phosphate buffer at pH 3 with octanol (0.25%); FR: 1.0 or 2.0 mL/min[80]
N-(Bemothiazol-2-yl)- α-amino alkyl phosphonic diestersCompare lipophilicity determination between RP-HPLC and RP-HPTLCRP-HPLC/
UV-Vis
(230 nm)
Column: ODS (250 × 4.6 mm, 5 μm); MP: mixtures of MeOH (70–90%) and water, FR: 1.0 mL/min[81]
Acyclovir estersDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/
UV-Vis
(240 nm)
Column: LiChrospher RP-18 (250 × 4.0 mm, 5.0 μm); MP: mixtures ACN or MeOH and 20 mM phosphate buffer pH 6.7; FR: 1.5 mL/min[82]
Diverse set of compoundsCompare the solvation equations obtained for different RP chromatographic retention parametersRP-HPLCColumn: ODS2-IK5 Inertsil (150 × 4.6 mm); MP: mixtures of ACN and 0.1% phosphoric acid or ammonium acetate buffer at pH 9.5 (gradient mode);
FR: 1.0 mL/min
[83]
[(5Z)-(5-Arylalkylidene-4-oxo-2-thioxo-1,3-thiazolidin-3-yl)]acetic acids derivativesDetermine lipophilicity by RP-HPLC and compare it with predicted valuesRP-HPLC/DAD
(210 nm)
Column: end-capped Symmetry C18 (250 × 4.6 mm, 5 μm); MP: mixture of MeOH (70%) and water;
FR: 0.9 mL/min
[84]
Weakly ionizable basic compounds and neutral compoundsDetermine lipophilicity by RP-HPLCRP-HPLC/DADColumn: Phenomenex Gemini C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH and 0.02 M ammonium chloride buffer at pH 7.4 and 9.0; FR: 1.0 mL/min[64]
4-Alkyl or alkoxy-4′-cyanobiphenyl derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-VIS
Column:C18 Nucleosil 7 (15 × 4.6 mm); MP: mixtures of MeOH (75–100%) and water; FR: 1.0 mL/min[85]
Structurally diverse acidic drugs both in the neutral and ionized formDetermine lipophilicity by RP-HPLC and compare three columns in the presence and absence of octan-1-ol as additiveRP-HPLC/
UV-Vis
(220, 254, 268 nm)
Columns: Supelcosil ABZ+ Plus (150 × 4.6 mm, 5 μm) and Supelcosil Aquasil (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (10–85%) and 20 m M MOPS buffer at pH 2.5 and 7.4 in the presence or absence of octan-1-ol; FR: 1.0 mL/min[86]
Quinoline derivativesDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/DAD
(210 nm)
Column: Symmetry C18 (250 × 4.6 mm, 5 μm); MP: mixtures of MeOH and water; FR: 0.9 mL/min[87]
Amidoximes or substituted heterocyclic coumarin derivativesStudy the relationship between log P and various chromatographic indices, including the extrapolated capacity factors derived by HPLC and reversed-phase TLCRP-HPLC/
UV-Vis
(254 nm)
Column: BDS C18 (250 × 4.6 mm, 5 μm); MP: mixtures of MeOH (40–90%) and water; FR: not specified[88]
Aliphatic hydrazide derivativesDetermine lipophilicityRP-HPLC/
UV-Vis
(320 nm)
Column: Hypersil BDS C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (45–95%) with acetate buffer at pH 4.0; FR: 0.5 mL/min[89]
Diverse set of compoundsDetermine lipophilicityRP-HPLC/DADColumns: Phenomenex Gemini NX (150 × 4.6 mm, 5 μm), Waters XTerra RP-18 (150 × 4.6 mm, 5 μm), Waters XTerra MS C18 (150 × 4.6 mm, 5 μm); MP: mixtures of ACN (40–50%) and pyrrolidine and ammonium hydrogenocarbonate buffer at pH 11.0;
FR: 1.0 mL/min
[90]
Selenazole derivativesDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/
UV-Vis
(210 nm)
Column: Zorbax SB-C18 (100 × 3.0 mm, 3.5 μm); MP: mixtures of MeOH (45–90%) and water; FR: not specified[91]
SpironolactoneDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/
UV-Vis
(238 nm)
Column: C18 Eurospher 100–5 (250 × 4.0 mm, 5 μm), MP: mixtures of MeOH or dioxane (55–95%) and water; FR: 1.0 mL/min[92]
6-Aminohexanoates derivativesDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/DAD
(204 nm)
Column: C18 Zorbax Eclipse XDB (150 × 4.0 mm, 5 μm); MP: mixture of MeOH (85%) and water;
FR: 0.4 mL/min
[93]
Marine natural productsDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/DADColumn: Hamilton polystyrene-divinylbenzene PRP-1 column (150 × 4.6 mm, 5 μm); MP: mixtures of ACN and 25 mM AcONH4 at pH 4.5, 7.2, 9.8 in the range 0–100%; FR: 1.0 mL/min[94]
Wide range of commercially available compoundsHigh-throughput log P measurement using parallel liquid chromatography/ultraviolet/mass spectrometry and sample-poolingRP-HPLC/UV/MSColumns: LUNA C18 (100 × 3.0 mm, 5 μm) for use on single channel LC/MS and LUNA C18 (50 × 3.0 mm, 5 μm) for high-throughput analysis on parallel LC/MS; MP: mixtures of MeOH (45–100%) with 20 mM ammonium carbonate at pH 8.0 or 20 mM ammonium formate at pH 1.0; FR: 0.45 mL/min[95]
Thiosemicarbazide and 1,2,4-triazole-3-thione derivativesDetermine lipophilicity by RP-HPLC and compare it with predicted valuesRP-HPLC/
UV-Vis
(254 nm)
Column: RP-18 Waters Symmetry (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (55–80%) or ACN (40–80%) and water; FR: 1.0 mL/min[96]
α-Asarone derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/DAD
(254 nm)
Columns: RP-18e Purospher STAR C18 (150 × 4.0 mm, 5 μm); RP-8e Purospher STAR C8 (150 × 4.6 mm, 5 μm); Zorbax Eclipse XDB-C (150 × 4.6 mm, 5 μm); Nucleosil Phenyl (250 × 4.6 mm, 7 μm); MP: mixtures of MeOH or ACN and water; FR: 1.0 mL/min[97]
Mesoionic 1,3,4-thuadiazolium-2-aminide derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
Column: C18 ODS-Shim-Pack (18.0 × 6.0 mm); MP: a mixture of MeOH (25–85%) and 0.005 M of phosphoric and glacial acetic acid buffer at pH 4.6;
FR: not specified
[98]
Arylamino-2-ethane-1,1-diyl- and benzoxazole-2-methylenebisphosphonatesDetermine lipophilicity by RP-HPLC and compare with predicted valuesRP-HPLC/DAD
(210 nm)
Column: Symmetry C18 (250 × 4.6 mm, 5μm); MP: mixture of MeOH (90%) and water; FR: 1.0 mL/min[99]
Pyrimidinic nucleoside derivativesDetermine lipophilicity by shake-flask, RP-TLC, RP-HPLC, and in silico methodsRP-HPLC/
UV-Vis
(265 nm)
Column: CLC-ODS(N)PN 228-17873-91(15 cm) QTY:2, packed with a C18 chemically bonded non-polar stationary phase; MP: mixtures of MeOH (30–80%) and water; FR: 1.0 mL/min[100]
Nicotinates estersDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(262 nm)
Column: LiChrosorb RP-18 (250 × 4 mm, 10 μm); MP: mixtures of MeOH or ACN and 0.02 M MOPS buffer at pH 7.0 wit N-decylamine (0.2%); FR: 1.5 mL/min[101]
Structurally diverse set of neutral, acidic, and basic compoundsDetermine lipophilicity by RP-HPLCRP-HPLC/variable wavelength detector
(220–300 nm)
Column: PRP-C18 (50 × 4.6 mm, 5 μm) or (33 × 2.1 mm, 5 μm); MP: mixtures of ACN and water with 0.2% phosphoric acid or 30 mM of diethylamine;
FR: 2.0 mL/min
[102]
Esters of alkoxyphenylcarbamic acidStudy the QSRR models for potential local anaesthetic drugsRP-HPLC/DAD
(210–290 nm)
Column: Separon SGX C18 (150 × 3.2 mm, 5 μm) or Separon SGX Phenyl (150 × 3.2 mm, 5 μm); MP: mixtures of MeOH or ACN (80%) and water;
FR: 1.0 mL/min
[103]
Methylated naphthalene derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/DAD
(254 nm)
Column: octadecylsilyl SUPELCOSIL LC-PAH (250 × 4.6 mm, 5 μm); MP: mixture of ACN (40–100%) and water; FR: 2.0 mL/min[104]
(Hetero)arylamides of 2-amino 4,6-dimethyl pyridineDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(254 or 285 nm)
Column: porous polystyrene/divinylbenzene copolymer PLRP-S (250 × 4.6 mm, 8 μm); MP: mixture of ACN (60%) and 0.02 M Na2HPO4 buffer at pH 9.4;
FR: 1.5 mL/min
[105]
Neutral compounds of varied structureEvaluate a RP column under isocratic and gradient elution conditions for estimating lipophilicityRP-HPLC/
UV-Vis
Column: Supelcosil LC-ABZ (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH and water; FR: 1.5 mL min[106]
Herbicides including triazines and phenylurea derivativesDetermine lipophilicity by shake-flask, RP-HPLC, and in silico methodsRP-HPLC/DAD
(218, 230, 245 or 270 nm)
Column: Spherisorb ODS 2 (125 × 4 mm, 5 μm); MP: mixtures of ACN (10–45%) and 1 mM ammonium acetate buffer at pH 6.7–7.3; FR: 0.8 mL/min[107]
Allyl thiosemicarbazide, N1-thiocarbamylamidrazone derivativesDetermine lipophilicity by RP-TLC, RP-HPLC and in silico methodRP-HPLC/
UV-Vis
(254 nm)
Column: RP-18 Waters Symmetry (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (50–80%) and water;
FR: 1.0 mL/min
[108]
Amphoteric compoundsDetermine lipophilicity RP-HPLCRP-HPLC/
UV-Vis
Column: C18 Hypersil 5 ODS (250 × 4.6 mm, 5 μm); MP: mixture of MeOH (50%) and phosphate buffer at pH range of 3.0 to 8.0; FR: 1.0 mL/min[109]
Succinimide derivativesDetermine lipophilicity by RP-TLC, RP-HPLC and in silico methodRP-HPLC/DADColumn: XTerra MS C-18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (5–100%) and water;
FR: 1.0 mL/min
[110]
2-Substituted phenylnitronyl nitroxidesDetermine lipophilicity RP-HPLCRP-HPLC/
UV-Vis
(275 nm)
Column: Rainbow C18 (150 × 4.6 mm); MP: mixture of MeOH (40%) and 0.001 M phosphate buffer;
FR: 0.25 mL/min
[111]
SorgoleoneDetermine lipophilicity RP-HPLCRP-HPLC/
UV-Vis
(280 nm)
Column: C18 (250 × 4.6 mm); MP: mixture of ACN (70%) and water; FR: 1.8 mL/min[112]
Solutes and drugs with well-defined solvatochromic parametersCompare two Stationary Phases Based on Retention MechanismsRP-HPLC/
UV-Vis
Columns: Supelcosil Silica-based Discovery-RP-Amide-C16 (50 × 4.6 mm, 5 μm) and Asahipak Polymer-based ODP-50-4B (50 × 4.6 mm, 5 μm); MP: mixtures of MeOH (10–80%) and 0.02 M phosphate buffer at pH 3.0, 4.0, or 7.0 with octan-1-ol (0.25%);
FR: 0.5 or 1 mL/min
[113]
[11C]Me-Halo-CGS 27023A analoguesDetermine lipophilicity RP-HPLCRP-HPLC/UV (240 nm) and γ-ray (NaI)Column: Prodigy C18 (250 × 4.6 mm, 5 μm); MP: mixture of ACN, MeOH, and 20 mM KHPO4 buffer at pH 6.7; FR: 1.5 mL/min[114]
Set of compounds covering a wide and regular range of structural parametersUnderstand the structural properties governing retention mechanisms on RP-HPLC stationary phasesRP-HPLC/
UV-Vis and RI
Column: ODS Supelcosil LC-ABZ (150 × 4.6 mm, 5 μm) pre-treated with electrostatic coating; MP: mixtures of MeOH (10–50%) and 0.02 M MOPS buffer at pH 7.4; FR: 1.0 mL/min[115]
Local anestheticsCompare a fast and direct method of determining lipophilicity against an indirect HPLC determinationRP-HPLC/UVColumn: C18 LiChroCART Purospher (125 × 3.0 mm, 5 μm); MP: not specified; FR: 0.5 mL min[54]
Large and diverse group of drugsDetermine lipophilicity and dissociation constantRP HPLC–ESI-TOF–MSColumn: XBridge-C18 (50 × 3.0 mm, 2.5 μm); MP: series of pH and organic modifier gradients;
FR: 0.5 mL/min
[116]
Estradiol derivativesEvaluate the predictive power of the calculation procedure for molecular hydrophobicityRP-HPLC/
UV-Vis
(254 nm)
Columns: LiChrosorb® RP-18 (250 × 4.0 mm, 5 μm) and LiChrosphert RP-8 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (60–95%) or ACN (70–90%) and water; FR: 1.0 mL/min[117]
Natural compounds (mycotoxins and alkaloids, and amines)Determine lipophilicity RP-HPLCRP HPLC/DAD/MSColumns: LiChroCART, Purospher RP-18e (125 × 3.0 mm; 5 μm) or LiChro-CART, LiChrospher RP-18e (250 × 4.0 mm; 5 μm) or Zorbax, Eclipse XDB-C8 (150 × 4.6 mm; 5 μm); MP: mixtures of MeOH or ACN and water at pH of 9.6 or 2.8; FR: 0.5 or 0.8 mL/min[118]
Pyrrolyl-acetic acid derivativesDetermine lipophilicity RP-HPLC at low pH and at pH 7.4RP-HPLC/
UV-Vis
(254 nm)
Column: Supelcosil ABZ+ (15 × 4.6 mm, 5 μm);
MP: different mixtures of MeOH (10–70%) and MOPS buffer at pH 3.0 and 7.4 with octan-1-ol (0–0.25%);
FR: not specified
[119]
Protonated basic compoundsDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(254 nm)
Column: LiChrosorb RP-18 (250 × 4 mm, 5 μm); MP: mixtures of MeOH and 0.02 M MOPS buffer with N-decylamine (0.2%); FR: 1.5 mL/min[120]
Drugs and flavonoidsStudy the influence of 1-octanol in the determination of lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(254 nm)
Column: Supelcosil Discovery-RP-Amide-C16 (50 × 4.6 mm, 5 μm); MP: mixtures of MeOH (10–70%) an 0.02 M phosphate buffer at pH 3.0 with octan-1-ol (0–0.25%); FR: 1.0 mL/min[121]
ParabensLipophilicity determination with mobile phases containing low and medium hydrophobic alcoholsRP-HPLC/MWD
(254 nm)
Column: double end-capped Zorbax Eclipse XDB-C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH with 1% of different alcohols and water with 0.1% H3PO4;
FR: 1.0 mL/min
[122]
DDT and related
compounds
Determine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
Column: Kromasil C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH and water; FR: 1.0 mL/min[123]
2-Substituted
alkyl-6-(2,5-dioxopyrrolidin-1-yl)hexanoates
Investigate the activity
as skin penetration enhancers
RP-HPLC/DAD (204 nm)Column: Zorbax Eclipse XDB (150 × 4.6 mm, 5 μm); MP: mixture of MeOH and water; FR: 0.4 mL/min[124]
Penicillins and cephalosporinsDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(254 nm)
Column: porous silica gel bonded chemically with octadecyl chains (250 mm); MP: mixture of MeOH and 0.035 M ammonium chloride buffer at pH 7.4;
FR: not specified
[125]
Set of diverse compoundsDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(210, 230 and 254 nm)
Column: Interaction ACT-l (150 × 4.6 mm, 10 μm), Nucleosil Cs (150 × 4.6 mm, 5 μm), Hamilton divinylbenzene-styrene copolymer (PRP-1) (150 × 4.1 mm, 10 μm); MP: mixtures of ACN (60–70%) and 0.1 M ammonium acetate buffer at pH 4.6;
FR: 0.75 or 2.0 mL/min
[126]
Acridinone derivativesDetermine the lipophilicity by RP-HPLCRP-HPLC/DAD
(254 nm)
Columns: Luna 5u C18(2) (150 × 4.6 mm, 5 μm), Candeza CD-C18 (150 × 4.6 mm, 3 μm), TSK-gel ODS-80TS (150 × 4.6 mm, 5 μm), Ascentis C18 (150 × 4.6 mm, 5 μm), Unison UK-C18 (150 × 4.6 mm, 3 μm), Zorbax SB-C8 column (75 × 4.6 mm, 3.5 μm);
MP: mixtures of ACN or MeOH and water with 0.1% of formic acid; FR: 1.0 mL/min
[127]
Neutral drugsDevelopment of a method to determine the lipophilicityRP-HPLC/DAD
(235, 255, 265 and 275 nm)
Column: Supelcosil LC-ABZ (50 × 4.6 mm, 5 μm); MP: mixture of MeOH (15–70%) and 20 mM MOPS buffer at pH 7.4 with 0.25% of octanol; FR: 0.5, 1.0, or 2.0 mL/min[28]
Structurally unrelated compoundsDetermine the lipophilicity by RP-HPLC with a fast generic gradientRP-HPLCColumn: ODS2-IK5 Inertsil (150 × 4.6 mm); MP: mixture of ACN (0–100%) and 50 mM ammonium acetate (pH ranging from 7.0 to 7.3); FR: 1.0 mL/min[128]
Drugs with diverse chemical natureDetermine the lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
Column: XBridgeTM Shield RP18 (50 × 4.6 mm, 5 μm); MP: mixtures of MeOH (70 to 10%) and 0.02 M phosphate buffer at pH 7.0 with 1-octanol/buffer partitioning a 0.25%; FR: 1.0 or 0.5 mL/min[129]
1H-Pyrazolo[3,4-b]pyridine derivativesDetermine the lipophilicity by RP-TLC and RP-HPLCRP-HPLC/
UV-Vis
(254 nm)
Column: LC8 SUPELCO (250 × 4.6 mm × 5μm); MP: mixtures of ACN (40–65%) and 0.1 M phosphate buffer at pH 7.4; FR: 1.5 mL/min[59]
Isochromanone derivativesDetermine the lipophilicity by RP-HPLCRP-HPLC/DADColumn: Hypersil 5 MOS (250 × 4.6 mm, 5 μm); MP: mixture of ACN (40%) and triethyl-ammonium phosphate buffer; FR: 1.0 mL/min[130]
Amphoteric compounds, amino acids, and small peptidesDetermine the lipophilicity by RP-HPLC in a broad pH rangeRP-HPLC/
UV-Vis
Column: Kromasil C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (10–90%) and ammonium acetate buffer at different pH; FR: 1.0 mL/min[131]
α-(4-Phenylpiperazine) derivatives of N-benzylamidesDetermine lipophilicity by RP-HPLC and in silico methodsRP-HPLC/
UV-Vis
(215 nm)
Column: Lichrospher RP18, (125 × 4.0 mm, 5 μm); MP: mixtures of ACN (45–65%) and water with addition of 0.1%TFA; FR: 1.0 mL/min[132]
N-Alkylbenzenes and some OECD reference substancesCorrelate log P shake-flask values and capacity factors derived from RP-HPLCRP-HPLC/
UV-Vis
Column: C18-SIL-X-5 (250 × 4.0 mm, 5 μm); MP: mixtures of MeOH (60–95%) and water; FR: 1.0 mL/min[133]
Natural secondary plant metabolitesDetermine lipophilicity by RP-HPLC and in silico methodsRP-HPLC/DADColumn: GraceSmart RP18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (90–30%) and 0.01 M phosphate buffer at pH 2.0; FR: 1.0 mL/min[134]
Derivatives of N-substituted amides of 3-(3-ethylthio-1,2,4-triazol-5-yl)propenoic acid)Determine the lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(254 nm)
Column: RP-18 Waters Symmetry (150 × 4.6 mm 5 μm); MP: mixtures of MeOH (50–75%) and water with or without ammonium buffer; FR: 1.0 mL/min[135]
CucurbitacinsCorrelate the hydrophobicity index and the basal cytotoxicity on HepG2 cellsRP-HPLC/DADColumn: Alltima C18 (250 × 4.6 mm, 5 μm) or Econosil C18 (150 × 4.6 mm, 5 μm); MP: mixtures of ACN and ammonium acetate buffer; FR: 1.0 mL/min[136]
CarotenoidsDetermine the lipophilicity by RP-HPLCRP-HPLC/DADColumn: Zorbax SB-C18 (250 × 3.0 mm, 5μm); MP: mixture of acetone (0–75%) and water (gradient mode); FR: 0.5 mL/min[137]
Highly lipophilic
benzoxazole-2ylphosphonates
Determine the lipophilicity by RP-HPLCRP-HPLC/DAD
(210 nm)
Column: Symmetry C18 (250 × 4.6 mm, 5 μm); MP: mixture of ethanol (90%) and water; FR: 1.0 mL/min[138]
Bispyridinium derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/DAD (210 nm)Column: Symmetry C18 (250 × 4.6 mm, 5μm); MP: mixture of MeOH (15%) and water; FR: 0.9 mL/min[139]
1-(1-Arylimidazolidyn-2-ylidyn)-3-arylalkyl urea derivativesCompare the conventional RP-HPLC and RP-HPLC enriched with room temperature imidazolium-based ionic liquidsRP-HPLC/DADColumn: Zorbax Extend-C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH or ACN and water with the addition of butyl-methyl imidazolium-based ionic liquids; FR: 1.0 mL/min[140]
Statin drugsDetermine lipophilicity by RP-HPLCRP-HPLC/DADColumns: Zorbax SB-C18 (250 × 4.6 mm, 5 μm), LiChrospher RP-C8 (250 × 4.0 mm, 5 μm); MP: mixtures of ACN (20–80%) and water; FR: 1.0 mL/min[141]
Basic and acidic analytesDetermine simultaneously pKa and lipophilicity by RP-HPLCRP-HPLC/DADColumn: XTerra MS C-18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH with buffers at different pH (gradient mode); FR: 1.0 mL/min[142]
Di-substituted 2-hydroxyacetamidesCorrelate different lipophilicity parameters and antimycobacterial activitiesRP-HPLC/
UV-Vis
(205 nm)
Column: RP-C18 (250 × 4.0 mm, 5 μm); MP: mixtures of MeOH (10–50%) and water; FR: 1.0 mL/min[143]
Fused 1,2,4-triazinones derivativesDetermine lipophilicity by RP-HPLC and in silico methodsRP-HPLC/
UV-Vis
(245 and 366 nm)
Column: Supelcosil LC-18 (150 × 4.6 mm, 5 μm); MP: mixtures of ACN (10–70%) or dioxane (5–80%) or MeOH (10–80%) and water; FR: 1.0 mL/min[144]
Polyprotic analytesDetermine pKa and lipophilicity by RP-HPLCRP-HPLC/DADColumn: XTerra MSC-18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH and different buffers;
FR: 1.0 mL/min
[145]
4,5-Dihydro-1H-1,2,4-triazol-5-one derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/
UV-Vis
(254 nm)
Column: Luna C18 (250 × 4.6 mm, 5 μm); MP: mixtures of MeOH (25–100%) and water; FR: 1.0 mL/min[146]
2-(4-Substituted phenyl)-3(2H)-isothiazolones derivativesDetermine lipophilicity by RP-HPLC and correlate it with biological activityRP-HPLC/
UV-Vis
(260 and 285 nm)
Column: octadecyl–poly(vinyl alcohol) (20 × 4.0 mm, 5 μm); MP: mixtures of MeOH and buffers at pH 2.0, 7.0 and 10.0 (gradient elution); FR: 1.5 mL/min[147]
Nitrazepam in bile acid micellesEvaluate the importance of temperature dependence of retention coefficient in QSAR modelRP-HPLC/DAD
(210 nm)
Column: Eclipse Plus C18 (250 × 3.0 mm, 5 μm); MP: mixture of MeOH and 0.01 M phosphate buffer at pH 7.0; FR: 1.0 mL/min[148]
Diazido Platinum(IV) complexesStudy the cellular accumulation, lipophilicity, and photocytotoxicityRP-HPLC/UVColumn: Agilent Zorbax eclipse plus C18 (250 × 4.6 mm, 5 μm); MP: mixtures of MeOH and water with 0.1% TFA; FR: 1.0 mL/min[149]
Organic compoundsDetermine lipophilicity by RP-HPLCRP-HPLC/DADColumns: Luna C18 (150 × 4.6 mm, 5 μm), Chromolith Performance RP-18e monolithic (100 × 4.6 mm); MP: mixtures of ACN or MeOH and buffers at different pH; FR: 1.0 or 2.0 mL/min[150]
β-BlockersApply perfluorinated acids as ion-pairing reagents for RP chromatography and retention-hydrophobicity relationshipsRP-HPLC/DAD
(220 nm)
Column: Zorbax Extend-C18 (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH and water with perfluorinated acids; FR: 1.0 mL/min[151]
Cyclen derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/DAD
(250 and 280 nm)
Column: Waters X-Bridge C-18 (150 × 4.6 mm, 5 μm); MP: mixture of ACN, MeOH and water (gradient mode); FR: 1.0 mL/min[152]
Naturally occurring bile acidsEvaluate the bile acid lipophilicityRP-HPLC/
UV-Vis
(200 nm)
Column: C-18 (100 mm, 5 μm); MP: mixture of MeOH (35–65%) and 0.01 M phosphate buffer at pH 7.0;
FR: 0.2 mL/min
[153]
Thiosemicarbazides and 1,2,4-triazole derivativesDetermine lipophilicity by RP-HPLCRP-HPLC/DADColumn: Zorbax 100 Extend-C18 (150 × 4.6 mm, 5 μm); MP: mixtures of can or MeOH and water;
FR: 1.0 mL/min
[154]
Simple model compoundsAssess the lipophilicity by different techniquesRP-HPLC/DADColumns: C18 (125 × 4.0 mm); CN (125 × 4.0 mm); C18e (125 × 4.0 mm); DIOL (125 × 4.0 mm); MP: mixtures of ACN or MeOH and water (gradient mode); FR: 1.0 mL/min[155]
Benzodiazepine-receptor ligandsMeasure the lipophilicity by RP-HPLC and RP-TLCRP-HPLC/
UV-Vis
(254 nm)
Column: μBondapak Cl8 (30 × 3.9 mm); MP: mixtures of ACN (30–70%) and phosphate buffer at pH 7.0;
FR: not specified
[156]
Halogenated organic pollutantsDetermine lipophilicity by RP-HPLCRP-HPLC/DADColumns: Venusil XBP C8 (150 × 2.1 mm, 5 μm); Venusil XBP C18 (150 × 2.1 mm, 5 μm); MP: mixtures of MeOH and water; FR: not specified[157]
Thiosemicarbazides and analoguesDetermine lipophilicity by chromatographic and in silico methodsRP-HPLC/
UV-Vis
(254 nm)
Column: RP-18 Waters Symmetry (150 × 4.6 mm, 5 μm); MP: mixtures of MeOH (40–65%) and water;
FR: 1.0 mL/min
[158]
Nicotine, cytisine and lobelineCorrelate the activity of cytisine and lobeline in ligand-binding and behavioural experiments and determine their lipophilicityRP-HPLC/
UV-Vis
(254 nm)
Column: Lichrosorb RP-18 (250 × 4.0 mm, 10 μm); MP: mixtures of MeOH and 0.02 M MOPS buffer at pH 7.4 with N-decylamine (0.2%); FR: 1.5 mL/min[159]
PreservativesDetermine lipophilicity by chromatographic and in silico methodsRP-HPLC/DAD/MS
(230, 254 and 366 nm)
Columns: endcapped C18 LiChroCART, Purosphere RP-18e (125 × 3.0 mm, 5 μm), double endcapped C8 Zorbax, Eclipse XDBC8, (150 × 4.6 mm, 5 μm), CN100 Lichrosphere, (250 × 4.0 mm, 5 μm), endcapped Supelcosil LC-NH2 (150 × 3.0 mm, 3 μm); MP: mixture of MeOH and water with 0.1% formic acid;
FR: 1.0 and 0.6 mL/min
[160]
8-Aryl-2,6,7,8-tetrahydroimidazo[2,1-c][1,2,4]triazine-3,4-dione derivativesDetermine lipophilicity by chromatographic and in silico methodsRP-HPLC/
UV-Vis
(254 nm)
Column: Zorbax Eclipse XDB C-18 (150 × 4.6 mm, 7 μm); MP: mixtures of MeOH and water;
FR: 1.0 mL/min
[161]
2-Methyl-4-oxo-3H-quinazoline-3-alkyl-carboxylic acid derivativesDetermine the lipophilicityRP-HPLC/
UV-Vis
(270 nm)
Column: Ultrasphere C8 (250 × 4.0 mm, 5 μm); MP: mixtures of MeOH and water; FR: 1.0 mL/min[162]
Neutral and basic drugsDetermine the lipophilicityRP-HPLC/
UV-Vis
Column: Hypersil BDS C-18 (250 × 4.6 mm, 5 μm); MP: mixtures of MeOH (85%) and 20 mM MOPS at pH 7.4 with N-decylamine (0.15–20%) or octan-1-ol (0.25%); FR: not specified[163]
Acidic and basic drugsEvaluate the effect of octan-1-ol in the mobile phaseRP-HPLC/DADColumn: Gemini C18 (150 × 4.6 mm, 5 μm), Gemini C18 (50 × 4.6 mm, 5 μm); MP: mixtures of MeOH with 0.25% of octan-1-ol (40–55%) and buffer saturated with octan-1-ol; FR: 1.0 mL/min[164]
p-Toluenesulfonyl-hydrazinothiazole and hydrazine-bis-thiazole derivativesCompare
between RP-HPLC and RP-HPTLC
RP-HPLC/
UV-Vis
(210 nm)
Column: Zorbax SB-C18 (100 × 3.0 mm, 3.5 µm); MP: mixtures of MeOH (45–90%) and water;
FR: not specified
[60]
1-Aryl-3-ethyl-3-methylsuccinimidesCompare
between RP-HPLC and RP-HPTLC
RP-HPLC
(254 nm)
Column: Luna C18 (250 × 4.6 mm, 5 µm); MP: mixtures of MeOH (55–55%) and water; FR: 1.0 mL/min[165]
Basic compounds such as β-blockers, local anesthetics, and piperazinesDetermine lipophilicityRP-UPLC/
UV-Vis or
RP-UPLC/MS
Columns: Acquity BEH Shield RP18, Acquity BEH C18, Acquity BEH C8, Acquity BEH Phenyl (30 mm × 2.1 mm, 1.7 μm); MP: mixtures of MeOH or ACN and aqueous buffer saturated in 1-octanol; FR: 0.5 mL/min[166]
Neutral, acidic, basic, and amphoteric drugsDetermine lipophilicityRP-UPLC/DAD
and RP-HPLC/UV-Vis
Columns: Gemini NX, (150 × 4.6 mm, 5 μm), Kinetex C18 (100 × 4.6 mm, 2.6 μm), Waters Acquity BEH C18 (50 × 2.1 mm, 1.7 μm); MP: mixtures of ACN (50%) and buffers at different pH values;
FR: 0.5 and 1.0 mL/min
[167]
Diverse set of compoundsDetermine the lipophilicityRP-UPLC/
UV-Vis
(215 or 220 nm)
Columns: Acquity BEH Shield RP18, Acquity BEH C18, Acquity BEH C8, AcquityBEH phenyl (30 × 2.1 mm, 1.7 μm); MP: mixtures of MeOH or ACN and aqueous buffer; FR: 0.5 mL/min[168]
Androstane derivativesDetermine the lipophilicityRP-UPLC/
UV-Vis
(210 and 230 nm)
Column: ZORBAX Eclipse C18 (2.1 × 50 mm, 1.8 µm); MP: mixtures of MeOH (30–60%) or ACN (30–60%) and water); FR: 0.3 mL/min[169]
N-[(4-Arylpiperazin-1-yl)alkyl]pyrrolidine-2,5-dione derivativesDetermine the lipophilicityRP-UPLC–MSColumn: Acquity UPLC BEH C18 (2.1 × 100 mm, 1.7 µm); MP: mixtures of methanol (30–90%) and 0.01 ammonium acetate at pH 7.4); FR: 0.3 mL/min[170]
Highly lipophilic compoundsDetermine lipophilicityRP-HPLC/
UV-Vis and UHPLC/UV-Vis
Columns: Discovery RPAmide C16 (20 × 4.0 mm, 5 μm) or Hypersil GOLD Javelin HTS (10 × 2.1 mm, 1.9 μm) or Acquity BEH Shield RP18 (30 × 2.1 mm, 1.7 μm); MP: mixtures of MeOH or ACN and water;
FR: 0.5 or 1.0 mL/min
[171]
Huperzine A derivativesDetermine the lipophilicityRP-HPLC/DAD and IAM-HPLC/DADColumns: ODS μBondapak C18 (300 × 3.9 mm) and IAM.PC.MG (150 × 4.6 mm); MP: mixtures of water and ACN; FR: 1.0 mL/min[172]
Thioquinoline derivativesDetermine lipophilicity by RP-HPLC and IAMRP-HPLC/UV-Vis and IAM-HPLC/UV-Vis (238, 215, 230, and 234 nm)Column: LiChrospher RP18 (125 × 4.0 mm, 5 μm), IAM. PC DD2 Regis (100 × 4.6 mm, 12 μm); MP: mixtures of ACN (55–90%) and water or mixtures of ACN (35–60%) and phosphate buffer at pH 5.4;
FR: 1.0 mL/min
[173]
ThiazolidinedionesStudy the lipophilic behaviour and the relationships with PPAR-γ activityRP-HPLC/UV-Vis and IAM-HPLC/UV-Vis (228, 267, or 254 nm)Columns: Hypersil BDS C-18 (250 × 4.6 mm, 5 μm), Supelcosil ABZ+ Plus (150 × 4.6 mm, 5 μm), Supelcosil Aquasil (150 × 4.6 mm, 5 μm), IAM.PC.DD2; MP: mixtures of MeOH (35–80%) and 20 mM MOPS buffer at 7.4 or mixtures of ACN (10 to 35%) and 0.01 M PBS buffer in the pH range 2.5–7.4; FR: 1.0, 2.0, or 3.0 mL/min[174]
2-(2,4-Dihydroxyphenyl)thieno-1,3-thiazin-4-one derivativesDetermine lipophilicity by RP-HPLC and IAMIAM-HPLC/
UV-Vis
(320 nm) and RP-HPLC/UV-Vis
(380 nm)
Column: HypersilGold C18 (100 × 3 mm, 3 μm) and IAM.PC.DD2 (100 × 4.6 mm, 12 μm); MP: mixtures of MeOH (30–95%) and 20 mM phosphate buffer at pH 4.0 or mixtures of ACN (10–50%) and 20 mM phosphate buffer at pH 7.4; FR: 0.5 or 1.0 mL/min[175]
Sα-(4-Phenylpiperazin-1-yl)-γ-phthalimido-butyramides derivativesDetermine lipophilicity by chromatographic and in silico methodsIAM-HPLC/UV-Vis and RP-HPLC/UV-Vis (210 nm)Column: LiChirospher 100 RP18 (100 × 4.6 mm, 5 μm) and IAM.PC.DD2 (100 × 4.6 mm); MP: mixtures of ACN (35–55%), water with 0.01% TFA; FR: 1.0 mL/min[176]
Solutes with sufficiently varied propertiesCharacterize an IAM column and study its similarity with common C18 chromatographic and biological systemsIAM-HPLC and RP-HPLC/
UV-Vis
(254 or 214 nm)
Columns: IAM.PC.DD2 (100 × 4.6 mm; 12 μm), XTerra MS C18 (150 × 4.6 mm, 5 μm), XTerra RP18 (150 × 4.6 mm, 5 μm); MP: mixtures of ACN and 0.01 M phosphate aqueous buffer adjusted to pH 7.0;
FR: 1.0 mL/min
[177]
PorphyrinsStudy the lipophilicity of and retention in IAM, C8-C18, and HILIC chromatographic systemsIAM-HPLC and RP-HPLC/DAD
(254 and 417 nm)
Columns: Gravity C18 (125 × 4.0 mm, 5 μm); Gravity C8 (125 × 4.0 mm, 5 μm); PolarTec (125 × 4.0 mm, 5 μm); HILIC (125 × 4.0 mm, 5 μm); IAM.PC.DD2 (30 × 4.6 mm, 10 μm); MP: mixtures of ACN (5–100%) and aqueous 20 mM ammonium acetate at pH 7.0 (gradient mode); FR: 1.0 mL/min[178]
4,5-Disubstituted-2,4-dihydro-3H- 1,2,4-triazole-3-thiones derivativesStudy the chromatographic behaviour on different reversed-phase materialsIAM-HPLC and RP-HPLC/DADColumns: Zorbax Extend-C18 (150 × 4.6 mm, 5 μm); Cogent UDC Cholesterol (150 mm × 4.6 mm, 4 μm); Regis IAM.PC. DD2 (150 × 4.6 mm, 10 μm); MP: mixtures of MeOH (75–90%) or ACN (55–70%) and water; FR: 1.0 mL/min[179]
Alkyl benzene derivatives and PAHs, flavonoids, nucleosides, and nucleic basesCompare retention properties of stationary phases imitated cell membrane in RP HPLCIAM-HPLC and RP-HPLC/DADColumn: IAM.PC.DD2 (150 × 4.6 mm, 12 μm); Amino-P-C18 (125 × 4.6 mm, 5 μm); MP: mixtures of MeOH or ACN and water; FR: 1.0 mL/min[180]
MacrolidesInvestigate the relationships between retention factors and cellular permeationRP-HPLC/MS IAM-HPLC/MSColumns: Luna C18 (50 × 3 mm), IAM PC DD; MP: mixture of ACN (0–100%) and 0.05 M ammonium acetate at pH 7.4; FR: 1.0 mL/min
Cephalosporins
Estimate lipophilicityIAM-HPLCColumns: IAM.PC.DD 2 (100 × 4.6 mm, 12 µm); MP: phosphate buffer at pH 6.9; FR: 1.0 mL/min[181]
Structurally diverse basic and neutral drugsCompare different IAM and reversed-phase retention factors and their relationships with lipophilicity dataIAM-HPLCColumn: IAM.PC.DD2, BDS C-18 (250 × 4.6 mm, 5 μm); MP: mixtures of ACN and 0.01 M PBS buffer or 0.02 M MOPS buffer at pH 7; FR: 1.0, 2.0, or 3.0 mL/min[182]
Structurally non-related orally administered drugsInvestigate possible relationships between pharmacokinetic parameters and intestinal permeation data.IAM-HPLCColumns: IAM.PC.MG (150 × 4.6 mm; 12 μm), IAM.PC.DD2 (100 × 4.6 mm; 10 μm); MP: mixtures of ACN and 0.1 M phosphate buffer at pH 7.0;
FR: not specified
[183]
Structurally diverse drugsInvestigate the relationships, retention factors, and volume of distributionIAM-HPLCColumn: IAM PC.DD2 (150 × 3.0 mm); MP:
Mobile phase A: mixtures of ACN (0–70%) and 0.05 M ammonium acetate at pH 7.4; FR: 1.0 mL/min
[184]
Quinolone drugsInvestigate possible relationships between IAM retention data and various lipophilicity scalesIAM-HPLCColumn: IAM.PC.MG (150 × 4.6 mm) and IAM.PC.DD2 (100 × 4.6 mm); MP: mixtures of ACN and 0.1 M phosphate buffer at pH 7.0; FR: 0.9 or 1.0 mL/min[185]
Structurally diverse drugsCharacterize column and study the relationships between IAM retention data and DIPLIAM-nano-HPLCMonolith column prepared via copolymerization of MDPC and MDPA; MP: 0.05 M ammonium acetate at pH 7.4 or 4.5; FR: 0.6 µL/min[186]
BBB permeant drugsInvestigate the relationships between BBB permeation data and lipophilicityIAM-HPLCColumns: IAM.PC.MG (150 × 4.6 mm; 12 μm); IAM.PC.DD2 (100 × 4.6 mm; 10 μm); MP: mixtures of ACN and 0.1 M phosphate buffer at pH 7.0;
FR: 1.0, 2.0 or 3.0 mL/min
[187]
Kv11.1 inhibitorsDetermine the lipophilicity by RP-HPLC and IAMIAM-HPLCColumn: Supelcosil LC-ABZ (50 × 4.6 mm, 5 μm), IAM.PC-DD2 (100 × 4.6 mm, 10 μm); MP: mixtures of water and MeOH or ACN; FR: not specified[188]
Set of basic, neutral, acidic and ampholytic drugsExplore the effect of electrostatic interactions of anionic speciesIAM-HPLCColumn: IAM.PC.DD2; MP: mixture of ACN (0–35%) and MOPS buffer pH: 7.4, 0.01 M PBS buffer at pH: 7.4, 0.01 M PBS buffer at pH: 5.0; FR: 1.2 or 3.0 mL/min[189]
Structurally diverse drugsCompare between IAM and equilibrium dialysisIAM-HPLC
(220 or 254 nm)
Column: IAM.PC.DD2 (100 × 4.6 m, 12 µm); MP: mixtures of methanol and 0.02 M phosphate buffer at pH 7.0; FR: 1.5 or 2.0 mL/min[190]
Neutral, acidic, and basic drugs, including steroid hormones, local anesthetics, β-blockers, and NSAIDsCharacterize a reversed-phase amide column coated with phosphatidylcholine-based liposomesILC/DAD
(230 or 254 nm)
Column: Ascentis® RP-amide HPLC column (50 × 4.6 mm, 5.0 μm) previously coated by immobilizing the PC liposomes on the column; MP: mixture of MeOH (20–80%) and phosphate buffer; FR: 1.0 and 2.0 mL/min[34]
β-blockers, imidazoline derivatives, and p-alkylphenolsCharacterize column coated with phosphatidylethanolamine-based liposomesILC/UV-Vis
(220 nm)
Column: 5 mm glass column containing PE liposomes bounded to Sephacryl S-1000 gel by avidin–biotin binding; MP: 0.01 M HEPES with NaCl (0.15 M,) at pH 7.4; FR: 0.3 mL/min[191]
Structurally diverse drugsEstimate lipophilicity and correlate it with absorptionILC/UV-Vis
(220 nm)
Column: 5–5.5 cm×5 mm glass column containing EPC, PS–chol, and EPC–PS–PE liposomes CNBr-Sepharose gel by avidin–biotin binding; MP: 0.01 M HEPES with NaCl (0.15 M,) at pH 7.4; FR: 0.3 mL/min[192]
Fused azaisocytosine analoguesCompare between RP-HPLC, IAM, and MLCIAM-HPLC MLC/UV-VisColumns: Spherisorb ODS-2 (125 × 4 mm, 5 μm), IAM.PC.DD2 (100 × 4.6 mm, 10 μm), Purosphere RP-18e column (125 × 4 mm, 5 μm); MP: 0.01 M phosphate citrate buffer at pH 7.4 with Brij 35 (0.15–0.075 M) and with isopropanol (7%); mixtures of ACN (20–50%) and 0.01 M phosphate citrate buffer at pH 7.4; FR: 0.1, 1 and 1.3 mL/min[193]
Series of pharmaceutically related compoundsStudy the effects of the variations in pH in MLC determinationMLC/UV-Vis
(221 or 285 nm)
Column: RP cyanopropyl Spherisorb (150 × 4.6 mm, 5μm); MP: 0.1 M phosphate buffer with SDS at pH 3.0, 4.0, 5.0, 5.8, 6.4 and 7.0; FR: 1.35 mL/min[194]
Benzene derivativesStudy the influence of mixtures of Triton X-100 and Brij 35MLC/DADColumn: Nucleodur C18 Gravity (2 × 125 mm, 5 μm); MP: HCl solution at pH 2 with different Brij 35/Triton X-100–ratios (0.05–0.30%); FR: 0.3 mL/min[195]
Structurally diverse drugsEstimate lipophilicity and correlate it with intestinal absorption and BBBMLC/UV-Vis (210 or 300 nm)Column: Grace GraceSmart C18 (150 × 2.1 mm, 3 μm); MP: 0.05 M phosphate buffer at pH 7.4 with 0.01 M miltefosine; FR: 0.2 mL/min[196]
Structurally diverse drugsEstimate lipophilicity and correlate it with absorptionMLC/UV-VisColumn: Kromasil octadecyl-silane C18 (150 × 4.6 mm, 5 μm); MP: 0.05 M phosphate buffer at pH 6.5 and 7.4 with 0.04 M Brij35; FR: 0.5 mL/min[197]
Structurally diverse drugsEstimate lipophilicity and correlate it with absorptionMLC/UV-Vis (221 or 360 nm)Column: Spherisorb (15 cm × 4.6 mm, 5 μm); MP: water with sodium deoxycholate (5–20 mM);
FR: 1.34 mL/min
[198]
Structurally diverse drugsEstimate lipophilicity and correlate it with absorptionMLC/UV-Vis (240 or 254 nm)Columns: Spherisorb octadecyl-silane ODS-2 C18 (120 × 4 mm, 5 μm), Kromasil octadecyl-silane ODS-2 C18 (50×4.6 mm, 5 μm), Kromasil octadecyl-silane C18 (150 × 4.6 mm, 5 μm); MP: 0.05 M phosphate buffer at pH 7.4 with 0.02 M Brij35; FR: 1 and 1.5 mL/min[199]
Anticonvulsant drugsEstimate lipophilicity and correlate it with pharmacokinetic parametersMLC/UV-Vis (220 or 240 nm)Column: Kromasil C18 column (150 × 4.6 mm, 5 μm); MP: 0.05 M phosphate buffer at pH 7.4 with 0.04 M Brij35; FR: 1 mL/min[200]
Quinolone drugsEstimate lipophilicity and correlate it with pharmacokinetic parametersMLC/UV-Vis (270 nm)Column: Kromasil C18 column (150 × 4.6 mm, 5 μm); MP: 0.05 M phosphate buffer at pH 7.4 with Brij35 and with NaCl (9.20 g/L); FR: 1 mL/min[201]
β-Hydroxy-β-arylalkanoic acids analoguesEstimate lipophilicity and correlate it with absorptionMLC/DADColumn: Zorbax Extend-C18 (150 mm × 4.6 mm, 5 μm); MP: 40 mM solution of Brij35 in a mixture of ACN (20%) and 0.07 M phosphate buffer at pH 5.50 with0.04 M Brij35; FR: 1 mL/min[202]
Structurally diverse drugsStudy the influence of pH, temperature, and ionic strength in MLC and compare it with IAMMLC/UV-Vis (220 nm)Column: High Stability C18 Supelco (150 × 4.6 mm, 5 μm); MP: 0.05 M phosphate buffer at pH 5.5 and 7.4 with NaCl (0–9.2 g/L) with 0.04 M Brij35;
FR: 1 mL/min
[203]
ACN—acetonitrile; BEH—bridged ethyl hybrid; Chol—cholesterol; DAD—diode-array detection; DIPL—drug-induced phospholipidosis; [EMIM][BF4]—1-ethyl-3 methylimidazolium tetrafluoroborate; EPC—egg phosphatidylcholine; TFA—trifluoroacetic acid; FR—flow rate; MeOH—methanol; MDPC—12-methacryloyl n-dodecylphosphocholine; MDPA—12-methacryloyl n-dodecylphosphoric acid; MLC—micellar liquid chromatography; MOPS—morpholinepropanesulphonic acid; MP—mobile phase; NSAIDs—non-steroidal anti-inflammatory drugs; PAHs—polycyclic aromatic hydrocarbons; PBS—phosphate buffer saline; PE—phosphatidylethanolamine; PS—phosphatidylserine; RI—refractive index; RP-HPLC—reversed-phase high-pressure liquid chromatography; RP-HPTLC—reversed-phase high-pressure thin layer chromatography; SDS—sodium dodecyl sulphate.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Soares, J.X.; Santos, Á.; Fernandes, C.; Pinto, M.M.M. Liquid Chromatography on the Different Methods for the Determination of Lipophilicity: An Essential Analytical Tool in Medicinal Chemistry. Chemosensors 2022, 10, 340. https://doi.org/10.3390/chemosensors10080340

AMA Style

Soares JX, Santos Á, Fernandes C, Pinto MMM. Liquid Chromatography on the Different Methods for the Determination of Lipophilicity: An Essential Analytical Tool in Medicinal Chemistry. Chemosensors. 2022; 10(8):340. https://doi.org/10.3390/chemosensors10080340

Chicago/Turabian Style

Soares, José X., Álvaro Santos, Carla Fernandes, and Madalena M. M. Pinto. 2022. "Liquid Chromatography on the Different Methods for the Determination of Lipophilicity: An Essential Analytical Tool in Medicinal Chemistry" Chemosensors 10, no. 8: 340. https://doi.org/10.3390/chemosensors10080340

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop