Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Dec 4.
Published in final edited form as: ACS Chem Neurosci. 2024 Nov 22;15(23):4322–4336. doi: 10.1021/acschemneuro.4c00571

Stereospecific properties and intracellular transport of novel intrinsically fluorescent neurosteroids

Vibeke Akkerman , Peter Reinholdt , Rasmus Schnoor-Madsen , Line Lauritsen , Jad Bader , Minxing Qian §, Yuanjiang Xu §, Gustav Akk , Holger Scheidt , Peter Müller , Douglas F Covey #, Alex S Evers #, Jacob Kongsted , Daniel Wüstner
PMCID: PMC11892034  NIHMSID: NIHMS2061931  PMID: 39574303

Abstract

Allopregnanolone (AlloP) is a neuroactive steroid (NAS), which is a potent allosteric activator of the gamma-aminobutyric acid A (GABAA) receptor. The mechanisms underlying the biological activity of AlloP and other NAS are only partially understood. Here, we present intrinsically fluorescent analogs of AlloP (MQ-323) and its 3β-epimer, E-AlloP (YX-11), and show, by a combination of spectroscopic and computational studies, that these analogs mimic the membrane properties of AlloP and E-AlloP very well. We found stereospecific differences in the orientation and dynamics of the NAS as well as in their impact on membrane permeability. However, all NAS are unable to condense the lipid bilayer, in stark contrast to cholesterol. Using Förster resonance energy transfer (FRET) and electrophysicological measurements, we show that MQ-323 but not YX-11 binds at the intersubunit site of the ELICα1GABAA receptor and potentiates GABA-induced receptor currents. In aqueous solvents, YX-11 forms aggregates at much lower concentrations than MQ-323, and loading both analogs onto cyclodextrin allows for their uptake by human astrocytes, where they become enriched in lipid droplets, as shown by quantitative fluorescence microscopy. Trafficking of the NAS analogs is stereospecific, as uptake and lipid droplet targeting is more pronounced for YX-11 compared to MQ-323. In summary, we present novel minimally modified analogs of AlloP and E-AlloP, which enable us to reveal stereospecific membrane properties, allosteric receptor activation, and intracellular transport of these neurosteroids. Our fluorescence design strategy will be very useful for the analysis of other NAS in the future.

Keywords: Allopregnanolone, fluorescence, probes, trafficking, microscopy, astrocytes

Graphical Abstract

graphic file with name nihms-2061931-f0001.jpg

Introduction

Neuroactive steroids (NAS) are steroids that modulate brain function in various ways, thereby impacting neuronal activity, stress, memory, and mental health. NAS are produced from cholesterol, either in the periphery to cross the blood-brain barrier, or they are synthesized locally in the brain.1 Upon cholesterol import into mitochondria, steroid-producing cells can convert it into pregnenolone by P450 side-chain cleaving enzyme.2 Pregnenolone is the precursor for a variety of other sterols, including the sex hormones testosterone and estradiol, but also of progesterone and its metabolite, allopregnanolone (AlloP). AlloP is a NAS that rapidly alters neuronal activity by being a potent allosteric activator of γ-aminobutyric acid A (GABAA) receptor, a chloride-conducting ligand-gated ion channel.3 By specific binding to distinct sites in the transmembrane domain (TMD) of GABAA receptors, AlloP causes an allosteric transition in the extracellular domain (ECD) of the receptor, leading to increased ligand (i.e., GABA) binding and, thereby enhanced chloride influx into cells.46 GABAA receptor activation thus hyperpolarizes the plasma membrane of neurons, thereby reducing the ability to generate action potentials.3,7 Mostly due to this non-genomic regulation of neuronal activity, AlloP and its formulations are used to treat patients with depression, and have been tested in clinical trials for treatment of insomnia, and anxiety disorders.3 This is because the balance between excitatory and inhibitory neuronal signaling is important for the regulation of mental health and sleep rhythm.8 AlloP and other NAS are also evaluated for clinical use to treat severe neurological and neurodegenerative diseases, such as epilepsy, major depressive disorder (MDD), multiple sclerosis (MS) as well as Parkinson’s disease (PD), and Alzheimer’s disease (AD).911 AlloP and related NAS are also considered potent regenerative agents to reduce neuroinflammation after injury or in the context of the above-mentioned neurodegenerative diseases.12 The potential of these NAS to reduce inflammation is at least partly a consequence of their action on microglia and astrocytes, either by binding to GABAA receptors in these cells, thereby inhibiting NFκB signaling, by regulating the activity of Toll-like receptors or by exerting control on autophagy.3,12,13 The ability of AlloP to stimulate repair could also become an important treatment option for MS.14 Mechanistically, AlloP can increase the activity of synaptic or extrasynaptic GABAA receptors leading to phasic or tonic inhibition of neuronal activity, respectively.3 Interestingly, epi-allopregnanolone (E-AlloP), a diastereomer of AlloP, inhibits rather than activates GABAA receptors and can reverse the stimulating effect of AlloP on GABAA receptor mediated chloride influx by binding to distinct sites in the receptor.4,6,15,16 To reach the different binding sites in the TMD of the various subunits of the GABAA receptor, NAS like AlloP and E-AlloP must first be inserted into the lipid bilayer.17 Since membrane-active steroid hormones can modulate membrane properties,18,19 it cannot be excluded that some of the effects of these two NAS are mediated by modulating bulk membrane properties, in addition to directly exerting control on receptor activity.1,17 Accordingly, the question arises, whether differing membrane interactions affect the distinct biological activities of AlloP and E-AlloP. Despite being of central importance, so far, only few studies have directly compared the membrane properties of neurosteroids.20,21 The impact of membrane interactions on function is a general feature of not only NAS but also a variety of other anesthetics and small molecules, whose absorption to the lipid bilayer can change the protein conformational energy landscape.2224 In fact, a recent hypothesis is that general anesthesia can impact ligand-gated ion channels, such as GABAA receptors, via their impact on the lateral pressure profile in the lipid bilayer.23,25 Theoretical studies suggest that this mechanism can even explain the complex electrophysiological behavior of the GABAA receptor in the presence of ligands and agonists.2224

Additionally, the availability of both NAS for GABAA receptor modulation might depend on their transport in cells, but almost nothing is known about the trafficking of AlloP and E-AlloP in neuronal cells. Since the effect of these NAS can last over hours and even days after single-dose application, intracellular sequestration and slow release from such depots could be possible. Studying intracellular transport of NAS depends on suitable analogs, which allow for tracking AlloP and E-AlloP in cells. One approach for that is to employ neurosteroid analogs tagged with an organic dye, such as nitrobenzoxadiazole (NBD) or a clickable Alexa488-azide, the latter linked via an extra alkyne group in the molecule.21,26,27 Dye-conjugated neurosteroids, however, will have very different physicochemical properties compared to the natural NAS, as found in a large body of work for the related sterol cholesterol (reviewed in Ref. 28). Even though these dye-tagged neurosteroid analogs might activate the GABAA receptor,27 their membrane properties and interaction with other lipid species can differ from natural NAS. Thus, the polarity and charge of the attached dye will compromise the precise action of NAS in the lipid bilayer, such that potential differences between stereoisomers of the same neurosteroid cannot be uncovered.

In this study, we present two new intrinsically fluorescent derivatives of AlloP (MQ-323) and E-AlloP (YX-11), which differ from their natural counterparts only by containing three conjugated double bonds in the ring system (Figure 1A and B). We and others have shown in previous studies that this approach of introducing extra double bonds into sterol molecules is very versatile and allows for studying sterol-protein interactions and intracellular trafficking for a variety of biologically relevant sterols without significantly compromising their activity.2830 Minimally modified sterols also allow for studying the impact of functional modifications of sterols on their membrane interaction, protein binding, or even intracellular transport, as we showed for analogs of cholesterol (i.e., cholestatrienol, CTL) versus oxysterol analogs, which differ from CTL only in having an extra hydroxy group in the side chain (i.e., 25- and 27-hydroxy-CTL).3134 We show here that the distinct properties of AlloP and E-AlloP for differential binding to and activation of the GABAA receptor are preserved among the new fluorescent neurosteroids. Furthermore, we provide evidence for similar membrane properties of MQ-323, the analog of AlloP, and YX-11, the analog of E-AlloP, but we also observe differences, such as stereospecific self-aggregation. Finally, we assess cellular uptake and trafficking of both analogs by UV-sensitive fluorescence microscopy and find quantitative differences in their subcellular transport. Our results demonstrate the potential of intrinsically fluorescent neurosteroid analogs for elucidating the molecular function and biological activity of NAS.

Figure 1:

Figure 1:

Allopregnanolone enantiomers and their fluorescent analogs can order phospholipid acyl chains. Chemical structures of allopregnanolone and epi-Allopregnalone (A) and their fluorescent analogs YX-11 and MQ-323 (B), respectively. The structure of allopregnanolone and epi-allopregnanolone differ from each other only by the orientation of the 3-hydroxyl group (highlighted in green and red, respectively). The fluorescent properties of the allopregnanolone and epi-allopregnanolone analogous, named YX-11 and MQ-323, are provided by the conjugated double bonds (highlighted in violet). Acyl chain order parameter calculated from MD trajectories for the sn-1 chain (palmitic acid, PA in panel C) and the sn-2 chain (oleic acid, OA in panel D) for membranes consisting of POPC and the indicated combinations of cholesterol (CHL) and neurosteroids.

Results

Membrane properties of AlloP, E-AlloP and their fluorescent analogs

AlloP and E-AlloP differ from each other only in the orientation of their 3-hydroxy group, and their analogs, MQ-323 and YX-11, show the same difference while containing three conjugated double bonds in the ring system, as shown in Figure 1A and B. Thus, the only difference between MQ-323 and YX-11 is the orientation of the 3-hydroxy group, exactly like the difference between AlloP and E-AlloP, allowing us to study the stereospecific effects of the NAS in a fluorescence setting. The parent molecule from which steroid hormones are synthesized, cholesterol, is well-known to condense fatty acyl chains in fluid lipid membranes, and it has been postulated that NAS can modulate membrane properties in a stereospecific manner.25,35 To address this point and validate our novel fluorescent steroid analogs, we used molecular dynamics (MD) simulations of AlloP, E-AlloP, as well as MQ-323 and YX-11, and determined the acyl chain order parameter for the host phospholipid POPC in cholesterol-containing membranes. We found that none of the NAS caused additional acyl chain ordering in cholesterol-containing membranes (Figure 1C and D). Additional simulations of the natural NAS, AlloP and E-AlloP, using higher steroid mole fractions (i.e., between 10 and up to 60 mol%) in POPC membranes in the absence of cholesterol showed that these neurosteroids could order the phospholipid acyl chain of POPC, though to a much lower degree than cholesterol (Figure S1). Also, under those conditions, their membrane orientation became more confined and similar to that of 30 mol% cholesterol, while their dynamics were dramatically slowed (Figure S2 and S3). For all conditions tested, except the highest sterol/steroid mole fraction of 0.6, cholesterol has the highest membrane condensing effect, as reflected by the reduction in area per lipid (Figure S3). Together, NAS can affect the properties of lipid bilayers but only at very high membrane concentrations, which are not observed under physiological conditions.

The differing orientation of the 3-hydroxy group in AlloP versus E-AlloP could affect the interaction of either NAS with membrane phospholipids. To test this hypothesis, we quantified the extent of tilting of the steroids from MD simulations of the bilayers at physiologically relevant concentrations, i.e., in membranes consisting of POPC with 25 mol% cholesterol and 5 mol% neurosteroid. We found that all four NAS studied have a rather broad tilt angle distribution in both membrane leaflets (Figure 2AD). This is in contrast to cholesterol, whose membrane orientation is much more confined around 15 degree tilt relative to the bilayer normal (Figure 2E). All sterols show excursions in their membrane orientation, which vary for cholesterol between 0 and max. 40 degrees, while the NAS can lie flat in the bilayer and even flip to the opposite membrane leaflet. We found a slightly more tilted orientation of AlloP and its fluorescent analog MQ-323 compared to E-AlloP and its analog YX-11. In contrast, both fluorescent neurosteroids resembled their natural counterparts closely (Figure 2AD). Thus, the orientation of the 3-hydroxy group has a slight effect on membrane anchoring, and a β-orientation, like in cholesterol, gives stronger interactions compared to an α-orientation, as in AlloP. To sustain this conclusion, we analyzed the hydrogen bonding pattern and reorientation dynamics of the four NAS from the MD simulations. We found that E-AlloP and YX-11 form slightly more hydrogen bonds to water and neighboring POPC molecules compared to AlloP and MQ-323, respectively (Figure 3AD). Furthermore, the larger tilting caused a much slower reorientation dynamics of AlloP and MQ-323 in the membrane compared to E-AlloP and YX-11 (Figure 3E). The fastest reorientation was found for cholesterol, which is a result of its upright orientation, allowing for unhindered rotation around its long molecular axis. The much larger tilting and slowed rotational dynamics of the four studied NAS compared to cholesterol is in line with our earlier observations with side chain oxysterols.32,34

Figure 2:

Figure 2:

Membrane localization and orientation of allopregnanolone enantiomers (A-B) and their fluorescent analogs (C-D) and cholesterol (E). Density heatmaps (2D-histograms) over the tilt angle and distance from the membrane center (z) are shown.

Figure 3:

Figure 3:

Hydrogen bonding capacity and membrane dynamics of allopregnanolone enantiomers and their fluorescent analogs. The number of hydrogen bonds formed between the allopregnanolone enantiomers and water (blue lines), the POPC carbonyl groups (orange lines), and cholesterol’s hydroxyl group (green lines) are shown for all neurosteroids (A-D). The black lines indicate the total number of hydrogen bonds formed by membrane-embedded neurosteroids. Membrane dynamics of neurosteroids were assessed in relation to that of cholesterol from the decay of the autocorrelation function as a function of time, ACF(t), as shown in (E). Panels (F-D) show snapshots from the MD simulations of MQ-323 and YX-11. The neurosteroid molecules are highlighted as spheres.

To follow up on these observations, we carried out experiments with the NAS in model membranes using 2H-NMR spectroscopy. Assessment of the acyl chain order parameter of POPC membranes containing either AlloP, E-AlloP, MQ-323, or YX-11 by NMR spectroscopy showed that none of these NAS can order phospholipid acyl chains to a significant extent (Figure 4A and B). This is, again, in stark contrast to cholesterol, while substituting some cholesterol with the NAS, also did not affect membrane ordering (Figure 4C). Based on these observations, we conclude that - at least at physiologically relevant concentrations (i.e., here 10 mol%) - neither AlloP nor E-AlloP or their fluorescent derivatives can cause structural ordering of lipid bilayers on the time scale accessible by NMR spectroscopy. There was a slight tendency of acyl chain ordering by 10 mol% of the NAS in POPC membranes as assessed by MD simulations (Figure S1). This minor discrepancy between NMR and MD results could be due to the different time scales accessible by both methods. Again, at un-physiologically high concentrations, we do observe membrane ordering by the NAS paralleled by less tilting and faster rotation around the long molecular axis (Figure S2A and B).

Figure 4:

Figure 4:

Order parameter of neurosteroids and their effect on membrane permeability to dithionite. The order parameter of the sn-1 chain of deuterated POPC in the indicated membranes was measured by 2H-NMR spectroscopy (A-C). Permeability of large unilamellar vesicles consisting of the indicated lipids and containing 0.5 mol% NBD-PC to sodium dithionite was measured by fluorescence spectroscopy (D). A bi-exponential fit to the data gives the rate constant for access of dithionite to the inner membrane leaflet.

While we did not see the opposite effect, i.e., a decreased ordering of acyl chains in the presence of NAS, the largely upright orientation of the neurosteroids could imply that they can reduce the membrane permeability barrier. To test this idea, we measured the permeation of dithionite across lipid membranes labeled with the fluorescence phospholipid NBD-PC. Reduction of the fluorescence of NBD-PC by dithionite is biphasic with a fast phase resembling the chemical reduction step and a slow phase resembling dithionite permeation and/or flip-flop of the NBD-PC across the membrane.36 This assay has been widely used to assess membrane perturbation by drug molecules,37,38 but also to determine the impact of structural modifications of cholesterol on membrane permeability.3942 Using this assay, we found that YX-11, and to a lower extent E-AlloP, reduced membrane permeability slightly (Figure 4D). This effect was limited to POPC membranes, while in POPC/cholesterol membranes, the permeability for dithionite was reduced irrespective of the added NAS (Figure 4D). The latter is a consequence of the presence of cholesterol, which is known to reduce the permeability of liposomes for dithionite.

Neurosteroid analogs bind to the ELICα1GABAA receptor and regulate its activity in a stereo-specific manner

Having ruled out major differences in membrane properties of the neurosteroid epimers, we assessed next their interaction with GABAA receptors. While AlloP is known as a positive allosteric modulator of GABAA receptors, Epi-AlloP does not bind to the same site as AlloP, and it is an important property for fluorescent analogs of both steroids to show the same stereospecific binding pattern.1 A particular advantage of intrinsically fluorescent sterols and steroids is that their excitation spectra significantly overlap with the emission spectra of aromatic amino acids, allowing for direct measurement of binding by Förster resonance energy transfer (FRET). This approach has been widely used to study protein-sterol interactions, e.g.,33,43,44 and we recently also applied it to two fluorescent neurosteroid analogs.45 To determine the affinity of the new fluorescent neurosteroids, purified ELICα1GABAA receptor was first incubated with increasing concentrations of the AlloP analog MQ-323. FRET was measured by exciting the donor fluorescence (i.e., that of the protein at 280 nm) and recording the emission of MQ-323 (Figure 5A). Due to energy transfer, the sensitized emission of MQ-323 increases upon binding to the receptor in a saturable manner (Figure 5B). Since FRET efficiency is a function of proximity and mutual orientation of donor and acceptor dipoles, not only specific binding but also non-specific association of MQ-323 or its insertion into the detergent micelle of the reconstituted receptor could result in a FRET signal. To account for those contributions, the binding experiment was repeated in the presence of 30 μM AlloP, and the resulting FRET signal was subtracted from the determined binding curve (Figure 5B). As a further test for binding specificity, the experiments were repeated with a mutant form of ELICα1GABAA receptor, in which Trp246, in transmembrane helix 1 of the receptor, was mutated to leucine (W246L). The Trp246 residue is essential for the binding of AlloP (analogs) in the canonical intersubunit site on GABAA receptors and for the potentiation of GABA-elicited currents by AlloP.4,4648 In the W246L mutant receptor, the FRET intensity of MQ-323 was indistinguishable from the non-specific FRET signal observed in the wild-type ELICα1GABAA receptor, supporting that MQ-323 specifically binds at the previously identified site involving transmembrane helix 1 of the receptor.4,45 In contrast to MQ-323, adding YX-11 to the wild-type ELICα1GABAA receptor did not result in any FRET signal, clearly showing that the fluorescent analog of E-AlloP (like E-AlloP) does not bind to the canonical neurosteroid site on the receptor (Figure 5C).4 Thus, the new intrinsically fluorescent analogs of AlloP and its epimer are powerful tools for detecting stereospecific interactions with the GABAA receptors.

Figure 5:

Figure 5:

Förster resonance energy transfer (FRET) of neurosteroid binding to GABAa receptors. Purified ELICα1GABAA receptor was incubated with either MQ-323 or YX-11 dissolved in a buffer containing 0.015% DDM, and FRET from the protein to the respective NAS was measured upon excitation at 280 nm. Raw emission spectra (A) and normalized FRET intensities (B, n = 3 for each plot) for increasing concentrations of MQ-323. Nonspecific binding in (B) is the FRET signal obtained in the presence of 30 μM allopregnanolone, and W246L shows that the FRET is reduced to non-specific binding when W246 in the receptor is replaced by leucine. Comparison of total and non-specific FRET obtained with MQ-323 and YX-11 (C), showing that YX-11 does not bind to the ELICα1GABAA receptor.

Next, we asked whether the stereospecific binding of the novel neurosteroid analogs also contributes to allosteric regulation of the GABAA receptor. We found that MQ-323 potentiates the human α1β3γ2L GABAA receptor activated by low (<EC10) GABA strongly, while YX-11 had a very weak potentiating effect on receptor function (Figure 6A and B). When the receptors were maximally activated in the presence of saturating GABA and 50 μM propofol, YX-11 weakly inhibited receptor function (Figure 6C). Together, these results demonstrate that the new intrinsically fluorescent analogs of AlloP and its epimer, E-AlloP, maintain the ability of endogenous NAS to modulate GABAA receptor activity. This makes these probes highly suitable for functional studies of neurosteroid function in the future.

Figure 6:

Figure 6:

Electrophysiological measurements of GABAA receptor currents. The α1β3γ2L receptors were activated by 2 μM GABA (A, B) or 1 mM GABA + 50 μM propofol (C), and modulated by 10 μM MQ-323 or YX-11. Drug applications are shown by horizontal lines above current traces. The dot plots show steroid effects in % of control response (100% = no effect) calculated as ratios of current amplitudes after and before application of steroid.

Neurosteroid analogs show stereo-specific solubility in aqueous solution

Based on their intrinsic fluorescence, the analogs of AlloP and E-AlloP, should be suitable for live-cell imaging studies, so we aimed for labeling human astrocytes with either MQ-323 or YX-11. We found that diluting both probes into the culture medium resulted in poor labeling but gave bright extracellular aggregates, particularly in the case of YX-11 (Figure S4). Aggregates show delayed photobleaching, which, besides their bright fluorescence, can be used to identify them in the microscope. To follow up on these observations, we studied the emission spectra of the fluorescent neurosteroids in water and found that YX-11 but not MQ-323 shows a pronounced red-shift and change in spectral shape when the steroid concentration exceeds ca. 9 μM (Figure 7A and B). By plotting the maximal emission at 378 nm for MQ-323 and 416 nm for YX-11 as a function of steroid concentration and using a bi-linear regression, a breakpoint at 8.99 μM was found for YX-11, which likely resembles the critical aggregation concentration (CAC) for this steroid in water (Figure 7C). As for MQ-323, no breakpoint was observed, and the emission spectra had the same shape over the entire studied concentration range (Figure 7B). The integrated emission of YX-11 and MQ-323 is similar in most solvents, and the red-shifted emission of YX-11 was only observed in water and PBS (Figure S5). These results suggest that the self-aggregation of the fluorescent steroids is stereospecific. In an attempt to uncover the molecular mechanisms underlying this stereospecific aggregation, we carried out MD simulations of YX-11 and MQ-323 monomers in water/ethanol mixtures at such high concentrations (0.1 M), that aggregate formation can be followed in a few μsec, achievable by atomistic molecular simulations. Under those conditions, the formation of aggregates was readily observed for both steroid probes once the ethanol content dropped below ca. 50% (v/v) (Figure S6). Thus, both neurosteroids can form aggregates in aqueous solution. To test whether self-aggregation might also take place in membranes, both NAS analogs were incorporated in large unilamellar vesicles (LUVs), and fluorescence spectra were recorded. The characteristic red-shift in emission was found for YX-11 but not for MQ-323, even at 20 mol% steroid probe in the membrane, but it was only observed for YX-11 at a high total lipid concentration of 200 μM (Figure S7).

Figure 7:

Figure 7:

Fluorescence emission spectra of YX-11 and MQ-323 titration in water. (A+B) Normalized fluorescence emission spectra (Ex. 328nm) of YX-11 and MQ-323 with various concentrations (0 to 50 μM) in water. The spectra represent the average from three independent measurements and are normalized to the maximum peak intensity. (C) The mean of the maximum emission at 416 nm and 378 nm from three individual experiments of YX-11 (orange) and MQ-323 (blue), respectively, plotted as a function of neurosteroid concentration and fitted with a straight-line model with breakpoints using a piece-wise linear regression routine.

Loading of AlloP onto cyclodextrin has been shown to improve its bioavailability in drug formulations10 and we, therefore, mixed MQ-323 and YX11 with low concentrations of MβCD to prevent self-aggregation of the neurosteroids prior to cell labeling. We found that the emission spectrum of YX-11 above its CAC in PBS/cyclodextrin solutions became identical to that below its CAC in PBS (Figure S8) and was indistinguishable from that of MQ-323 (Figure S8B). In line with these observations, we observed by UV-sensitive fluorescence microscopy more intense and homogeneous labeling of human astrocytes by both fluorescent NAS probes upon loading them onto cyclodextrin (Figure 8) but not when loading them onto albumin (not shown). Strikingly, the uptake of YX-11 was higher than that of the bioactive AlloP analog, MQ-323, when comparing cell-associated intensities of both probes with autofluorescence (Figure 8). We chose human astrocytes for these experiments, as these glia cells are very important in GABAergic signaling and use neurosteroids in their communication with neurons.49

Figure 8:

Figure 8:

Uptake of neurosteroid analogs loaded on methyl-β-cyclodextrin (MCβD) by astrocytes. Human astrocytes were loaded with either YX-11 and MQ-323 on MCβD (final concentration of 20 μM) and incubated for 1 hr before imaging on a UV-sensitive wide-field microscope (A-B). Scale bar, 20 μm. Cells were segmented using Cellpose (see Image analysis for fluorescence microscopy data) in order to find the mean UV fluorescence intensity for cells loaded with YX-11 (n = 580) or MQ-323 (n = 488) (C). Control in panel C are cells without loading of neurosteriods (n = 290). By uing a Mann-Whitney U test on the distributions for YX-11 and MQ-323 in panel C, a p-value of 9.510 × 10−32 was found.

Neurosteroid analogs accumulate in lipid-droplets in astrocytes

To determine the trafficking itineraries of our novel fluorescent NAS analogs, we carried out co-staining experiments with known organelle markers. NAS were administered in MβCD and allowed to equilibrate for 30 minutes before loading to cells. Both YX-11 and MQ-323 showed little - if any - co-localization with NBD-Ceramide, a marker for the Golgi apparatus and trans-Golgi network (TGN, Figure S9). Co-staining with the Golgi marker was slightly higher for YX-11 compared to MQ-323, but this could also be due to the increased uptake and, therefore, higher signal for the E-AlloP analog. Both NAS probes became enriched in lipid droplets co-stained with the droplet marker BODIPY 493/503 (Figure 9). Importantly, the extent of this droplet targeting was much more pronounced for YX-11 compared to MQ-323, as validated by image quantification (Figure 9). No colocalization was found with a marker for late endosomes and lysosomes (LE/LYSs), ruling out that any of the punctuate stainings of the fluorescent neurosteroids are lysosomal (Figure S10). Similarly, the staining of the PM was rather low compared to the intracellular accumulation of both NAS probes. We conclude that fluorescent analogs of AlloP and E-AlloP are transported to lipid droplets and that the extent of this intracellular accumulation is stereospecific. We repeated the uptake experiment with an ACAT inhibitor (Avasimibe) and found a strongly reduced formation of lipid droplets and consequently no or very little droplet targeting of YX-11 and MQ-323 to the few remaining droplets, despite a slightly higher uptake of YX-11 by the cells upon ACAT inhibition (Figure S11). Thus, droplet formation requires ACAT activity. A hypothesis for the preferred droplet targeting of YX-11 is that this E-AlloP analog is preferentially esterified by ACAT and therefore enriched in lipid droplets. Supporting this hypothesis are studies carried out for cholesterol and its 3’-hydroxy α-epimer, epi-cholesterol, which showed that the β-orientation of the 3’-hydroxy group, as in cholesterol, is necessary for sterol esterification.51,52 Thus, it is possible that YX-11 with its 3’-hydroxy group in β-orientation is recognized by ACAT and therefore esterified and stored in droplets, while MQ-323 is no ACAT substrate. Unfortunately, the fluorescence of the neurosteroid probes in the very few remaining droplets after ACAT inhibition is too low to test this model directly. Together, our results show that transport of the fluorescent neurosteroids to droplets requires ACAT activity, and they suggest that the stereospecific accumulation of allopregnanolone epimers is at least partly the result of different esterification e iciency of fluorescent analogs of AlloP and E-AlloP.

Figure 9:

Figure 9:

Neurosteriods are localized in lipid droplets (LDs). Human astrocytes were loaded with either YX-11 or MQ-323 on MCβD (final concentration of 20 μM), incubated for 1 hr and afterward stained with bodipy for 5 min before imaging on a UV-sensitive wide-field microscope (A-B). Scale bar, 20 μm. Cells and LDs were segmented as described in Image analysis for fluorescence microscopy datas, and the fraction of YX-11 (n = 290) or MQ-323 (n = 223) found inside LDs per cell was quantified (C). A fraction above 1 indicates the upregulation of neurosteroids in LDs. By using a Mann-Whitney U test on the distributions in panel C, a p-value of 7.66 × 10−30 was found.

Discussion

NAS, like AlloP, are potent regulators of mood, stress handling, and sleep rhythm due to their ability to potentiate the effect of GABA on the GABAA-mediated chloride currents. The effects of these NAS are site-specific and stereo-specific, largely due to specific receptor-ligand interactions. Dissecting such interactions at a molecular level requires suitable NAS probes, which mimic endogenous neurosteroids closely but preserve the key pharmacological characteristics of each steroid analog. Here, we present two novel analogs of AlloP and its 3β-epimer, E-AlloP, which can be used to study not only neurosteroid binding to specific sites in the GABAA receptor by an on-line fluorescence spectroscopy assay, but also to directly observe intracellular trafficking of NAS by UV-sensitive fluorescence microscopy. We show that the membrane properties of MQ-323, the fluorescent derivative of AlloP, and of YX-11, the analog of E-AlloP, are very similar at low steroid concentrations, which is in line with previous studies.1,20 MD simulations and permeation studies with dithionite indicate that E-AlloP and its analog YX-11 are slightly better anchored in lipid membranes compared to AlloP and its analog MQ-323, suggesting that the β-orientation of the 3-hydroxy group, as also found for cholesterol, is important for hydrogen bonding of sterols and their interaction with phospholipids. However, the difference is minor, and the capacity to order fatty acyl chains is negligible for all studied NAS, in strong contrast to cholesterol. Since neurosteroids have a much shorter side chain compared to cholesterol, the alkyl chain of cholesterol is essential for its strong membrane condensing effect, in line with previous studies.40 Unexpectedly, we found that YX-11 self-aggregates at much lower concentrations than MQ-323 in aqueous solutions and likely also in membranes. Whether aggregates are inserted into the lipid bilayer or directly formed once the E-AlloP analog accumulates in the membrane remains to be shown. Similarly, the molecular mechanisms underlying stereospecific aggregation of AlloP analogs need to be determined.

Our new intrinsically fluorescent NAS not only allow for studying properties of neurosteroids in solutions and in membranes, but also enabled us to infer stereospecific interactions with GABAA receptors. We demonstrate that the stereospecific interaction and regulation of the ELICα1GABAA receptor by AlloP is preserved for its fluorescent analog MQ-323, while the analog of E-AlloP cannot bind or activate the receptor to a comparable extent. Thus, due to their minimal chemical modification compared to the natural NAS epimers, MQ-323 and YX-11 are sutiable probes for studying stereospecific allosteric regulation of GABAA receptors by combined FRET spectroscopy and electrophysicology.

When loading the fluorescent NAS onto cyclodextrin, their aqueous solubility was comparable, and efficient labeling of human astrocytes could be achieved. This underlines the importance of proper formulations for neurosteroid targeting to cells.10 It also supports and explains previous findings showing that the slow action of NAS on GABAA receptors can be accelerated with cyclodextrin:53 based on our imaging results, we conclude that increased retention of NAS in cells, when loaded onto cyclodextrin, results in a larger cellular pool being available for receptor activation. Thus, our findings support a model of membrane-mediated access of AlloP to the GABAA receptor, which is in full support of the lipophilic nature of the NAS.26 While we could load the NAS probes onto albumin, labeling of cells with steroid-albumin complexes was inefficient (not shown). Future studies could attempt cellular delivery by other proteins, such as steroid carrier proteins, which have been proposed to be key regulators of trafficking and activity of NAS.2 Using multi-color live-cell microscopy, we could show that human astrocytes internalize more of the E-AlloP analog, YX-11, compared to the AlloP analog, MQ-323. This cellular enrichment with the E-AlloP analog is largely due to its preferred transport to lipid droplets.

Our results are in line with previous findings, showing that (analogs of) NAS accumulate at intracellular sites,26,53,54 and extend these observations by showing that the intracellular pool is primarily consisting of lipid droplets and to a smaller extent the Golgi/TGN and likely the endoplasmic reticulum (ER). Accumulation of a clickable analog of AlloP in the Golgi has been shown previously in neurons.27 Based on these findings, a model can be envisioned, according to which AlloP and other NAS are continuously supplied to the plasma membrane (PM) from lipid droplets, either directly by non-vesicular trafficking or by secretory membrane traffic via the Golgi/TGN. Future studies using MQ-323 and YX-11 in combination with genetic or chemical inhibition of selected trafficking pathways will provide further insight into the transport pathways of NAS in cells.

Materials and Methods

Computational details

Molecular dynamics simulations

Membrane bilayers containing in each leaflet 70 POPC lipids, 25 cholesterol molecules, and 5 neurosteroid molecules were assembled using packmol. Membranes were assembled for allopregnanolone, epiallopregnanolone, and their fluorescent analogs. Additionally, a membrane with a 70:30 ratio of POPC:cholesterol was generated for comparison purposes. The systems also included 10,000 water molecules, 22 K+, and 22 Cl ions, corresponding to a concentration of about 120 mM KCl. The water molecules were described by the TIP3P force field,55 while the POPC lipids and cholesterol were described by the amber lipid14 force field.56 Force field parameters for the neurosteroids were derived using the QForce57 program package, based on r2scan-3c58 Hessian calculations carried out using the Orca program,59 version 5.0.3. Charges were derived using electrostatic potential fitting. Lennard-Jones parameters were assigned from GAFF.60 Flexible dihedral parameters were fitted to relaxed dihedral scans, also using r2scan-3c.

All molecular dynamics simulations were carried out using the Gromacs program,61 version 2023.1. The membranes were equilibrated in a three-stage procedure. First, the structures were minimized for 5000 steps with a steepest descent minimizer to remove any bad contacts. Next, we carried out NVT equilibration (200 ps), and NPT equilibration (2 ns) was carried out. A time-step of 2 fs was used. Long-range electrostatics were treated with the particle mesh Ewald method62 with a cutoff distance of 12 Å. Bonds involving hydrogens were constrained using LINCS.63 The temperature was controlled by a Nose-Hoover thermostat64,65 (towards 298.15K), while the pressure (in the NPT simulations only) was controlled by a semi-isotopic Berendsen barostat66 (towards 1 bar). After equilibration, production molecular dynamics runs (1000 ns) were carried out. In the production simulations, we switched to a Parrinello-Rahman barostat.67 The resulting trajectories were analyzed using Gromacs (for deuterium order parameters), and in-house Python scripts relying on the MDAnalysis library68 for tilt angle distributions, hydrogen bonding, and autocorrelations of the tilting angle.

Aggregation simulations of MQ-323 and YX-11 were carried out in ethanol/water solvent mixtures of varying solvent composition (0%, 20%, 40%, 60%, 80%, and 100% ethanol concentration by volume). In each simulation, 100 neurosteroid molecules were solvated in a volume of about 1570 nm3, corresponding to a neurosteroid concentration of about 0.1M. Force-field parameters were assigned as described above, with parameters for ethanol being derived using Q-Force based on r2scan-3c QM calculations. Solvent/neurosteroid mixtures were assembled using the Packmol program, with the neurosteroids were randomly distributed in the simulation cell. The assembled systems were equilibrated using the same protocol as described above. After this, production molecular dynamics runs (200 ns) were carried out, and the resulting trajectories were analyzed for aggregation.

Experimental procedures

Reagents

1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC #850457P) and cholesterol (#700000P) were purchased from Avanti Polar Lipids, Inc. (Alabaster, Al). Bodipy 493/503(#D3922), NBD-C6-Ceramide (#N1154, NBD-Cer) and Dextran, Tetramethylrhodamine (#D1819) was purchased from ThermoFisher. Methyl-β-cyclodextrin.(MCβD)(#332615) was purchased from Sigma-Aldrich. Immortalized human astrocytes (IM-NHA) (#P10251) and the acquired growth media (astrocyte medium kit (#P60101)) were purchased from Innoprot.

Preparation of large unilamellar vesicles (LUVs)

LUVs were made by extrusion using a 400nm pore diameter filter and a micro extruder from Avanti Polar Lipids (Alabaster, AL), as described in.69 LUVs were made with different lipid compositions: POPC: neurosteroids (80:20mol%) and POPC: cholesterol: neurosteroids (70:20:10mol%), and made with a final lipid concentration of either 50 μM or 200 μM corresponding to 100mol%).

Assessment of membrane permeability

Access of dithionite to the phospholipid analog NBD-PC was measured in LUVs exactly as described.38 Briefly, the fluorescence of LUVs of the respective lipid composition with 1mol% NBD-PC was continuously monitored for an excitation wavelength of 470 nm and an emission wavelength of 540 nm. Sodium dithionite was added from a freshly prepared 1 M stock solution in 100 mM Tris (pH 10) giving a final dithionite concentration of 50 mM. After monitoring the reduction kinetics, Triton X-100 (final concentration of 0.5% (w/v) was added, resulting in disruption of the liposomes and full access of the quencher to the lipid analog. After normalization to the intital intensity, the decay curves were fit with a bi-exponential model, and from the rate constant of the second phase, the kinetics of access of dithionite to NBD-PC in the inner membrane leaflet was inferred.

Fluorescence spectroscopy

Emission spectra were obtained using an ISS Chronos BH spectrofluorometer (Urbana-Champaign, IL). The excitation wavelength was set to 328 nm, and recorded from 350–500 nm. For all measurements, a slit width of 0.5 mm (excitation and emission path) with no polarization was used.

Photo-physical properties of YX-11 and MQ-323

For each experiment, a fresh stock of neurosteroids with a final concentration of 2.5 mM in absolute ethanol was prepared. Emission spectra were obtained in various solvents with different polarities with a final concentration of 20 μM. Each spectrum represents the average from three independent measurements. The averaged spectra were smoothed using the Savitzky-Golay method in Python. The integrated intensity (area under the curve) was calculated for the raw spectra and displayed as the mean with its correlation standard error (sem). The titration experiments were conducted in water with a final concentration ranging from 0–50 μM. The emission maximum of the raw spectra for YX-11 was found at 416 nm and for MQ-323 at 378 nm. The maximum emission from three individual experiments was plotted as a function of neurosteroid concentration and fitted with a straight-line model with breakpoints using piece-wise regression in Python.

FRET measurements of MQ-323 and YX-11 binding to ELICα1GABAA receptor

Binding of MQ-323 and YX-11 to ELIC- 1GABAAR was measured as previously described.45 Briefly, 100 μl of ELICα1GABAA receptor (0.3 μM) was transferred into a 50 μl window quartz cuvette, which was maintained at 4°C. NaI (final concentration of 100 mM) was added to the protein sample to reduce the emission of hydrophilic tryptophan residues. 1 μL aliquots of 100X stock solutions of MQ-323 or YX-11 (prepared in Buffer A + 0.015% DDM) were sequentially added to the protein sample and incubated for 30 min after each addition. The emission spectra (background spectrum) of MQ-323 or YX-11 in the absence of protein were obtained and subtracted from the protein-MQ-323 (or YX-11) emission spectra to obtain the background-corrected spectra. Finally, the contribution of ELICα1GABAA receptor tryptophan emission to the FRET spectra was calculated as previously described45 and subtracted from the background-corrected spectrum to obtain the emission spectrum of the signal resulting from energy transfer (FRET spectrum) from tryptophan to MQ-323 or YX-11. The 370 nm intensity values of the FRET spectra were plotted against MQ-323 concentration to yield the MQ-323 binding curve. Non-specific MQ-323 binding was measured by determining MQ-323 binding (as described above) in the presence of 30 μM allopregnanolone and specific binding was obtained by subtracting non-specific binding from total binding. Data is presented as mean ± S.D. from three independent experiments.

Electrophysiology

The electrophysiological experiments were conducted on human α1β3γ2L GABAA receptors. Receptor expression in Xenopus laevis oocytes and details of two-electrode voltage clamp have been described in detail previously.4,70 The modulatory effects of MQ-323 and YX-11 were tested on receptors activated by 2 μM (<EC10) GABA. In addition, YX-11 was tested on receptors activated by the combination of 1 mM GABA + 50 μM propofol. This drug combination fully activates the receptors.70 The drug application protocol consisted of activating the receptors with GABA (or GABA + propofol), followed by co-application of steroid, and wash in GABA (or GABA + propofol). The effect of steroid was calculated from the ratio of current amplitudes in the presence and absence of steroid.

NMR spectroscopy

The desired molar ratio of respective molecules was first mixed in ethanol. For overnight lyophilization the solvent was evaporated and the samples were redissolved in cyclohexane. The obtained fluffy powder was were hydrated with 50 wt% H2O-buffer (10 mM Hepes, 100 mM NaCl, pH 7.4) and equilibrated by ten freeze-thaw cycles. The static 2H NMR spectra were acquired on a Bruker (Bruker Biospin, Rheinstetten, Germany) DRX300 NMR spectrometer using the quadrupole echo pulse sequence71 on a high-power probe with a 5-mm solenoid sample coil. The relaxation delay was 1 s, the delays between the 90° pulses (of ~ 3.2 μs) were 50 μs. For the depaked spectra72 smoothed order parameter profiles were calculated.73 The measurements were carried out at a temperature of 25°C.

Cell culture and labeling

Immortalized human astrocytes (IM-NHA) were cultured in T25 Nunc EasyFlask (#156367, Thermo Scientific) and grown in astrocyte medium (AM) supplemented with 10% Fetal Bovine Serum (FBS) and 1% pencilin/streptomycin solution (P/S solution) and incubated at 37°C, 5% CO2, and 95% humidity. Media was changed every second day and subcultured at 90% confluence using 1X trypsin-EDTA (#T47174, Sigma-Aldrich). For microscopy cells were plated on 35mm microscope dishes (#TKO-P351184–408, MatTek) and ready for imaging when they reached 70–80% confluence.

To prevent aggregation of the neurosteroid analogs they were loaded on methyl-β-cyclodextrin (MβCD) in PBS (final concentration of YX-11 and MQ-323 (20 μM) and MβCD (40 μM)). The MβCD-sterol complex was thoroughly vortexed and settled for minimum 30 minutes at RT before loading to cells in Medium 1 (M1) (containing 150 mM NaCl, 5 mM KCl, 1 mM CaCl2, 1 mM MgCl2, 5 mM glucose and 20 mM HEPES at pH 7.4) for 1 hr at 37°C. For labelling of lipid droplets bodipy (final concentration of 0.2 μg/mL) was added to cells for 5 minutes at 37°C. For Golgi and TGN labelling cells were stained with NBD-Cer (final concentration of 5 μM) for 10 minutes at 37°C. For Labelling late endosomes and lysosomes (LE/LYSs) cells were stanied with 0.25 mg/mL Rh-dextran overnight. Before imaging, cells were washed with M1 media.

Fluorescence microscopy

Widefield epifluorescence microscopy was carried out on a Leica DMIRBE microscope with a 63 × 1.4 NA oil immersion objective (Leica Lasertechnik GmbH) with a Lambda SC smart shutter (Sutter Instrument Company) as illumination control. Images were acquired with an Andor Ixon blue EMCCD camera operated at −75°C and driven by a Solis software. The microscope contains an 10× extra magnification lens in the emission light path, resulting in a final pixel size of 193 nm for the 63× objective. YX-11 and MQ-323 were imaged in the UV range of the spectra using a specially designed filter cube obtained from Chroma Technology (Corp., Brattleboro, VA, USA) with a 335-nm (20-nm bandpass) excitation filter, 365-nm dichromatic mirror, and 405-nm (40-nm bandpass) emission filter. For the neurosteriods a bleach stack with 100 frames were recorded with 400 msec acquisition time. Bodipy was imaged using a fluorescein filter cube with 480 nm (40 nm bandpass) excitations filter, 505 nm longpass dichromatic mirror, and 527 nm (30 nm bandpass) emission filter.

Image analysis for fluorescence microscopy data

Image correction and deconvolution

Autofluorescence correction is done for all UV images for the neurosteriods by subtracting the last image of each bleach stack from the first one in an automated manner using batch processing in Python. The resulting image only contain the bleaching fluorescent probe signal. For visualization purposes images obtained for the lipid droplets, Golgi and TGN, and Rd-dextran (green and red emission) were deconvolved in an automated manner using batch processing in Python. The Richard-Lucy algorithm was used with 20 iterations and a theoretical PSF for each specific channel was used for deconvolution and generated using the Diffraction PSF 3D plugin in ImageJ (https://imagej.net/plugins/diffraction-psf-3d) Analysis of photobleaching kinetics was carried out using the ImageJ plugin PixBleach, as described.74

Segmentation of cells using Cellpose

Cells were segmented using the deep learning-based segmentation method Cellpose, as described in.75 The images were segmented using the built-in Cytoplasm 2.0 model (‘cyto2’) in Cellpose with a diameter of 100 in an automated manner in Python. The mean fluorescence intensity of the neurosteriods per cell was found using the masked area. To quantify the fluorescence intensity of the neurosteriods in lipid droplets (LDs), the LDs were segmented using a custom training pipeline. A ratio of how much of the neurosteroids found in LDs per cell was displayed in boxplots. A ratio above 1 indicates that more of neurosteriods is found in LDs than outside.

Supplementary Material

Supplementary material
1

Acknowledgments

This research was funded by the Lundbeck Foundation grant nr. R366-2021-226 (D.W.), by the Deutsche Forschungsgemeinschaft (DFG), MU 1017/14-1 (P.M. and D.W. as Mercator Fellow), from the NIH/NIGMS R35GM140947 (G.A.) and R35GM149287 (A.S.E.) and grant NIH 1 P50 MH122379 (D.F.C.), and all funding sources are kindly acknowledged. We also thank Mr. Spencer Pierce for help with electrophysiology experiments.

Footnotes

Author Competing Interests

The authors declare no competing financial interest.

Supporting Information

Molecular dynamics simulations of order parameter of POPC, of orientation and dynamics of neurosteroids as well as of membrane area as function of steroid concentration; photobleaching microscopy of YX-11 aggregates in cell culture medium, emission spectra and integrated fluorescence of the neurosteroid analogs in various solvents; molecular simulations of aggregation of fluorescent neurosteroids; emission spectra of neurosteroids in lipid membranes; effect of cyclodextrin on emission spectra of YX-11; co-localization microcopy experiment of fluorescent neurosteroids with Golgi marker and with marker for late endosomes/lysosomes, respectively; effect of ACAT inhibitor avasimibe on uptake and trafficking of fluorescent neurosteroids.

References

  • (1).Covey DF; Evers AS; Izumi Y; Maguire JL; Mennerick SJ; Zorumski CF Neurosteroid enantiomers as potentially novel neurotherapeutics. Neuroscience & Biobehavioral Reviews 2023, 149, 105191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (2).Diotel N; Charlier TD; Lefebvre d’Hellencourt C; Couret D; Trudeau VL; Nicolau JC; Meilhac O; Kah O; Pellegrini E Steroid Transport, Local Synthesis, and Signaling within the Brain: Roles in Neurogenesis, Neuroprotection, and Sexual Behaviors. Frontiers in Neuroscience 2018, 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Maguire JL; Mennerick S Neurosteroids: mechanistic considerations and clinical prospects. Neuropsychopharmacology 2023, 49, 73–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (4).Sugasawa Y; Cheng WW; Bracamontes JR; Chen Z-W; Wang L; Germann AL; Pierce SR; Senneff TC; Krishnan K; Reichert DE et al. Site-specific effects of neurosteroids on GABAA receptor activation and desensitization. eLife 2020, 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (5).Sun C; Zhu H; Clark S; Gouaux E Cryo-EM structures reveal native GABAA receptor assemblies and pharmacology. Nature 2023, 622, 195–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).Tateiwa H; Chintala SM; Chen Z; Wang L; Amtashar F; Bracamontes J; Germann AL; Pierce SR; Covey DF; Akk G et al. The Mechanism of Enantioselective Neurosteroid Actions on GABAA Receptors. Biomolecules 2023, 13, 341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Cutler AJ; Mattingly GW; Maletic V Understanding the mechanism of action and clinical effects of neuroactive steroids and GABAergic compounds in major depressive disorder. Translational Psychiatry 2023, 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Selten M; van Bokhoven H; Nadif Kasri N Inhibitory control of the excitatory/inhibitory balance in psychiatric disorders. F1000Research 2018, 7, 23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).Brinton RD Neurosteroids as regenerative agents in the brain: therapeutic implications. Nature Reviews Endocrinology 2013, 9, 241–250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).Irwin RW; Solinsky CM; Brinton RD Frontiers in therapeutic development of allopregnanolone for Alzheimerâ€s disease and other neurological disorders. Frontiers in Cellular Neuroscience 2014, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (11).Bryson A; Reid C; Petrou S Fundamental Neurochemistry Review: GABAA receptor neurotransmission and epilepsy: Principles, disease mechanisms and pharmacotherapy. Journal of Neurochemistry 2023, 165, 6–28. [DOI] [PubMed] [Google Scholar]
  • (12).Yilmaz C; Karali K; Fodelianaki G; Gravanis A; Chavakis T; Charalampopoulos I; Alexaki VI Neurosteroids as regulators of neuroinflammation. Frontiers in Neuroendocrinology 2019, 55, 100788. [DOI] [PubMed] [Google Scholar]
  • (13).Kim HN; Lee S-J; Koh J-Y The neurosteroids, allopregnanolone and progesterone, induce autophagy in cultured astrocytes. Neurochemistry International 2012, 60, 125–133. [DOI] [PubMed] [Google Scholar]
  • (14).Noorbakhsh F; Baker GB; Power C Allopregnanolone and neuroinflammation: a focus on multiple sclerosis. Frontiers in Cellular Neuroscience 2014, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Lundgren P; Strömberg J; Bäckström T; Wang M Allopregnanolone-stimulated GABA-mediated chloride ion flux is inhibited by 3-hydroxy-5-pregnan-20-one (isoallopregnanolone). Brain Research 2003, 982, 45–53. [DOI] [PubMed] [Google Scholar]
  • (16).Legesse DH; Fan C; Teng J; Zhuang Y; Howard RJ; Noviello CM; Lindahl E; Hibbs RE Structural insights into opposing actions of neurosteroids on GABAA receptors. Nature Communications 2023, 14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).Crowley J; Withana M; Deplazes E The interaction of steroids with phospholipid bilayers and membranes. Biophysical Reviews 2021, 14, 163–179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Alsop RJ; Khondker A; Hub JS; Rheinstädter MC The Lipid Bilayer Provides a Site for Cortisone Crystallization at High Cortisone Concentrations. Scientific Reports 2016, 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Atkovska K; Klingler J; Oberwinkler J; Keller S; Hub JS Rationalizing Steroid Interactions with Lipid Membranes: Conformations, Partitioning, and Kinetics. ACS Central Science 2018, 4, 1155–1165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Alakoskela J-M; Covey DF; Kinnunen PK Lack of enantiomeric specificity in the effects of anesthetic steroids on lipid bilayers. Biochimica et Biophysica Acta (BBA) - Biomembranes 2007, 1768, 131–145. [DOI] [PubMed] [Google Scholar]
  • (21).Mennerick S; Lamberta M; Shu H-J; Hogins J; Wang C; Covey DF; Eisenman LN; Zorumski CF Effects on Membrane Capacitance of Steroids with Antagonist Properties at GABAA Receptors. Biophysical Journal 2008, 95, 176–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Cantor RS; Twyman KS; Milutinovic PS; Haseneder R A kinetic model of ion channel electrophysiology: bilayer-mediated effects of agonists and anesthetics on protein conformational transitions. Soft Matter 2009, 5, 3266. [Google Scholar]
  • (23).Cantor RS Chemical and Biochemical Approaches for the Study of Anesthetic Function, Part A; Elsevier, 2018; p 97–110. [Google Scholar]
  • (24).Cantor RS Kinetic Model of Adsorption of Aqueous Solutes onto Lipid Bilayers: Modulation of the Activity of Membrane Proteins. The Journal of Physical Chemistry B 2023, 127, 1598–1606. [DOI] [PubMed] [Google Scholar]
  • (25).Cantor RS Receptor Desensitization by Neurotransmitters in Membranes: Are Neurotransmitters the Endogenous Anesthetics? Biochemistry 2003, 42, 11891–11897. [DOI] [PubMed] [Google Scholar]
  • (26).Chisari M; Eisenman LN; Krishnan K; Bandyopadhyaya AK; Wang C; Taylor A; Benz A; Covey DF; Zorumski CF; Mennerick S The Influence of Neuroactive Steroid Lipophilicity on GABAAReceptor Modulation: Evidence for a Low-Affinity Interaction. Journal of Neurophysiology 2009, 102, 1254–1264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Jiang X; Shu H-J; Krishnan K; Qian M; Taylor AA; Covey DF; Zorumski CF; Mennerick S A clickable neurosteroid photolabel reveals selective Golgi compartmentalization with preferential impact on proximal inhibition. Neuropharmacology 2016, 108, 193–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (28).Solanko KA; Modzel M; Solanko LM; Wüstner D Fluorescent Sterols and Cholesteryl Esters as Probes for Intracellular Cholesterol Transport. Lipid Insights 2015, 8s1, LPI.S31617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (29).Rogers J; Lee AG; Wilton DC The organisation of cholesterol and ergosterol in lipid bilayers based on studies using non-perturbing fluorescent sterol probes. Biochim. Biophys. Acta 1979, 552, 23–37. [DOI] [PubMed] [Google Scholar]
  • (30).McIntosh AL; Atshaves BP; Huang H; Gallegos AM; Kier AB; Schroeder F Fluorescence Techniques Using Dehydroergosterol to Study Cholesterol Trafficking. Lipids 2008, 43, 1185–1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (31).Iaea DB; Gale SE; Bielska AA; Krishnan K; Fujiwara H; Jiang H; Max-field FR; Schlesinger PH; Covey DF; Schaffer JE et al. A novel intrinsically fluorescent probe for study of uptake and trafficking of 25-hydroxycholesterol. J. Lipid Res 2015, 56, 2408–2419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (32).Nåbo LJ; Modzel M; Krishnan K; Covey DF; Fujiwara H; Ory DS; Szomek M; Khandelia H; Wüstner D; Kongsted J Structural design of intrinsically fluorescent oxysterols. Chem. Phys. Lipids 2018, 212, 26–34. [DOI] [PubMed] [Google Scholar]
  • (33).Petersen D; Reinholdt P; Szomek M; Hansen SK; Poongavanam V; Dupont A; Heegaard CW; Krishnan K; Fujiwara H; Covey DF et al. Binding and intracellular transport of 25-hydroxycholesterol by Niemann-Pick C2 protein. Biochim. Biophys. Acta 2020, 1862, 183063. [DOI] [PubMed] [Google Scholar]
  • (34).Szomek M; Moesgaard L; Reinholdt P; Hald SBH; Petersen D; Krishnan K; Covey DF; Kongsted J; Wüstner D Membrane organization and intracellular transport of a fluorescent analogue of 27-hydroxycholesterol. Chem. Phys. Lipids 2020, 233, 105004. [DOI] [PubMed] [Google Scholar]
  • (35).Lee DK; Albershardt DJ; Cantor RS Exploring the Mechanism of General Anesthesia: Kinetic Analysis of GABAA Receptor Electrophysiology. Biophysical Journal 2015, 108, 1081–1093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (36).Langner M; Hui S Dithionite penetration through phospholipid bilayers as a measure of defects in lipid molecular packing. Chemistry and Physics of Lipids 1993, 65, 23–30. [DOI] [PubMed] [Google Scholar]
  • (37).Haralampiev I; Scheidt HA; Abel T; Luckner M; Herrmann A; Huster D; Müller P The interaction of sorafenib and regorafenib with membranes is modulated by their lipid composition. Biochimica et Biophysica Acta (BBA) - Biomembranes 2016, 1858, 2871–2881. [DOI] [PubMed] [Google Scholar]
  • (38).Akkerman V; Scheidt HA; Reinholdt P; Bashawat M; Szomek M; Lehmann M; Wessig P; Covey DF; Kongsted J; Müller P et al. Natamycin interferes with ergosterol-dependent lipid phases in model membranes. BBA Advances 2023, 4, 100102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (39).Scheidt HA; Müller P; Herrmann A; Huster D The Potential of Fluorescent and Spin-labeled Steroid Analogs to Mimic Natural Cholesterol. Journal of Biological Chemistry 2003, 278, 45563–45569. [DOI] [PubMed] [Google Scholar]
  • (40).Scheidt HA; Meyer T; Nikolaus J; Baek DJ; Haralampiev I; Thomas L; Bittman R; Müller P; Herrmann A; Huster D Cholesterol’s Aliphatic Side Chain Modulates Membrane Properties. Angewandte Chemie International Edition 2013, 52, 12848–12851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (41).Meyer T; Baek DJ; Bittman R; Haralampiev I; Müller P; Herrmann A; Huster D; Scheidt HA Membrane properties of cholesterol analogs with an unbranched aliphatic side chain. Chemistry and Physics of Lipids 2014, 184, 1–6. [DOI] [PubMed] [Google Scholar]
  • (42).Hilsch M; Haralampiev I; Müller P; Huster D; Scheidt HA Membrane properties of hydroxycholesterols related to the brain cholesterol metabolism. Beilstein Journal of Organic Chemistry 2017, 13, 720–727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (43).Schroeder F; Butko P; Nemecz G; Scallen TJ Interaction of fluorescent delta 5, 7, 9(11), 22-ergostatetraen-3 beta-ol with sterol carrier protein-2. Journal of Biological Chemistry 1990, 265, 151–157. [PubMed] [Google Scholar]
  • (44).Yang H; Tong J; Lee CW; Ha S; Eom SH; Im YJ Structural mechanism of ergosterol regulation by fungal sterol transcription factor Upc2. Nature Communications 2015, 6. [DOI] [PubMed] [Google Scholar]
  • (45).Chintala SM; Tateiwa H; Qian M; Xu Y; Amtashar F; Chen Z; Kirkpatrick CC; Bracamontes J; Germann AL; Akk G et al. Direct measurements of neurosteroid binding to specific sites on GABAA receptors. British Journal of Pharmacology 2024, [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Hosie AM; Wilkins ME; da Silva HM; Smart TG Endogenous neurosteroids regulate GABAA receptors through two discrete transmembrane sites. Nature 2006, 444, 486–489. [DOI] [PubMed] [Google Scholar]
  • (47).Wang L; Covey DF; Akk G; Evers AS Neurosteroid Modulation of GABAA Receptor Function by Independent Action at Multiple Specific Binding Sites. Current Neuropharmacology 2022, 20, 886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (48).Chen Q; Wells MM; Arjunan P; Tillman TS; Cohen AE; Xu Y; Tang P Structural basis of neurosteroid anesthetic action on GABAA receptors. Nature Communications 2018, 9, 3972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Magnaghi V GABA and Neuroactive Steroid Interactions in Glia: New Roles for Old Players? Current Neuropharmacology 2007, 5, 47–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (50).Smolič T; Tavčar P; Horvat A; Černe U; Halužan Vasle A; Tratnjek L; Kreft ME; Scholz N; Matis M; Petan T et al. Astrocytes in stress accumulate lipid droplets. Glia 2021, 69, 1540–1562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (51).Liu J; Chang CCY; Westover EJ; Covey DF; Chang T-Y Investigating the allosterism of acyl-CoA:cholesterol acyltransferase (ACAT) by using various sterols: in vitro and intact cell studies. Biochemical Journal 2005, 391, 389–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (52).Rogers MA; Liu J; Kushnir MM; Bryleva E; Rockwood AL; Meikle AW; Shapiro D; Vaisman BL; Remaley AT; Chang CC et al. Cellular Pregnenolone Esterification by Acyl-CoA:Cholesterol Acyltransferase. Journal of Biological Chemistry 2012, 287, 17483–17492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (53).Shu H-J; Eisenman LN; Jinadasa D; Covey DF; Zorumski CF; Mennerick S Slow Actions of Neuroactive Steroids at GABAAReceptors. The Journal of Neuroscience 2004, 24, 6667–6675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (54).Li P; Shu H; Wang C; Mennerick S; Zorumski CF; Covey DF; Steinbach JH; Akk G Neurosteroid migration to intracellular compartments reduces steroid concentration in the membrane and diminishes GABA-A receptor potentiation. The Journal of Physiology 2007, 584, 789–800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (55).Jorgensen WL; Chandrasekhar J; Madura JD; Impey RW; Klein ML Comparison of simple potential functions for simulating liquid water. The Journal of chemical physics 1983, 79, 926–935. [Google Scholar]
  • (56).Dickson CJ; Madej BD; Skjevik ÅA; Betz RM; Teigen K; Gould IR; Walker RC Lipid14: the amber lipid force field. Journal of chemical theory and computation 2014, 10, 865–879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (57).Sami S; Menger MF; Faraji S; Broer R; Havenith RW Q-Force: Quantum mechanically augmented molecular force fields. Journal of Chemical Theory and Computation 2021, 17, 4946–4960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Grimme S; Hansen A; Ehlert S; Mewes J-M r2SCAN-3c: A “Swiss army knife” composite electronic-structure method. The Journal of Chemical Physics 2021, 154. [DOI] [PubMed] [Google Scholar]
  • (59).Neese F Software update: The ORCA program system—Version 5.0. Wiley Interdisciplinary Reviews: Computational Molecular Science 2022, 12, e1606. [Google Scholar]
  • (60).Wang J; Wolf RM; Caldwell JW; Kollman PA; Case DA Development and testing of a general amber force field. Journal of computational chemistry 2004, 25, 1157–1174. [DOI] [PubMed] [Google Scholar]
  • (61).Abraham MJ; Murtola T; Schulz R; Páll S; Smith JC; Hess B; Lindahl E GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1, 19–25. [Google Scholar]
  • (62).Darden T; York D; Pedersen L Particle mesh Ewald: An N log (N) method for Ewald sums in large systems. The Journal of chemical physics 1993, 98, 10089–10092. [Google Scholar]
  • (63).Hess B; Bekker H; Berendsen HJ; Fraaije JG LINCS: a linear constraint solver for molecular simulations. Journal of computational chemistry 1997, 18, 1463–1472. [Google Scholar]
  • (64).Nosé S A unified formulation of the constant temperature molecular dynamics methods. The Journal of chemical physics 1984, 81, 511–519. [Google Scholar]
  • (65).Hoover WG Canonical dynamics: Equilibrium phase-space distributions. Physical review A 1985, 31, 1695. [DOI] [PubMed] [Google Scholar]
  • (66).Berendsen HJ; Postma J. v.; Van Gunsteren WF; DiNola A; Haak JR Molecular dynamics with coupling to an external bath. The Journal of chemical physics 1984, 81, 3684–3690. [Google Scholar]
  • (67).Parrinello M; Rahman A Polymorphic transitions in single crystals: A new molecular dynamics method. Journal of Applied physics 1981, 52, 7182–7190. [Google Scholar]
  • (68).Michaud-Agrawal N; Denning EJ; Woolf TB; Beckstein O MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. Journal of computational chemistry 2011, 32, 2319–2327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (69).Nåbo LJ; Modzel M; Krishnan K; Covey DF; Fujiwara H; Ory DS; Szomek M; Khandelia H; Wüstner D; Kongsted J Structural design of intrinsically fluorescent oxysterols. Chemistry and physics of lipids 2018, 212, 26–34. [DOI] [PubMed] [Google Scholar]
  • (70).Pierce SR; Germann AL; Steinbach JH; Akk G The sulfated steroids pregnenolone sulfate and dehydroepiandrosterone sulfate inhibit the α1β3γ2L GABAA receptor by stabilizing a novel nonconducting state. Molecular pharmacology 2022, 101, 68–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (71).Davis J; Jeffrey K; Bloom M; Valic M; Higgs T Quadrupolar echo deuteron magnetic resonance spectroscopy in ordered hydrocarbon chains. Chemical Physics Letters 1976, 42, 390–394. [Google Scholar]
  • (72).Sternin E; Bloom M; MacKay A De-Paking of NMR-Spectra. J. Magn. Reson 1983, 55, 274–282. [Google Scholar]
  • (73).Lafleur M; Fine B; Sternin E; Cullis P; Bloom M Smoothed orientational order profile of lipid bilayers by 2H-nuclear magnetic resonance. Biophysical Journal 1989, 56, 1037–1041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (74).Wüstner D; Landt Larsen A; Faergeman NJ; Brewer JR; Sage D Selective Visualization of Fluorescent Sterols in Caenorhabditis elegans by Bleach-Rate-Based Image Segmentation. Traffic 2010, 11, 440–454. [DOI] [PubMed] [Google Scholar]
  • (75).Stringer C; Wang T; Michaelos M; Pachitariu M Cellpose: a generalist algorithm for cellular segmentation. Nature methods 2021, 18, 100–106. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary material
1

RESOURCES