Abstract
Aquaporins (AQPs), transmembrane water-conducting channels, have earned a great deal of scrutiny for their critical physiological roles in healthy and disease cell states, especially in the biomedical field. Numerous methods have been implemented to elucidate the involvement of AQP-mediated water transport and downstream signaling activation in eliciting whole cell, tissue, and organ functional responses. To modulate these responses, other methods have been employed to investigate AQP druggability. This review discusses standard in vitro, in vivo, and in silico methods for studying AQPs, especially for biomedical and mammalian cell biology applications. We also propose some new techniques and approaches for future AQP research to address current gaps in methodology.
Keywords: aquaporins, methods, in vitro, permeability, inhibitors, cell behaviors
Introduction
Aquaporins (AQPs), transmembrane protein channels responsible for facilitating passive water flux across cell membranes, are critical for regulating cell shape, cell volume, and fluid homeostasis in nearly all organisms. Thirteen AQP isoforms (AQP0–AQP12) have been identified in mammals [1], some of which are classified as aquaglyceroporins that transport glycerol in addition to water (AQP3, 7, 9, and 10) or super AQPs (AQP11 and AQP12) that have a unique cysteine residue in their pore-forming asparagine-proline-alanine (NPA) box [2]. AQPs are known to drive water-essential processes such as tear production, saliva secretion, renal reabsorption, and vaginal lubrication, but they also have other nonobvious functions in cell migration, cell invasion, proliferation, angiogenesis, and neuroexcitation. Growing evidence indicates AQP involvement in the pathophysiology of multiple clinical conditions including cancer, endometriosis, glaucoma, and epilepsy [3, 4], thus prompting further investigation into AQP-mediated processes and the manipulation of AQPs in the design of therapeutics.
Over the past 30 years, the diverse cellular roles of AQPs and their mechanisms in physiological and pathophysiological systems have been studied in vitro, in vivo, and in silico. Although the research performed with these methods has provided a foundational understanding of AQPs, further study is still warranted to clarify AQP-mediated biological processes. In vitro methods may be used to induce or measure AQP expression and subcellular localization or to analyze AQP function in regulating cell behaviors. Several in vivo models drive many of the pre-clinical efforts in understanding the relationship between AQPs and disease progression. Finally, there are several in silico methods that can elucidate AQP structure, function, and interactions with other molecules and proteins, as well as model the resulting cell behavior. The goal of this review is to collate and discuss established in vitro, in vivo, and in silico methods for studying mammalian AQPs in biomedical science and engineering applications. A condensed collection of these methods is depicted in Fig. 1. We also suggest some new techniques and models that can be applied for future AQP-related research.
Figure 1.
Graphical summary of in vitro, in vivo, and in silico methods for AQP biology as discussed in this review.
In vitro methods for studying AQPs
Once AQPs were discovered in 1992 by Peter Agre et al. [5, 6], X-ray crystallography [7], nuclear magnetic resonance (NMR) spectroscopy [8], and 3D electron microscopy [9] techniques defined the structures of AQP isoforms. Since then, the majority of work conducted on AQP biology and druggability has been conducted in vitro; this prevalence occurs because in vitro models allow for the fine-tuned manipulation of experimental conditions, greater accessibility in observing mechanistic changes, and overall ease of use. Moreover, in vitro techniques are sometimes preferred over in vivo experiments for their practicality, cost–effectiveness, and ethicality. A summary of in vitro AQP methods and example references is detailed in Table 1.
Table 1.
In vitro methods
| Study area | Goal | Method | Example Ref(s) |
|---|---|---|---|
| Expression & Localization | AQP Expression | siRNA or shRNA lentiviral vector transfections | [10–13] |
| Overexpression vector transfection | [14, 15] | ||
| AQP Localization | IF | [16–18] | |
| Fluorescently labeled AQPs | [15, 19, 20] | ||
| Osmotic stimuli | [21] | ||
| AVP | [22, 23] | ||
| AQP Function | Permeability Assays | Proteoliposomes | [24] |
| Stopped-flow spectrophotometry | [25–29] | ||
| X. laevis oocyte swelling | [30, 31] | ||
| Fluorescence (calcein) | [32, 33] | ||
| Epithelial assays | [34] | ||
| Yeast freeze–thaw challenge | [35, 36] | ||
| Ion conductance using electrophysiology | [37–42] | ||
| Cell swelling with glycerol gradient | [25, 27–29] | ||
| H2O2 fluorescent label | [26] | ||
| Modulating AQP activity | Small molecule inhibitors | [43] | |
| Heavy metal inhibitors | [44] | ||
| Bumetanide derivative | [45, 46] | ||
| Plant-based compound inhibitors | [46, 47] | ||
| Drug AQP translocation | [48] | ||
| Modlate ion conductance | [26, 38, 41] | ||
| Optogenetics | [49] | ||
| AQP–Protein Interactions | co-IP | [50–54] | |
| PDAs | [45, 55] | ||
| Y2H screening | [56] | ||
| Cell swelling assays | [53, 56, 58] | ||
| Cross-linking mass spectrometry | [59–62] | ||
| Overlay assays | [63, 64] | ||
| Microscale thermophoresis | [27, 65, 66] | ||
| Proximity ligation assays | [66, 67] | ||
| AQP–Lipid Interactions | X-ray crystallography/electron crystallography | [68–71] | |
| Phosphorylation Site Identification & Characterization | Liquid chromatography-mass spectrometry | [72–74] | |
| Radioactively labeled ATP & immunoprecipitation | [53, 56] | ||
| Phospho-specific antibodies | [72] | ||
| Functional mutant AQPs | [23, 67, 75] | ||
| Cell Behavior | Migration & Invasion | Wound healing | [12, 66, 76] |
| Transwell invasion | [12, 66, 76] | ||
| 2D random migrationa | [77] | ||
| Monolayer incorporationa | [78] | ||
| Spheroid invasion | [47] | ||
| 3D random invasiona | [79] | ||
| Confining microfluidic devices | [16, 80–82] | ||
| Actin cytoskeleton visualization using ICC | [11, 18, 76, 83, 84] | ||
| Fluorescently labeled filamentous actin probe | [15, 85] | ||
| Western blotting for invasive markers | [76, 83, 85, 86] | ||
| Protrusion analysis | [15, 76, 85, 87–89, 90] | ||
| Proliferation | AlamarBlue | [13, 83] | |
| Bromodeoxyuridine | [91] | ||
| MTT | [92] | ||
| Cell Counting Kit 8 | [93] | ||
| Crystal violet assays | [47] | ||
| Apoptosis | Caspase-3 activity assay | [94–96] | |
| Flow cytometry with a viability dye & annexin V | [47, 97] | ||
| Terminal deoxynucleotidyl transferase dUTP nick end labeling | [96, 98, 99] | ||
| Western blotting for apoptotic factors | [95, 98, 99, p. 4, 100] | ||
| Swelling assay | [94] | ||
| Mitochondrial membrane potential analysis | [94, 100] | ||
| Cell Adhesion | Crystal violet staining | [83] | |
| AFM | [10] | ||
| ICC visualization of junction proteins | [18, 101] | ||
| Dispase-based dissociation | [101] | ||
| Phenotyping cell–cell junctionsa | [102] | ||
| Fluorescence-based microtubule polymerization | [103] | ||
| Angiogenesis | Tube formation | [69, 81, 88] | |
| HUVEC cell assays | [13, 104–107] | ||
| Western blotting for angiogenic factors | [98] | ||
| IF | [108] |
Assay that, to our knowledge, has not been employed yet to study AQPs.
Expression and localization
Since AQP localization and expression are essential to the roles they play in the regulation of cellular processes, it is worth identifying their alterations across healthy and disease states. Here, we summarize traditional methods for measuring and modulating AQP expression and subcellular distribution.
AQP expression
AQP genes are commonly knocked down using RNA interference. siRNA transfections and shRNA lentiviral vector transductions are common techniques for gene knockdowns [109, 110] as they are low-cost and quick to perform compared to a knockout. Full AQP knockouts have been created primarily in mice and will be discussed in ‘Observing AQP expression and activity in vivo’ section. To induce or increase AQP expression, cells can be transfected with a constructed overexpression vector [76]. Western blots and RT-PCR/RT-Qpcr [14, 91–111] are used to quantify AQP protein and gene expression levels, respectively. These studies continuously add to a wide range of healthy and pathophysiological microenvironmental components that have been shown to alter AQP expression [83, 112–115].
AQP localization
The ability to visualize the localization of AQPs, which primarily exist on the plasma membrane or within the cytoplasm, can lend additional insight into AQP abundance, trafficking, and function with respect to cell behavior. Immunofluorescence (IF), known as immunocytochemistry (ICC) for cells in vitro, is the gold standard method to visualize protein localization. ICC can be employed for cells situated in different environments, including 2D culture and 3D culture such as within microfluidic devices [16]. Although IF produces qualitative data, multiple parameters can easily be quantified from microscopy images. For example, the mean fluorescence intensity (MFI) of a region of interest (ROI) is often used as a proxy for relative protein expression in a sample compared to a control. Measuring MFI values and selecting an ROI are easily accomplished using the Fiji-ImageJ software [116]. Additionally, the nuclear to cytoplasmic MFI ratio is one metric of subcellular AQP distribution that can be informative for cells expressing intranuclear AQPs, a phenomenon that has been previously reported [17]. The main disadvantage of ICC is that it does not allow for dynamic visualization of AQP stimulus/response trafficking since the samples must be fixed. Instead, transfecting cells with a construct to tag AQPs with a fluorescent protein are the standard technique for imaging live subcellular AQP distribution over time [15, 19, 20]. Furthermore, observing the co-localization of AQPs with other markers associated with cellular trafficking can provide insight into the mechanisms underlying subcellular distribution [22, 117].
In addition to observing AQP localization, it is possible to induce a particular subcellular distribution if needed. AQPs are predominantly found in the cytoplasm or on the cell surface, which is often a result of trafficking activities that are regulated by molecules such as arginine vasopressin (AVP), adenosine 3′,5′-cyclic monophosphate (cAMP), protein kinase C (PKC), and protein kinase A (PKA) [118]. These molecules use phosphorylation to regulate the endocytosis, storage, degradation, or exocytosis of AQPs in response to a stimulus. For example, redistribution of AQP1 to the cell membrane has been observed in human umbilical vein endothelial cells (HUVECs) upon exposure to osmotic stimuli, which was immediately mediated by PKC and then mediated by calcium signaling in the following minutes [21]. Similarly, AVP can induce a translocation of AQP2 to the membrane during dehydration [23] or AQP2 internalization in other instances [22, 117]. Subcellular AQP distribution patterns may differ across different tissues, and due to dynamic trafficking activities, AQPs are generally not solely situated in one localization [25, 48, 119–121]. Even AQP6, which is predominantly found intracellularly [122], has recently been reported in discrete areas of the plasma membrane [123].
Analyzing the AQP function
AQPs are traditionally known for their function in transmembrane water transport to regulate cell volume and support a wide variety of physiological phenomena, but some have been shown to play a second role in the activation of multiple signaling pathways such as Ras/Raf [14], MAPK/p38 [124], Wnt [50], and PI3K/Akt [125, 126]. To fully understand the impact of AQPs in biological systems, it is important to have methods to characterize both water permeation and signaling mechanisms.
Permeability assays
Permeability assays leverage an osmotic gradient to measure AQP-mediated water flux. For benefits, limitations, and further details of each AQP permeability assay, we direct the reader to other reviews [127, 128].
Stopped-flow
Water permeation through AQP pores can be measured in AQP-containing proteoliposomes using stopped-flow spectrophotometry. In this technique, the scattered light intensity correlates with the change in vesicle volume after an osmotic shock. Although this method requires a challenging sample preparation process, Yue et al. devised an Escherichia coli Cell-Free Protein Synthesis (CFPS) system that does not require cells to generate AQPs, circumventing the need for an additional purification step [24]. The CFPS system efficiently synthesizes a large quantity of AQPs that are then inserted into the liposomes in order to study water permeability. In addition to proteoliposomes, AQP-expressing yeast cells have also been used with stopped-flow measurements [26].
Cell-based permeability assays
More commonly, cell swelling assays are utilized to measure permeability. Usually, human AQPs are heterologously expressed in Xenopus laevis oocytes, and the cell volume is monitored with videomicroscopy—such as phase contrast, confocal or fluorescence microscopy—during a hypo-osmotic shock [30]. The oocytes do have other endogenous membrane proteins [24], but the advantage of this model organism is its low intrinsic water and glycerol permeability [129]. Using fluorescence microscopy techniques, cell volume can be measured with fluorescent dyes that undergo quenching during cell swelling or shrinking from an osmotic shock [130]. To achieve high-throughput measurements, Fenton et al. adapted a calcein-quenching-based assay for a plate reader that can be employed for a variety of adherent cell types: briefly, the calcein-AM fluorophore was loaded into cells and cell volume changes were calculated from plate reader fluorescence intensity readings [32, 131]. A step-by-step protocol for this assay has been written by Kitchen et al. [132]. Similarly, Mola et al. developed an automated cell-based AQP inhibitor screening assay, which utilizes a microplate reader to make dynamic fluorescence measurements of cell volume; the assay was validated using calcein-AM-loaded primary astrocytes and fibroblasts that strongly express AQP4 and AQP1, respectively [33].
Epithelial permeability assays, where an epithelial monolayer on a permeable surface is mounted in an Ussing chamber, are also used to measure water transport [34]. An osmotic gradient is applied to one compartment and the fluid height is measured in a capillary tube connected to the other compartment; alternatively, a fluorescent dye is added to the hyperosmotic compartment and the rate at which the fluorescence intensity decreases is measured. This method is best for characterizing AQP-mediated cellular water permeability in epithelial cell monolayers [129].
Yeast has also been used as a permeability measurement tool: AQPs of any origin are heterologously expressed in yeast cells and a freeze–thaw challenge is applied [35, 36]. The survival of the yeast post-challenge depends on AQP-mediated water flux, making this an effective and simple inhibitor screening method.
Measuring permeability of unconventional permeants
In addition to facilitating water flux, certain AQP isoforms are capable of transporting ions, with the most notable example being AQP1. AQPs are known to form a quaternary tetrameric structure, resulting in a central pore between the monomers. The role of this central pore has been a topic of debate in the AQP biology field, but it is the current understanding that the central pore of AQP1 acts as a cGMP-gated ion channel [133–135]. Ionic conductance is typically measured with electrophysiology techniques, including two-electrode voltage clamp [37–39], black lipid membrane [40, 136], and patch-clamp experiments [39, 41]. Furthermore, some AQPs can transport small solutes, such as glycerol and hydrogen peroxide (H2O2) [42]. To measure glycerol permeability, a cell swelling assay on the X. laevis oocytes can be performed by replacing the sodium chloride in the culture medium to glycerol at the same tonicity. The cell volume increase represents a solute influx first, followed by water influx [10, 130, 137]. The stopped-flow light scattering method is another technique used to measure glycerol permeability in cells under an applied glycerol gradient [25, 27–29]. H2O2 flux has been measured under live cell imaging after transfecting cells with a construct containing a fluorescent tag such as the HyPer biosensor [11, 29, 45] or the H2O2-specific small molecule indicator Peroxy Yellow 1 Methyl-Ester [11]. Similarly, live cell imaging can be performed on cells incubated with the cell-permeant indicator 2′,7′-dichlorodihydrofluorescein diacetate (H2-DCFDA) that fluoresces upon oxidation [26]. In the same study, yeast cells expressing human AQPs, H2O2 uptake was measured with an electrochemical assay which measures its conversion to oxygen [26].
Modulating AQP activity
Modulating the activity of AQPs by blocking their transport functions or their trafficking is desirable for in vitro experiments as well as for future clinical applications in treating AQP-mediated diseases. The following subsections describe current strategies for reducing AQP activity.
Drugging AQP-mediated water permeation
The majority of putative AQP inhibitors are small molecules, including tetraethylammonium (TEA), acetazolamide, anti-epileptic drugs, TGN-020, and phloretin, which generally work by blocking water transport through AQP pores. Heavy metal compounds such as nickel chloride, copper compounds, and silver are used as cytotoxic, nonspecific AQP water inhibitors. Bumetanide derivative AqB013 [45] and plant-based compound bacopaside II [30] have shown AQP1 water channel inhibiting abilities. Additionally, AuPhen is capable of significantly reducing glycerol permeability in AQP3. A detailed description of these inhibitors and their putative AQP binding sites can be found in other reviews [127, 128]. Unfortunately, these inhibitors range in efficacy for a limited number of human, mice, or rat AQP isoforms and produce variable outcomes across different permeability assays [127]. The lack of specific, effective, and nontoxic AQP inhibitors means that further high-throughput screening is necessary [138]. Once more AQP inhibitor candidates are identified, multiple permeability assays should be performed reproducibly to validate that they remain robust under a variety of test conditions. A discussion on computational screening methods is included in ‘AQP drug screening’ section.
Targeting AQP translocation
Aside from blocking water flux with AQP inhibitors, cell water permeability could be controlled by drugging signaling pathways that regulate AQP trafficking to the cell surface, such as PKC-dependent pathways for AQP1, and cAMP- and PKA-dependent pathways for AQP5 [118]. Recently, trifluoperazine blocked AQP4 localization to the blood-spinal cord barrier in astrocytes by inhibiting calmodulin. Since calmodulin likely binds directly with AQP4 to induce its translocation, trifluoperazine successfully neutralized the hypoxia-driven increase in astrocyte water flux [48].
Drugging AQP-mediated ion conductance
Lastly, a group of AQP-targeting drugs that modulate ion conductance through the putative central ion pore of AQP1 has been identified. The central ion channel is of interest due to its potential involvement in volume regulation and signal transduction [139], and indeed, AQP1 ion conductance has been shown to facilitate human colon cancer cell migration [37]. To attenuate this migratory behavior, ion conductance through the central pore can be blocked with the bumetanide derivative, AqB011, alone or in conjunction with one of the aforementioned AQP (water flux) inhibitors such as bacopaside II [37, 46]. In another study, AqB013 and AqB050, which are also AQP1 drugs, inhibited HUVEC tube formation via induction of apoptosis and impairment of cell migration [45]. Finally, work by Palethorpe et al. showed that bacopasides I and II can synergistically inhibit breast cancer cell growth, migration, and invasion by blocking AQP1-mediated ion and water transport [47].
Optogenetics to modulate AQPs
Optogenetic techniques have recently emerged as tools for the precise manipulation of protein function and localization in specific, photosensitive cells using light [140]. While optogenetics has not been widely employed for AQP modulation, they have been used to control light-gated ion channels that create an osmotic gradient and trigger a water influx or efflux through AQP1 [49]. Future investigations should continue to explore optogenetics as a means to control AQP expression, localization, and function in cells.
AQP-protein and AQP–lipid interactions
Although individual AQP monomers are functional water channels, they often interact with each other to form tetramers: each monomer transports water and the role of the central pore in between the monomers is still controversial. AQPs of the same or different isoforms may combine to form these tetramers [65]. AQPs may interact with other proteins, such as ion channels that establish osmotic gradients to drive water transport or upstream or downstream molecules in signaling pathways. AQPs can trigger several pathways, including Ras/Raf/ERK [14, 66], MAPK/p38 [124], Wnt/β-catenin [50], and PI3K/Akt [125, 126]. On the other hand, upstream signals regulate cell water permeability by triggering AQP membrane translocation or internalization [67, 141]. Identification of protein–protein interactions (PPIs) is valuable for understanding AQP activities within the cell, and the regulatory mechanisms behind these activities can be further elucidated by studying AQP post-translational modification sites. AQP–lipid interactions are also critical in biological membranes, as these interactions impact both integral protein function and lipid organization. AQP–lipid interactions are much less characterized than AQP–protein interactions due to our incomplete knowledge of atomic-level AQP isoform structures but are worth studying due to their impact on membrane properties.
AQP–protein interaction detection methods
Of the in vitro techniques for extracting AQP binding partners, co-immunoprecipitation (co-IP) [50, 51, 52, 103], pull-down assays (PDAs) [63, 142, 143], and yeast two-hybrid (Y2H) screening [56] are the gold standards. Co-IP and PDA involve isolating the protein of interest along with their putative binding partners from a cell lysate. They are usually used in conjunction with IF to ensure that the proteins are co-localized in the native cell, and with mass spectrometry to identify the protein complexes. Y2H employs a reporter gene and a transcription activator with an activating domain and a DNA binding domain. In addition to co-IP and IF [53], cell swelling assays have been used to characterize AQP collaboration with ion channels, especially transient receptor potential vanilloid 4 [53, 57, 58]. Other less commonly used methods that have been employed to reveal AQP–protein interactions include cross-linking mass spectrometry [59–62], overlay assays [63, 64], microscale thermophoresis [48, 141, 144, 145], and proximity ligation assays [143, 145, 146]. The AQP interactome, including binding partners and detection method(s) used, has been thoroughly discussed in other reviews [65, 147, 148]. Computational PPI methods can be found in ‘Identification of AQP–protein and AQP–lipid interactions’ section.
Identification of post-translational modification sites on AQPs
Once putative binding partners are identified, their interactions with AQPs can be further characterized by determining likely phosphorylation sites via a phosphoproteomic analysis. Knowledge of AQP phosphorylation sites lends a greater mechanistic understanding of how AQPs receive regulatory signals from other proteins, eventually resulting in altered water permeability or subcellular localization [118]. These regulatory events could be targeted in order to control AQP activities or downstream signaling pathways. AQP phosphorylation sites are already well-defined experimentally and are tabulated on bioinformatics databases [72, 149, 150]. Phosphoproteomic analyses are usually carried out by various phosphopeptide enrichment methods. For example, a prominent study combined immobilized metal affinity chromatography and liquid chromatography–mass spectrometry (LC–MS) neutral loss scanning to enrich phosphopeptides from rat inner medullary collecting duct (IMCD) tissues treated with AVP, leading to the detection of four phosphorylation sites on AQP2 [72]. After advancements in the sensitivity and accuracy of mass spectrometers, these results were confirmed over a decade later using LC–MS in a similar experimental setup [73]. Another group used an in vitro phosphorylation assay in which rat tissue homogenate was incubated with radioactively labeled ATP and either PKA or PKC. The phosphorylated proteins were then immunoprecipitated with an AQP1 or AQP4 antibody, respectively, showing that AQP1 is phosphorylated by cAMP-dependent PKA [151] and that AQP4 is phosphorylated by PKC in the presence of a PKC activator [151, 152]. The same assay was also utilized to study AQP7 phosphorylation by PKA [143]. These interactions can be further quantified by measuring protein levels using phospho-specific antibodies; Hoffert et al. used an AQP2 antibody that recognized site S256 phosphorylation and found that samples treated with AVP showed an increased phospho-AQP2 abundance compared to the control [72].
In addition to phosphorylation sites, some AQP isoforms contain ubiquitylation [67], glycosylation [75], SUMOylation [153], and acetylation [74] sites. These sites have been identified similarly to phosphorylation sites; for example, an AQP3 SUMOylation site was found by wild-type α-lytic protease digestion of murine tissues and subsequent mass spectrometry [153], and an AQP3 acetylation site was identified using liquid chromatography–tandem mass spectrometry on rat IMCD samples [74]. AQP2 ubiquitylation and N-glycosylation were studied using mutant forms in Madin–Darby canine kidney cells [67, 75]. This method of using site-directed mutagenesis on AQPs is widely used to understand how the loss-of-function of a PPI (phosphorylation, ubiquitylation, glycosylation, acetylation, etc.) of interest affects AQP regulation and cell behavior [23].
AQP–lipid interaction characterization methods
The AQP–lipid interactions that have been studied thus far come from X-ray crystallography and electron crystallography, which allows for imaging of atomic-level structures of AQP–lipid complexes [7]. In particular, the structure of AQP0 has been solved by electron crystallography, so its integration and organization with various lipids have been observed in detail [68–71]. Structures obtained from crystallography are often then used in computational simulations as described in ‘Identification of AQP–protein and AQP–lipid interactions’ section .
Understanding cell behavior with respect to AQP activity
AQPs have gained significant interest for their roles in cell activities such as fluid homeostasis, proliferation, migration, invasion, angiogenesis, and apoptosis. These functions are fundamental to healthy processes including kidney water reabsorption, brain-water homeostasis, neuroexcitation, development, and wound healing; and are also crucial for disease processes like cancer metastasis, tumor growth, seizures, and endometriosis development [3, 4]. Having a wide variety of in vitro cell-based assays enables us to study the involvement of AQPs in both healthy and pathologic cell states. Although this is not a comprehensive review of cell-based assays, here we discuss common techniques used in AQP studies and highlight those that are particularly relevant to AQP regulation of cellular processes. These experiments are typically performed with AQP knockdowns or overexpressions, AQPs treated with inhibitors, or functionally mutated AQPs, along with controls.
Cell migration and invasion
AQPs have been scrutinized for their involvement in the migration and invasion of healthy and diseased cells. Specifically, AQPs are responsible for facilitating rapid water fluxes that allow cells to change volume and shape and provide space for actin polymerization. AQPs co-localize with ion channels that establish an osmotic gradient at the leading edge of a cell, thereby inducing a water influx that creates space for actin to polymerize. Similarly, an AQP-mediated outflux of water at the trailing edge of the cell can help the membrane retract [154]. When studying the impacts of AQPs on cell motility, the most simplistic experiment that countless studies have included is a wound healing or gap closure assay, where collective cell migration is measured by the change over time in the width of a scratch or gap made in a monolayer. In parallel, many researchers use transwell invasion assays in which individual cells invade from an upper chamber through a semipermeable membrane into a lower chamber containing a chemoattractant. To give a few examples, knocking down AQP1 [66], AQP3 [76], and AQP5 [12] expression resulted in reduced wound closure and number of invaded cells for gastric cancer cells, endometrial carcinoma cells, and nonsmall cell lung carcinoma cells, respectively.
Although wound healing and transwell invasions are useful low-cost methods, they only provide insight into the involvement of AQPs in directed cell migration; it is unknown whether AQPs have a significant role in undirected single-cell migration. Given that AQPs facilitate rapid cell volume changes and polarize to the leading and trailing edges of moving cells, studying 2D random migration may shed light on the contribution of AQPs to innate cell motility. This can be achieved with time-lapse tracking of individual cells with altered AQP expression or function; this would yield data on cell speed, directional persistence, mean square displacement, and morphological parameters. Another simple AQP-mediated cell invasion method could be an incorporation assay [78], which has not yet been employed for studying AQPs. Briefly, invasive cells such as MDA-MB-231s are stained and seeded on top of an endothelial or epithelial monolayer; the percentage of cells that are “incorporated” and disappear into the monolayer is quantified. This method could be used to characterize the effect of the AQPs in the invading cells or in the cell monolayer. In other words, do AQPs enhance the capability of the invading cells to disrupt the monolayer, and/or do the AQPs in the monolayer affect its barrier function? This assay can provide a preliminary answer to this question. Moreover, the method is relevant to AQP biology because the invading cells have to deform and squeeze through cell–cell junctions in the monolayer, representing a form of confined migration which AQPs may facilitate.
In addition to these basic migration assays, it is important to use models that are more representative of the in situ physiological microenvironment that cells experience (and which AQPs may help them traverse) during healthy and disease processes. A spheroid invasion assay more closely mimics how cells can metastasize away from a primary tumor, for example. One group found that treating breast cancer cells with bacopasides I and II, putative AQP1 inhibitors, significantly reduced MDA-MB-231 triple-negative breast cancer spheroid invasion into the surrounding matrix [47]. Heterotypic spheroids with leader and follower cells would be particularly fascinating for studying AQP-mediated invasion; follower cells can become invasive and travel through microtracks made by malignant leader cells [155]. Spheroid invasion assays would shed light on whether AQP expression enhances the ability of leader cells to proteolytically degrade the extracellular matrix or the ability of follower cells to deform and migrate through pre-made tracks. And, similar to the 2D random migration assay, there have been no reports of the role of AQPs in undirected cell invasion: a random invasion assay on cells evenly distributed within a 3D matrix such as collagen or Matrigel without a chemoattractant could be informative.
Stroka et al. also utilized a more physiologically relevant 3D model for cell migration in confined spaces that exist in vivo, such as tissue microtracks. Within a polydimethylsiloxane microchannel device, they blocked actin polymerization and myosin II-mediated contractility and administered osmotic shocks at the front or back of the cell. They found that water permeation drives tumor cell migration in confined spaces in the absence of actin polymerization and myosin II-mediated contractility, a phenomenon termed the “Osmotic Engine Model” [16]. In light of these findings, it makes sense to use methods that further clarify how AQPs might enable confined migration of invasive cell types, especially since AQPs facilitate rapid cell shape changes. Although several different models of confinement exist [156], they have not been widely employed to study AQPs. However, three papers after the establishment of the Osmotic Engine Model have reported research on confined migration with respect to AQPs. Two studies similarly performed time-lapse imaging of confined cell migration in a microfluidic device: one of them studied live AQP4 localization [80], and the other measured migration speed during an AQP5 knockdown [81]. The third investigation used an electro-osmotic microfluidic system that applies precise osmotic gradients to cells in microchannels, leading them to find that knocking down AQP4 resulted in a decreased confined cancer cell migration speed [82]. Broader use of different types of confined migration models across more AQP isoforms will help elucidate the influence of AQPs in enabling cells to move and deform through tight spaces.
Apart from directly studying cell migration and invasion, AQP expression has been linked to other hallmarks of invasion such as actin cytoskeleton rearrangement, focal adhesion dynamics, and epithelial–mesenchymal transition (EMT) [154]. The standard for studying the actin cytoskeleton and focal adhesion kinase organization is ICC [18, 50, 76, 84, 85, 157], but for unfixed cells, actin dynamics are usually observed by transfecting cells with a fluorescently labeled filamentous actin probe [15, 85], such as LifeAct. In many AQP studies, expression levels of cytoskeletal proteins and EMT markers have been measured with simple western blots [76, 83–86, 157], although this type of data can only provide correlations with AQP expression. A more AQP-relevant hallmark of cell invasiveness would be to visualize water-induced membrane protrusions, which multiple cell motility studies have done using microscopy techniques [15, 76, 84, 85, 87–90]. One study quantified bleb-like protrusions and filopodia from fluorescence microscopy images and found that overexpression of AQP9 in HEK-293 cells promoted the formation of membrane protrusions and therefore directed actin polymerization [15]. Similarly, knocking down AQP1 in rat gastric epithelial cells impaired lamellipodia formation during wound repair [87]. Continued use of membrane protrusion analyses with respect to other AQP isoforms and migratory cell types would help diversify our understanding of AQPs’ function in the formation of migratory membrane structures. Future work could also determine how inhibiting AQP water flux impacts cell morphology.
Cell proliferation, apoptosis, adhesion, angiogenesis, and more
AQPs have been implicated in other cell behaviors, including proliferation, adhesion, angiogenesis, and apoptosis. AQPs facilitate volume changes necessary for cell division, and some isoforms participate in proliferative signaling pathways such as Ras/Raf/ERK [14] and MAPK/p38 [124]. Moreover, aquaglyceroporins transport glycerol, an important intermediate for ATP production and biosynthesis [158]. The alamarBlue [13, 83], BrdU [91], MTT [92], CCK-8 [93], and crystal violet assays [47] have been widely used to understand the effect of AQP expression on cell proliferation. Most often, AQP knockdowns have been correlated with reduced cell proliferation [13, 83, 91, 93]. But of the studies on AQPs and proliferation, there have been markedly fewer that utilize the two AQP modulation strategies other than expression: inhibitors and mutants. Performing more thorough testing of cell proliferation with respect to AQP water/glycerol permeability and AQP phosphorylation would lend a more holistic understanding of which mechanism(s) different AQP isoforms use to impact proliferation. Along with proliferation assays, dynamic AQP localization should be analyzed during cell division via live imaging of fluorescently labeled AQPs to gain more clues on their role in proliferation.
AQPs are linked to apoptosis as well by allowing a water efflux during the apoptotic volume decrease (AVD); after AVD, which is driven by an efflux of K+, AQPs are thought to be inactivated to allow intracellular K+ concentrations to drop enough for the activation of apoptotic enzymes [94]. However, the relationship between AQPs and cell death is unclear due to mixed results [10, 94, 95, 97–100, 159, 160, 163] and should be further investigated. There are several ways apoptosis can be characterized in AQP experiments, including a caspase-3 activity assay [94–96], flow cytometry with a viability dye and annexin V [97, 163], a terminal deoxynucleotidyl transferase-mediated dUTP nick end labeling assay [96, 98, 99], western blotting for apoptotic factors [95, 98–100], a swelling assay for measuring a reduced cell water permeability post-AVD [94], and a mitochondrial membrane potential analysis [94, 100]. Like proliferation, most of these methods are performed in the context of an AQP knockdown; however, the cell swelling assay is the most direct way to measure the effect of the AQP water transport function on apoptotic shrinkage. Jablonski et al. used flow cytometry to perform a cell swelling size distribution analysis after a hypotonic insult on shrunken (apoptotic) cells and healthy cells. The shrunken cells were significantly less water-permeable when compared to healthy cells, supporting the idea that AQPs may become inactivated after AVD to enable apoptotic enzyme activation. Since the AQP inactivation does not seem to be a result of AQP degradation or removal from the cell membrane [94], future use of mutant AQPs (with modified phosphorylation sites) in the apoptosis methods listed above would clarify the mechanism of this inactivation. The wide variety of findings in the literature on AQP and apoptosis might be partially due to the fact that the timing of the biochemical events during cell death varies significantly depending on the cell type, drug concentration, apoptotic stimulus, and exposure time [161]. More systematic and time-dependent experimentation is needed to understand how these parameters might affect the function of AQPs in apoptosis, especially if their function changes during the course of cell death.
Less prominent but still fascinating is how AQPs have been implicated in a variety of other cellular processes. For example, cell adhesion onto a Matrigel-coated surface with respect to AQP expression has been estimated with absorbance readings after crystal violet staining of washed and fixed samples [83]. Atomic force microscopy (AFM) with a tipless cantilever can quantify the work required to overcome cell–cell adhesion and has shown that AQP knockdown cells have weaker cell–cell adhesion [10]. Cell aggregation and expression of membrane-associated junctional proteins are other metrics for intercellular adhesion, which can be visualized with ICC [18, 101]. Conversely, cell–cell dissociation has been observed with video microscopy of migrating epithelial cell sheets and quantified with a dispase-based dissociation assay [101].
Related to cell–cell adhesion is junction phenotype, which is especially relevant in the blood–brain barrier. Our lab has developed a novel Junction Analyzer Program (JAnaP) for phenotyping cell–cell junctions from IF images as an indicator of endothelial barrier integrity [102]. This method has not yet been used to investigate the effect of AQP expression or function on endothelial junction presentation and would be informative given that some AQPs are involved in cell–cell adhesion and junction formation [162, 163]. JAnaP is not limited to endothelial cells—epithelial monolayers can be analyzed as well. In the same vein, AQPs have been studied in the context of paracellular permeability in epithelial cells. Cell–cell junctions are anchored to cytoskeletal components including microtubules, and thus a cytoskeletal rearrangement can alter the permeability between the junctions. Microtubule stability has been linked to AQP5 expression via ICC visualization of microtubules, soluble and insoluble microtubule extraction, and a fluorescence-based microtubule polymerization assay [162].
Finally, AQP expression has been linked to angiogenesis in multiple cell types, and investigations have most commonly used tube formation assays [12, 13, 163, 105]. These are often performed in parallel with other cell migration, invasion, and proliferation assays [13, 163, 105–107] using endothelial cells such as HUVECs. The relationship between AQPs and other angiogenesis promoters, such as erythropoietin [108] and vascular endothelial growth factor [98], can also be examined using previously discussed western blotting, IF, or PPI detection methods. Again, these studies mainly provide some correlative insight on the effect of AQP expression on angiogenic cell behavior, but not AQP function. Using functional AQP mutants and AQP inhibitors in these methods would demonstrate the impact of AQP-triggered signaling cascades and AQP water blockade on angiogenesis, respectively.
In vivo methods for studying AQPs
Though in vitro methods are important for identifying the specific effects of individual cues on cells, it is essential to understand the translatability of these findings into complex in vivo systems. Here we provide an abridged compilation of the numerous ways in vivo models can be used to study AQP expression, AQP-induced cellular function, AQP-targeted drug development, and AQP-based anatomical imaging modalities.
Observing AQP expression and activity in vivo
The most simplistic studies conducted in vivo plainly observed AQP expression or localization in samples from animal models or patients, typically via immunohistochemistry [164–169]. Building upon this baseline, physiological AQP expression and localization can be compared to pathological tissue samples excised from patients or animal models. Common disease states studied for pathological changes in AQP dynamics include cancer (colorectal [170], lung [171], liver [96], breast [172]), diabetes [173, 174], arthritis [175], and chronic inflammation [176–181]. Other animal models can be employed to identify changes in AQP expression during events like development [182, 183].
Additionally, the opposite could be done, where AQPs are knocked out of in vivo models, and the functional changes that accompany this knockout can be observed. This idea was epitomized by much of the work done by Dr. Alan Verkman: their lab developed numerous AQP isoform knockout models and conducted experiments to identify differences among them [131, 184–188]. To our knowledge, murine knockout models have been created for the majority of AQP isoforms (AQP0, 1, 2, 3, 4, 5, 7, 8, 9, 11, and 12) [184–194], with additional models across other species. By observing the functional changes that occur amidst these knockout models, AQPs have been shown to regulate numerous functional pathways in the brain [184, 195–198], immune system [185, 191], the urinary tract [193, 199, 200], diabetes [194, 201–204], development [131, 205], and more [206, 207]. A functional correlation between AQPs and other pathways can be determined by developing knockout models of AQP-related proteins or regulators [208–213]. Changes in AQP functionality in the absence of these regulators provide a more holistic perspective of the complex interplay of signaling in vivo.
AQP modulation in vivo
After using the above methods to identify the abundant functional roles of AQPs across different systems, the next step is to identify if these roles can be modulated in vivo. Using molecular [214–218], siRNA [219], and immuno-based drugs [220], researchers have successfully inhibited AQP functionality or expression. In many cases, this inhibition reversed the burden and symptoms of the pathological system in question. Additionally, other molecular drugs [221] or viral vectors [222] have been used to increase the expression of AQPs, thereby enhancing their function. For instance, an adenoviralus vector with cDNA to increase AQP1 expression enhances saliva secretion in Wistar rats with hypofunctional salivary glands.
AQPs applied to imaging
Finally, AQP’s water transport function is a property that can be appropriated for the use of anatomical imaging. Diffusion-weighted imaging (DWI) uses a pair of pulsed magnetic field gradients to relate nuclear spin to diffusion; this can relate to water, as molecules with the ability to diffuse more appear darker. In cells, this diffusion is related to the water’s ability to move intracellularly and extracellularly, regulated by AQPs [223]. By tagging specific cells to overexpress AQP isoforms, DWI magnetic resonance imaging (MRI) is capable of tracking their dynamic behaviors in vivo in a noninvasive manner [224–226]. Aside from DWI analysis of MRI, other MRI methods have been used to identify functional AQP characteristics, like the tagging of O water molecules analyzed by JJ vicinal coupling proton exchange MRI [227] or AQP4 glymphatic drainage following gadoteric acid through t1 mapping [212]. Fluorescent, positron emission tomography, and hyperspectral imaging can also be utilized to identify the functional properties of AQPs [228, 229].
In silico methods for studying AQPs
While in vitro and in vivo methods have provided us with a wealth of knowledge on AQP expression and function in biological systems, most of our understanding of AQP structure and function on a molecular level is due to in silico work. Computational models and simulations have the obvious benefits of saving time and laboratory resources, but they are also the most effective way to handle large amounts of data. Molecular dynamics (MD) and molecular docking, while computationally expensive, have revealed crucial information regarding AQP structures, the mechanism of water transport, AQP drug candidates and their putative binding sites, and PPIs. Stochastic and deterministic mathematical cell models have also begun to shed light on the impact of AQPs on cell activities.
AQP structure and water flux
Much of our current understanding of AQP structure stems from X-ray crystallography and electron microscopy, but the structures of AQP3 [43], AQP7 [230], AQP9 [231], AQP11, and AQP12 [232] have additionally been predicted using structural bioinformatics in silico technique called homology or comparative modeling. This involves building a 3D model of a protein from its amino acid sequence and is compared to the known structures of a similar protein. More recently, machine learning has also been implemented in sequence-based AQP prediction [233, 234]. However, most of our initial knowledge of AQPs and AQP-mediated water flux comes from MD, a computational method in which the movements of atoms and molecules are simulated according to Newton’s equations of motion. Extensive MD simulations have been performed to elucidate AQP folding [9], proton exclusion from AQP pores [9, 235, 236], gating capabilities [237–239], ion conductivity [238, 240], voltage sensitivity [241, 242], lipid interactions [243, 244], permeability [28, 245–250], effects of mutations [251, 252], effects of phosphorylation [253, 254], and suppression of water flux by inhibitors [231, 255, 256]. For a more thorough review of MD simulations for the AQP study, we direct the reader to other reviews [257, 258]. Hand-in-hand with MD is meta-dynamics, which estimates the free energy of a system and simulates rare events, particularly accelerating the sampling of protein conformations within an MD simulation [259]. Meta-dynamics has been used to simulate water, glycerol, and H2O2 transport through AQP3 [260, 261], and to study the effects of electric fields on AQP4 [262, 263].
AQP–molecular interactions
Once protein structures are defined, in silico methods can provide much quicker insight into how AQPs interact with other molecules compared to in vitro methods. Predicting AQP–ligand interactions allows for expedited drug screening; AQP–protein interactions shed light on AQP regulation and signaling activities; and AQP–lipid interactions elucidate key aspects of plasma membrane properties and how they affect AQP function.
AQP drug screening
Over time, the revelation that AQPs play key roles in several diseased cell behaviors has driven the search for AQP drugs that can block water transport, glycerol permeation, or ion conductance. In vitro methods for drug validation can be tedious and time-consuming. Molecular docking is a computational tool that has been widely used to expedite the screening of AQP inhibitors [30, 231, 264–267], predicting how small molecules will fit with a protein at the atomic level. In most cases, the protein and ligand are prepared and docked using automated docking software such as AutoDock. Once inhibitors are identified with molecular docking, binding modes may be further analyzed with MD simulations [256, 264].
Identification of AQP–protein and AQP–lipid interactions
Molecular docking can predict PPIs in addition to protein–ligand interactions and has been used to model binding between AQP5 and ezrin [146]. While it would be desirable to discover more AQP–protein binding partners using molecular docking, it remains a challenge to manage the heavy computational cost of docking two proteins while capturing their conformational dynamics. However, a simpler method exists for a broader detection of putative PPIs; predictions can be made based on consensus sites in amino acid sequences [268] that can be compared using software such as ProteinPrompt [269]. To predict AQP–lipid interactions, structural AQP data from crystallography can be used in MD simulations. For example, MD studies have investigated lipid organization around AQP0—an isoform whose structure has been solved from crystallography—to better understand membrane properties, protein mobility, and cholesterol-mediated tetrameric assembly [70, 71].
Computational and mathematical cell models
Another in silico approach to studying AQPs is to model them within cells. To our knowledge, few studies have developed computational models showing the interplay of AQPs and cell processes. One group ran simulations of 2D neural crest cell migratory streams involving cell speed and filopodia dynamics associated with an AQP1 overexpression or downregulation [270]. Another study built a multiscale compartmental model of AQP2 trafficking via the vesicular transport system in renal principal cells. Their numerical simulations incorporate membrane agents, filaments, vesicles, and a rule-based reaction system for the chemical entities [271]. Numerical methods have also been used to model AQP1 in ion and fluid transport across parotid duct cells [272], AQP4 in brain swelling in meningitis [273], and AQP4 supramolecular assembly into orthogonal arrays within cell membranes [274]. Future work modeling AQPs in intracellular signaling pathways would elucidate how AQPs can trigger downstream cascades affecting cell migration, proliferation, differentiation, etc.
Concluding remarks and outlook
In this review, we cover frequently used in vitro, in vivo, and in silico methods for studying AQP expression, localization, function, effects on cell behavior, and more. While these methods have earned us a great deal of basic knowledge of AQP structure and functions, they still have their limitations in their robustness, reproducibility, and physiological relevance. Hopefully, future method design will include more sensitive permeability assays, more reproducible inhibitor screening, further characterization of AQPs within in vivo systems, and wider use of computational modeling tools. Since much of AQP biology research is performed in vitro, there is also a need for more physiologically relevant models that help clarify the role of AQPs in both healthy and pathologic cell behaviors. Addressing these areas for improvement for current methods, exploring new uses of existing methods, and designing novel methods may well lead to the discovery of more effective AQP-targeting therapeutics and a greater understanding of the impact of AQPs in biological processes.
Acknowledgments
Figure 1 was created using BioRender.com.
Contributor Information
Shohini Banerjee, Fischell Department of Bioengineering, University of Maryland, MD 20742, United States.
Ian M Smith, Fischell Department of Bioengineering, University of Maryland, MD 20742, United States.
Autumn C Hengen, Fischell Department of Bioengineering, University of Maryland, MD 20742, United States.
Kimberly M Stroka, Fischell Department of Bioengineering, University of Maryland, MD 20742, United States; Marlene and Stewart Greenebaum Comprehensive Cancer Center, University of Maryland, Baltimore MD 21201, United States; Biophysics Program, University of Maryland, MD 20742, United States; Center for Stem Cell Biology and Regenerative Medicine, University of Maryland, Baltimore MD 21201, United States.
Disclaimers
The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.
Authors' contributions
Shohini Banerjee (Conceptualization [lead], Visualization [equal], Writing—original draft [lead], Writing—review & editing [lead]), Ian M. Smith (Conceptualization [supporting], Writing—original draft [supporting], Writing—review & editing [equal]), Autumn C. Hengen (Visualization [lead], Writing—review & editing [supporting]), and Kimberly Stroka (Conceptualization [supporting], Funding acquisition [lead], Project administration [lead], Supervision [lead], Writing—review & editing [equal]).
Conflict of interest statement. No conflicts of interest, financial or otherwise, are declared by the authors.
Funding
The authors acknowledge funding from an NIGMS MIRA #R35GM142838 (to K.M.S.), from the Clark Doctoral Fellowship (to S.B. and I.M.S.), and from a Maryland Summer Scholar Award (to A.C.H.) and MTech ASPIRE Award (to A.C.H.). Research reported in this publication was supported by the National Institute of General Medical Sciences of the National Institutes of Health under Award Number R35GM142838.
Data availability
No new data were generated or analysed in support of this research.
References
- 1. Azad AK, Raihan T, Ahmed J. et al. Human aquaporins: functional diversity and potential roles in infectious and non-infectious diseases. Front Genet 2021;12:654865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Ishibashi K, Tanaka Y, Morishita Y.. The role of mammalian superaquaporins inside the cell. Biochim Biophys Acta 2014;1840:1507–12. [DOI] [PubMed] [Google Scholar]
- 3. Smith IM, Banerjee S, Moses AK. et al. Aquaporin-mediated dysregulation of cell migration in disease states. Cell Mol Life Sci 2023;80:48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Verkman AS. Aquaporins in clinical medicine. Annu Rev Med 2012;63:303–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Preston GM, Carroll TP, Guggino WB. et al. Appearance of water channels in Xenopus oocytes expressing red cell CHIP28 protein. Science 1992;256:385–7. [DOI] [PubMed] [Google Scholar]
- 6. Carbrey JM, Agre P.. Discovery of the aquaporins and development of the field. Handb Exp Pharmacol 2009;190:3–28. [DOI] [PubMed] [Google Scholar]
- 7. Andrews S, Reichow SL, Gonen T.. Electron crystallography of aquaporins. IUBMB Life 2008;60:430–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Wang S, Ing C, Emami S. et al. Structure and dynamics of extracellular loops in human aquaporin-1 from solid-state NMR and molecular dynamics. J Phys Chem B 2016;120:9887–902. [DOI] [PubMed] [Google Scholar]
- 9. Murata K, Mitsuoka K, Hirai T. et al. Structural determinants of water permeation through aquaporin-1. Nature 2000;407:599–605. [DOI] [PubMed] [Google Scholar]
- 10. Silva PM, da Silva IV, Sarmento MJ. et al. Aquaporin-3 and aquaporin-5 facilitate migration and cell–cell adhesion in pancreatic cancer by modulating cell biomechanical properties. Cells 2022;11:1308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Miller EW, Dickinson BC, Chang CJ.. Aquaporin-3 mediates hydrogen peroxide uptake to regulate downstream intracellular signaling. Proc Natl Acad Sci USA 2010;107:15681–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Elkhider A, Wang B, Ouyang X. et al. Aquaporin 5 promotes tumor migration and angiogenesis in non-small cell lung cancer cell line H1299. Oncol Lett 2020;19:1665–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Zou LB, Shi S, Zhang RJ. et al. Aquaporin-1 plays a crucial role in estrogen-induced tubulogenesis of vascular endothelial cells. J Clin Endocrinol Metab 2013;98:E672–82. [DOI] [PubMed] [Google Scholar]
- 14. Woo J, Lee J, Kim MS. et al. The effect of aquaporin 5 overexpression on the Ras signaling pathway. Biochem Biophys Res Commun 2008;367:291–8. [DOI] [PubMed] [Google Scholar]
- 15. Karlsson T, Bolshakova A, Magalhães MAO. et al. Fluxes of water through aquaporin 9 weaken membrane-cytoskeleton anchorage and promote formation of membrane protrusions. PLoS One 2013;8:e59901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Stroka KM, Jiang H, Chen SH. et al. Water permeation drives tumor cell migration in confined microenvironments. Cell 2014;157:611–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Wan S, Jiang J, Zheng C. et al. Estrogen nuclear receptors affect cell migration by altering sublocalization of AQP2 in glioma cell lines. Cell Death Discov 2018;4:49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Min S, Choe C, Roh S.. AQP3 increases intercellular cohesion in NSCLC A549 cell spheroids through exploratory cell protrusions. Int J Mol Sci 2021;22:4287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Umenishi F, Verbavatz JM, Verkman AS.. cAMP regulated membrane diffusion of a green fluorescent protein-aquaporin 2 chimera. Biophys J 2000;78:1024–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Kosugi-Tanaka C, Li X, Yao C. et al. Protein kinase A-regulated membrane trafficking of a green fluorescent protein-aquaporin 5 chimera in MDCK cells. Biochim Biophys Acta 2006;1763:337–44. [DOI] [PubMed] [Google Scholar]
- 21. Jiang Y, Wang C, Ma R. et al. Aquaporin 1 mediates early responses to osmotic stimuli in endothelial cells via the calmodulin pathway. FEBS Open Bio 2021;11:75–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Moeller HB, Praetorius J, Rützler MR. et al. Phosphorylation of aquaporin-2 regulates its endocytosis and protein–protein interactions. Proc Natl Acad Sci USA 2010;107:424–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. van Balkom BWM, Savelkoul PJM, Markovich D. et al. The role of putative phosphorylation sites in the targeting and shuttling of the aquaporin-2 water channel. J Biol Chem 2002;277:41473–9. [DOI] [PubMed] [Google Scholar]
- 24. Yue K, Jiang J, Zhang P. et al. Functional analysis of aquaporin water permeability using an Escherichia coli-based cell-free protein synthesis system. Front Bioeng Biotechnol 2020;8:1000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Laforenza U, Scaffino MF, Gastaldi G.. Aquaporin-10 represents an alternative pathway for glycerol efflux from human adipocytes. PLoS One 2013;8:e54474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Rodrigues C, Pimpão C, Mósca AF. et al. Human aquaporin-5 facilitates hydrogen peroxide permeation affecting adaption to oxidative stress and cancer cell migration. Cancers 2019;11:932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Yang B, Zhao D, Verkman AS.. Evidence against functionally significant aquaporin expression in mitochondria. J Biol Chem 2006;281:16202–6. [DOI] [PubMed] [Google Scholar]
- 28. Rodriguez RA, Liang H, Chen LY. et al. Single-channel permeability and glycerol affinity of human aquaglyceroporin AQP3. Biochim Biophys Acta Biomembr 2019;1861:768–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Watanabe S, Moniaga CS, Nielsen S. et al. Aquaporin-9 facilitates membrane transport of hydrogen peroxide in mammalian cells. Biochem Biophys Res Commun 2016;471:191–7. [DOI] [PubMed] [Google Scholar]
- 30. Pei JV, Kourghi M, De Ieso ML. et al. Differential inhibition of water and ion channel activities of mammalian aquaporin-1 by two structurally related bacopaside compounds derived from the medicinal plant Bacopa monnieri. Mol Pharmacol 2016;90:496–507. [DOI] [PubMed] [Google Scholar]
- 31. Yang B, Verkman AS.. Water and glycerol permeabilities of aquaporins 1–5 and MIP determined quantitatively by expression of epitope-tagged constructs inXenopus oocytes. J Biol Chem 1997;272:16140–6. [DOI] [PubMed] [Google Scholar]
- 32. Fenton RA, Moeller HB, Nielsen S. et al. A plate reader-based method for cell water permeability measurement. Am J Physiol Renal Physiol 2010;298:F224–30. [DOI] [PubMed] [Google Scholar]
- 33. Mola MG, Nicchia GP, Svelto M. et al. Automated cell-based assay for screening of aquaporin inhibitors. Anal Chem 2009;81:8219–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Levin MH, Sullivan S, Nielson D. et al. Hypertonic saline therapy in cystic fibrosis. J Biol Chem 2006;281:25803–12. [DOI] [PubMed] [Google Scholar]
- 35. To J, Yeo CY, Soon CH. et al. A generic high-throughput assay to detect aquaporin functional mutants: potential application to discovery of aquaporin inhibitors. Biochim Biophys Acta 2015;1850:1869–76. [DOI] [PubMed] [Google Scholar]
- 36. Groszmann M, De Rosa A, Chen W. et al. A high-throughput yeast approach to characterize aquaporin permeabilities: profiling the Arabidopsis PIP aquaporin sub-family. Front Plant Sci 2023;14:1078220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Kourghi M, Pei JV, De Ieso ML. et al. Bumetanide derivatives AqB007 and AqB011 selectively block the aquaporin-1 ion channel conductance and slow cancer cell migration. Mol Pharmacol 2016;89:133–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Kourghi M, Nourmohammadi S, Pei JV. et al. Divalent cations regulate the ion conductance properties of diverse classes of aquaporins. Int J Mol Sci 2017;18:2323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Hazama A, Kozono D, Guggino WB. et al. Ion permeation of AQP6 water channel protein. J Biol Chem 2002;277:29224–30. [DOI] [PubMed] [Google Scholar]
- 40. Saparov SM, Kozono D, Rothe U. et al. Water and ion permeation of aquaporin-1 in planar lipid bilayers. J Biol Chem 2001;276:31515–20. [DOI] [PubMed] [Google Scholar]
- 41. Boassa D, Stamer WD, Yool AJ.. Ion channel function of aquaporin-1 natively expressed in choroid plexus. J Neurosci 2006;26:7811–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Wu B, Beitz E.. Aquaporins with selectivity for unconventional permeants. Cell Mol Life Sci 2007;64:2413–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Brooks HL, Regan JW, Yool AJ.. Inhibition of aquaporin-1 water permeability by tetraethylammonium: involvement of the loop E pore region. Mol Pharmacol 2000;57:1021–6. [PubMed] [Google Scholar]
- 44. Martins AP, Marrone A, Ciancetta A. et al. Targeting aquaporin function: potent inhibition of aquaglyceroporin-3 by a gold-based compound. PLoS One 2012;7:e37435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Tomita Y, Palethorpe HM, Smith E. et al. Bumetanide-derived aquaporin 1 inhibitors, AqB013 and AqB050 inhibit tube formation of endothelial cells through induction of apoptosis and impaired migration in vitro. Int J Mol Sci 2019;20:1818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. De Ieso ML, Pei JV, Nourmohammadi S. et al. Combined pharmacological administration of AQP1 ion channel blocker AqB011 and water channel blocker Bacopaside II amplifies inhibition of colon cancer cell migration. Sci Rep 2019;9:12635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Palethorpe HM, Smith E, Tomita Y. et al. Bacopasides I and II act in synergy to inhibit the growth, migration and invasion of breast cancer cell lines. Molecules 2019;24:3539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Kitchen P, Salman MM, Halsey AM. et al. Targeting aquaporin-4 subcellular localization to treat central nervous system edema. Cell 2020;181:784–99.e19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Lin F, Tang R, Zhang C. et al. Combining different ion-selective channelrhodopsins to control water flux by light. Pflugers Arch - Eur J Physiol 2023. 10.1007/s00424-023-02853-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Monzani E, Bazzotti R, Perego C. et al. AQP1 is not only a water channel: it contributes to cell migration through lin7/beta-catenin. PLoS One 2009;4:e6167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Bhattacharya D, Yu L, Wang M.. Expression patterns of conjunctival mucin 5AC and aquaporin 5 in response to acute dry eye stress. PLoS One 2017;12:e0187188. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. De Bellis M, Pisani F, Mola MG. et al. A novel human aquaporin-4 splice variant exhibits a dominant-negative activity: a new mechanism to regulate water permeability. Mol Biol Cell 2014;25:470–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Benfenati V, Caprini M, Dovizio M. et al. An aquaporin-4/transient receptor potential vanilloid 4 (AQP4/TRPV4) complex is essential for cell-volume control in astrocytes. Proc Natl Acad Sci USA 2011;108:2563–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Reppetti J, Reca A, Seyahian EA. et al. Intact caveolae are required for proper extravillous trophoblast migration and differentiation. J Cell Physiol 2020;235:3382–92. [DOI] [PubMed] [Google Scholar]
- 55. Krüger C, Waldeck-Weiermair M, Kaynert J. et al. AQP8 is a crucial H2O2 transporter in insulin-producing RINm5F cells. Redox Biol 2021;43:101962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. van Balkom BWM, Boone M, Hendriks G. et al. LIP5 interacts with aquaporin 2 and facilitates its lysosomal degradation. J Am Soc Nephrol 2009;20:990–1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Toft‐Bertelsen TL, Križaj D, MacAulay N.. When size matters: transient receptor potential vanilloid 4 channel as a volume‐sensor rather than an osmo‐sensor. J Physiol 2017;595:3287–302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Galizia L, Pizzoni A, Fernandez J. et al. Functional interaction between AQP2 and TRPV4 in renal cells. J Cell Biochem 2012;113:580–9. [DOI] [PubMed] [Google Scholar]
- 59. Lindsey Rose KM, Wang Z, Magrath GN. et al. Aquaporin 0−Calmodulin interaction and the effect of aquaporin 0 phosphorylation. Biochemistry 2008;47:339–47. [DOI] [PubMed] [Google Scholar]
- 60. Wang Z, Schey KL.. Aquaporin-0 interacts with the FERM domain of ezrin/radixin/moesin proteins in the ocular lens. Invest Ophthalmol Vis Sci 2011;52:5079–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Wang Z, Schey KL.. Identification of a direct aquaporin-0 binding site in the lens-specific cytoskeletal protein filensin. Exp Eye Res 2017;159:23–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Wang Z, Friedrich MG, Truscott RJW. et al. Cleavage C-terminal to Asp leads to covalent crosslinking of long-lived human proteins. Biochim Biophys Acta Proteins Proteom 2019;1867:831–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Pietrement C, Da Silva N, Silberstein C. et al. Role of NHERF1, cystic fibrosis transmembrane conductance regulator, and cAMP in the regulation of aquaporin 9. J Biol Chem 2008;283:2986–96. [DOI] [PubMed] [Google Scholar]
- 64. Misawa T, Arima K, Mizusawa H, Satoh J.. Close association of water channel AQP1 with amyloid-β deposition in Alzheimer disease brains. Acta Neuropathol 2008;116:247–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Roche JV, Törnroth-Horsefield S.. Aquaporin protein-protein interactions. Int J Mol Sci 2017;18:2255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Wang Z, Wang Y, He Y. et al. Aquaporin-1 facilitates proliferation and invasion of gastric cancer cells via GRB7-mediated ERK and Ras activation. Anim Cells Syst (Seoul) 2020;24:253–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Kamsteeg E-J, Hendriks G, Boone M. et al. Short-chain ubiquitination mediates the regulated endocytosis of the aquaporin-2 water channel. Proc Natl Acad Sci USA 2006;103:18344–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Gonen T, Cheng Y, Sliz P. et al. Lipid–protein interactions in double-layered two-dimensional AQP0 crystals. Nature 2005;438:633–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Hite RK, Gonen T, Harrison SC, Walz T.. Interactions of lipids with aquaporin-0 and other membrane proteins. Pflugers Arch 2008;456:651–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Chiu PL, Orjuela JD, De Groot BL. et al. Structure and dynamics of cholesterol-mediated aquaporin-0 arrays and implications for lipid rafts. Elife 2023;12:RP90851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Briones R, Aponte-Santamaría C, De Groot BL.. Localization and ordering of lipids around aquaporin-0: protein and lipid mobility effects. Front Physiol 2017;8:124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Hoffert JD, Pisitkun T, Wang G. et al. Quantitative phosphoproteomics of vasopressin-sensitive renal cells: regulation of aquaporin-2 phosphorylation at two sites. Proc Natl Acad Sci USA May 2006;103:7159–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Deshpande V, Kao A, Raghuram V. et al. Phosphoproteomic identification of vasopressin V2 receptor-dependent signaling in the renal collecting duct. Am J Physiol Renal Physiol 2019;317:F789–F804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Hyndman KA, Yang CR, Jung HJ. et al. Proteomic determination of the lysine acetylome and phosphoproteome in the rat native inner medullary collecting duct. Physiol Genomics 2018;50:669–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Hendriks G, Koudijs M, van Balkom BWM. et al. Glycosylation is important for cell surface expression of the water channel aquaporin-2 but is not essential for tetramerization in the endoplasmic reticulum. J Biol Chem 2004;279:2975–83. [DOI] [PubMed] [Google Scholar]
- 76. Cui D, Sui L, Han X. et al. Aquaporin-3 mediates ovarian steroid hormone-induced motility of endometrial epithelial cells. Hum Reprod 2018;33:2060–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Gau DM, Roy P.. Single cell migration assay using human breast cancer MDA-MB-231 cell line. Bio-Protoc 2020;10:e3586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Pranda MA, Gray KM, DeCastro AJL. et al. Tumor cell mechanosensing during incorporation into the brain microvascular endothelium. Cell Mol Bioeng 2019;12:455–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Hetmanski JHR, de Belly H, Busnelli I. et al. Membrane tension orchestrates rear retraction in matrix-directed cell migration. Dev Cell 2019;51:460–75.e10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Zhang Y, Li Y, Thompson KN. et al. Polarized NHE1 and SWELL1 regulate migration direction, efficiency and metastasis. Nat Commun 2022;13:6128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Bera K, Kiepas A, Godet I. et al. Extracellular fluid viscosity enhances cell migration and cancer dissemination. Nature 2022;611:365–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Hui TH, Cho WC, Fong HW. et al. An electro-osmotic microfluidic system to characterize cancer cell migration under confinement. J R Soc Interface 2019;16:20190062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Zou LB, Zhang RJ, Tan YJ. et al. Identification of estrogen response element in the aquaporin-2 gene that mediates estrogen-induced cell migration and invasion in human endometrial carcinoma. J Clin Endocrinol Metab 2011;96:E1399–408. [DOI] [PubMed] [Google Scholar]
- 84. Jensen HH, Holst MR, Login FH. et al. Ectopic expression of aquaporin-5 in noncancerous epithelial MDCK cells changes cellular morphology and actin fiber formation without inducing epithelial-to-mesenchymal transition. Am J Physiol Cell Physiol 2018;314:C654–C661. [DOI] [PubMed] [Google Scholar]
- 85. Thiagarajah JR, Chang J, Goettel JA. et al. Aquaporin-3 mediates hydrogen peroxide-dependent responses to environmental stress in colonic epithelia. Proc Natl Acad Sci USA 2017;114:568–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Zhang WG, Li CF, Liu M. et al. Aquaporin 9 is down-regulated in hepatocellular carcinoma and its over-expression suppresses hepatoma cell invasion through inhibiting epithelial-to-mesenchymal transition. Cancer Lett 2016;378:111–9. [DOI] [PubMed] [Google Scholar]
- 87. Hayashi S, Takahashi N, Kurata N. et al. Involvement of aquaporin-1 in gastric epithelial cell migration during wound repair. Biochem Biophys Res Commun 2009;386:483–7. [DOI] [PubMed] [Google Scholar]
- 88. Di Giusto G, Pizzoni A, Rivarola V. et al. Aquaporin‐2 and Na+/H+ exchanger isoform 1 modulate the efficiency of renal cell migration. J Cell Physiol 2020;235:4443–54. [DOI] [PubMed] [Google Scholar]
- 89. Hara-Chikuma M, Verkman AS.. Aquaporin-1 facilitates epithelial cell migration in kidney proximal tubule. J Am Soc Nephrol 2006;17:39–45. [DOI] [PubMed] [Google Scholar]
- 90. Loitto VM, Huang C, Sigal YJ. et al. Filopodia are induced by aquaporin-9 expression. Exp Cell Res 2007;313:1295–306. [DOI] [PubMed] [Google Scholar]
- 91. Jung HJ, Park JY, Jeon HS. et al. Aquaporin-5: a marker protein for proliferation and migration of human breast cancer cells. PLoS One 2011;6:e28492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Woo J, Lee J, Chae YK. et al. Overexpression of AQP5, a putative oncogene, promotes cell growth and transformation. Cancer Lett 2008;264:54–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Wei X, Dong J.. Aquaporin 1 promotes the proliferation and migration of lung cancer cell in vitro. Oncol Rep 2015;34:1440–8. [DOI] [PubMed] [Google Scholar]
- 94. Jablonski EM, Webb AN, McConnell NA. et al. Plasma membrane aquaporin activity can affect the rate of apoptosis but is inhibited after apoptotic volume decrease. Am J Physiol Cell Physiol 2004;286:C975–C985. [DOI] [PubMed] [Google Scholar]
- 95. Lakner AM, Walling TL, McKillop IH. et al. Altered aquaporin expression and role in apoptosis during hepatic stellate cell activation: aquaporin expression and function in HSCs. Liver Int 2011;31:42–51. [DOI] [PubMed] [Google Scholar]
- 96. Jablonski EM, Mattocks MA, Sokolov E. et al. Decreased aquaporin expression leads to increased resistance to apoptosis in hepatocellular carcinoma. Cancer Lett 2007;250:36–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Ding T, Zhou Y, Sun K. et al. Knockdown a water channel protein, aquaporin-4, induced glioblastoma cell apoptosis. PLoS One 2013;8:e66751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Shu C, Shu Y, Gao Y. et al. Inhibitory effect of AQP1 silencing on adhesion and angiogenesis in ectopic endometrial cells of mice with endometriosis through activating the Wnt signaling pathway. Cell Cycle 2019;18:2026–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Chu H, Xiang J, Wu P. et al. The role of aquaporin 4 in apoptosis after intracerebral hemorrhage. J Neuroinflammation 2014;11:184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Kannan A, Anbarasu K, Mohideen AP. et al. Role of aquaporin 9 in hyperglycaemia-induced testicular leydig cell apoptosis. J King Saud Univ Sci 2022;34:102160. [Google Scholar]
- 101. Login FH, Jensen HH, Pedersen GA. et al. Aquaporins differentially regulate cell‐cell adhesion in MDCK cells. FASEB J 2019;33:6980–94. [DOI] [PubMed] [Google Scholar]
- 102. Gray KM, Katz DB, Brown EG. et al. Quantitative phenotyping of cell–cell junctions to evaluate ZO-1 presentation in brain endothelial cells. Ann Biomed Eng 2019;47:1675–87. [DOI] [PubMed] [Google Scholar]
- 103. Sidhaye VK, Chau E, Srivastava V. et al. A novel role for aquaporin-5 in enhancing microtubule organization and stability. PLoS One 2012;7:e38717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Palethorpe HM, Tomita Y, Smith E. et al. The aquaporin 1 inhibitor bacopaside II reduces endothelial cell migration and tubulogenesis and induces apoptosis. Int J Mol Sci 2018;19:653. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Saadoun S, Papadopoulos MC, Hara-Chikuma M. et al. Impairment of angiogenesis and cell migration by targeted aquaporin-1 gene disruption. Nature 2005;434:786–92. [DOI] [PubMed] [Google Scholar]
- 106. Huebert RC, Vasdev MM, Shergill U. et al. Aquaporin-1 facilitates angiogenic invasion in the pathological neovasculature that accompanies cirrhosis. Hepatology 2010;52:238–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Huebert RC, Jagavelu K, Hendrickson HI. et al. Aquaporin-1 promotes angiogenesis, fibrosis, and portal hypertension through mechanisms dependent on osmotically sensitive MicroRNAs. Am J Pathol 2011;179:1851–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Maltaneri RE, Schiappacasse A, Chamorro ME. et al. Aquaporin-1 plays a key role in erythropoietin-induced endothelial cell migration. Biochim Biophys Acta Mol Cell Res 2020;1867:118569. [DOI] [PubMed] [Google Scholar]
- 109. Han H. RNA interference to knock down gene expression. Methods Mol Biol 2018;1706:293–302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Taxman DJ, Moore CB, Guthrie EH. et al. Short hairpin RNA (shRNA): design, delivery, and assessment of gene knockdown. In Sioud M. (ed.), RNA Therapeutics, vol. 629, Methods in Molecular Biology, vol. 629. Totowa, NJ: Humana Press, 2010, 139–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. He R, Han W, Hu Y. et al. AQP2 is regulated by estradiol in human endometrium and is associated with spheroid attachment in vitro. Mol Med Rep 2019;20:1306–12. [DOI] [PubMed] [Google Scholar]
- 112. Liu W, Wang K, Gong K. et al. Epidermal growth factor enhances MPC-83 pancreatic cancer cell migration through the upregulation of aquaporin 3. Mol Med Rep 2012;6:607–10. [DOI] [PubMed] [Google Scholar]
- 113. Hasler U, Vinciguerra M, Vandewalle A. et al. Dual effects of hypertonicity on aquaporin-2 expression in cultured renal collecting duct principal cells. J Am Soc Nephrol 2005;16:1571–82. [DOI] [PubMed] [Google Scholar]
- 114. Huo Z, Lomora M, Kym U. et al. AQP1 is up-regulated by hypoxia and leads to increased cell water permeability, motility, and migration in neuroblastoma. Front Cell Dev Biol 2021;9:605272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Luo L, Yang R, Zhao S. et al. Decreased miR-320 expression is associated with breast cancer progression, cell migration, and invasiveness via targeting Aquaporin 1. Acta Biochim Biophys Sin (Shanghai) 2018;50:473–80. [DOI] [PubMed] [Google Scholar]
- 116. Shihan MH, Novo SG, Le Marchand SJ. et al. A simple method for quantitating confocal fluorescent images. Biochem Biophys Rep 2021;25:100916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Maekawa C, Kitahara T, Kizawa K. et al. Expression and translocation of aquaporin-2 in the endolymphatic sac in patients with Meniere’s disease: AQP2 in Meniere’s disease. J Neuroendocrinol 2010;22:1157–64. [DOI] [PubMed] [Google Scholar]
- 118. Nesverova V, Törnroth-Horsefield S.. Phosphorylation-dependent regulation of mammalian aquaporins. Cells 2019;8:82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Kreda SM, Gynn MC, Fenstermacher DA. et al. Expression and localization of epithelial aquaporins in the adult human lLung. Am J Respir Cell Mol Biol 2001;24:224–34. [DOI] [PubMed] [Google Scholar]
- 120. Schey KL, Wang Z, Wenke JL, Qi Y.. Aquaporins in the eye: expression, function, and roles in ocular disease. Biochim Biophys Acta 2014;1840:1513–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Gao J, Tan M, Gu M. et al. Cellular localization of aquaporin-1 in the human and mouse trigeminal systems. PLoS One 2012;7:e46379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Yasui M, Kwon T-H, Knepper MA. et al. Aquaporin-6: an intracellular vesicle water channel protein in renal epithelia. Proc Natl Acad Sci USA 1999;96:5808–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Pellavio G, Martinotti S, Patrone M. et al. Aquaporin-6 may increase the resistance to oxidative stress of malignant pleural mesothelioma cells. Cells 2022;11:1892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Choi YS, Park JH, Yoon JK. et al. Potential roles of aquaporin 9 in the pathogenesis of endometriosis. Mol Hum Reprod 2019;25:373–84. [DOI] [PubMed] [Google Scholar]
- 125. Mlinarić M, Lučić I, Milković L. et al. AQP3-dependent PI3K/Akt modulation in breast cancer cells. Int J Mol Sci 2023;24:8133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Jiang XX, Fei XW, Zhao L. et al. Aquaporin 5 plays a role in estrogen-induced ectopic implantation of endometrial stromal cells in endometriosis. PLoS One 2015;10:e0145290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Abir-Awan M, Kitchen P, Salman MM. et al. Inhibitors of mammalian aquaporin water channels. Int J Mol Sci 2019;20:1589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Wang S, Solenov EI, Yang B.. Aquaporin inhibitors. Adv Exp Med Biol 2023;1398:317–30. [DOI] [PubMed] [Google Scholar]
- 129. Madeira A, Moura TF, Soveral G.. Detecting aquaporin function and regulation. Front Chem 2016;4:3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Madeira A, Camps M, Zorzano A. et al. Biophysical assessment of human aquaporin-7 as a water and glycerol channel in 3T3-L1 adipocytes. PLoS One 2013;8:e83442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Levin MH, Verkman AS.. Aquaporin-3-dependent cell migration and proliferation during corneal re-epithelialization. Invest Ophthalmol Vis Sci 2006;47:4365–72. [DOI] [PubMed] [Google Scholar]
- 132. Kitchen P, Salman MM, Abir-Awan M. et al. Calcein fluorescence quenching to measure plasma membrane water flux in live mammalian cells. STAR Protoc 2020;1:100157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133. Campbell EM, Birdsell DN, Yool AJ.. The activity of human aquaporin 1 as a cGMP-gated cation channel is regulated by tyrosine phosphorylation in the carboxyl-terminal domain. Mol Pharmacol 2012;81:97–105. [DOI] [PubMed] [Google Scholar]
- 134. Anthony TL, Brooks HL, Boassa D. et al. Cloned human aquaporin-1 is a cyclic GMP-gated ion channel. Mol Pharmacol 2000;57:576–88. [DOI] [PubMed] [Google Scholar]
- 135. Boassa D, Yool AJ.. Single amino acids in the carboxyl terminal domain of aquaporin-1 contribute to cGMP-dependent ion channel activation. BMC Physiol 2003;3:12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Lee JW, Zhang Y, Weaver CD. et al. Phosphorylation of nodulin 26 on serine 262 affects its voltage-sensitive channel activity in planar lipid bilayers. J Biol Chem 1995;270:27051–7. [DOI] [PubMed] [Google Scholar]
- 137. Abrami L, Berthonaud V, Rousselet G. et al. Glycerol permeability of mutant aquaporin 1 and other AQP-MIP proteins: inhibition studies. Pflugers Arch 1996;431:408–14. [DOI] [PubMed] [Google Scholar]
- 138. Salman MM, Kitchen P, Yool AJ, Bill RM.. Recent breakthroughs and future directions in drugging aquaporins. Trends Pharmacol Sci 2022;43:30–42. [DOI] [PubMed] [Google Scholar]
- 139. Yool AJ, Campbell EM.. Structure, function and translational relevance of aquaporin dual water and ion channels. Mol Aspects Med 2012;33:553–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Benedetti L. Optogenetic tools for manipulating protein subcellular localization and intracellular signaling at organelle contact sites. Curr Protoc 2021;1:e71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141. Roche JV, Survery S, Kreida S. et al. Phosphorylation of human aquaporin 2 (AQP2) allosterically controls its interaction with the lysosomal trafficking protein LIP5. J Biol Chem 2017;292:14636–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Ohashi Y, Tsuzaka K, Takeuchi T. et al. Altered distribution of aquaporin 5 and its C-terminal binding protein in the lacrimal glands of a mouse model for Sjögren’s syndrome. Curr Eye Res 2008;33:621–9. [DOI] [PubMed] [Google Scholar]
- 143. Hansen JS, Krintel C, Hernebring M. et al. Perilipin 1 binds to aquaporin 7 in human adipocytes and controls its mobility via protein kinase A mediated phosphorylation. Metabolism 2016;65:1731–42. [DOI] [PubMed] [Google Scholar]
- 144. Kreida S, Roche JV, Olsson C. et al. Protein–protein interactions in AQP regulation—biophysical characterization of AQP0–CaM and AQP2–LIP5 complex formation. Faraday Discuss 2018;209:35–54. [DOI] [PubMed] [Google Scholar]
- 145. Chivasso C, Nesverova V, Järvå M. et al. Unraveling human AQP5-PIP molecular interaction and effect on AQP5 salivary glands localization in SS patients. Cells 2021;10:2108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Chivasso C, Hagströmer CJ, Rose KL. et al. Ezrin is a novel protein partner of aquaporin-5 in human salivary glands and shows altered expression and cellular localization in Sjögren’s syndrome. Int J Mol Sci 2021;22:9213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Törnroth-Horsefield S, Chivasso C, Strandberg H. et al. Insight into the mammalian aquaporin interactome. Int J Mol Sci 2022;23:9615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Sjöhamn J, Hedfalk K.. Unraveling aquaporin interaction partners. Biochim Biophys Acta 2014;1840:1614–23. [DOI] [PubMed] [Google Scholar]
- 149. Lundby A, Secher A, Lage K. et al. Quantitative maps of protein phosphorylation sites across 14 different rat organs and tissues. Nat Commun 2012;3:876. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Hornbeck PV, Zhang B, Murray B. et al. PhosphoSitePlus, 2014: mutations, PTMs and recalibrations. Nucleic Acids Res 2015;43:D512–D520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151. Han Z, Patil RV.. Protein kinase A-dependent phosphorylation of aquaporin-1. Biochem Biophys Res Commun 2000;273:328–32. [DOI] [PubMed] [Google Scholar]
- 152. Han Z, Wax MB, Patil RV.. Regulation of aquaporin-4 water channels by phorbol ester-dependent protein phosphorylation. J Biol Chem 1998;273:6001–4. [DOI] [PubMed] [Google Scholar]
- 153. Lumpkin RJ, Gu H, Zhu Y. et al. Site-specific identification and quantitation of endogenous SUMO modifications under native conditions. Nat Commun 2017;8:1171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Smith IM, Stroka KM.. The multifaceted role of aquaporins in physiological cell migration. Am J Physiol Cell Physiol 2023;325:C208–C223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Carey SP, Starchenko A, McGregor AL. et al. Leading malignant cells initiate collective epithelial cell invasion in a three-dimensional heterotypic tumor spheroid model. Clin Exp Metastasis 2013;30:615–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Paul CD, Hung WC, Wirtz D. et al. Engineered models of confined cell migration. Annu Rev Biomed Eng 2016;18:159–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Huang Y-T, Zhou J, Shi S. et al. Identification of estrogen response element in aquaporin-3 gene that mediates estrogen-induced cell migration and invasion in estrogen receptor-positive breast cancer. Sci Rep 2015;5:12484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Galán-Cobo A, Ramírez-Lorca R, Echevarría M.. Role of aquaporins in cell proliferation: what else beyond water permeability. Channels (Austin) 2016;10:185–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159. Shimizu H, Shiozaki A, Ichikawa D. et al. The expression and role of Aquaporin 5 in esophageal squamous cell carcinoma. J Gastroenterol 2014;49:655–66. [DOI] [PubMed] [Google Scholar]
- 160. Das A, Banik NL, Ray SK.. Modulatory effects of acetazolomide and dexamethasone on temozolomide mediated apoptosis in human glioblastoma T98G and U87MG Cells”. Cancer Invest 2008;26:352–8. [DOI] [PubMed] [Google Scholar]
- 161. Sundquist T, Moravec R, Niles A. et al. Timing your apoptosis assays. Cell Notes 2006;16:18–21. [Google Scholar]
- 162. Gonen T, Cheng Y, Kistler J. et al. Aquaporin-0 membrane junctions form upon proteolytic cleavage. J Mol Biol 2004;342:1337–45. [DOI] [PubMed] [Google Scholar]
- 163. Hiroaki Y, Tani K, Kamegawa A. et al. Implications of the aquaporin-4 structure on array formation and cell adhesion. J Mol Biol 2006;355:628–39. [DOI] [PubMed] [Google Scholar]
- 164. Jungersted JM, Bomholt J, Bajraktari N. et al. In vivo studies of aquaporins 3 and 10 in human stratum corneum. Arch Dermatol Res 2013;305:699–704. [DOI] [PubMed] [Google Scholar]
- 165. Falco MD et al. Down-regulation of aquaporin 4 in human placenta throughout pregnancy. In Vivo 2007;21:813–7. [PubMed] [Google Scholar]
- 166. Satoh J, Tabunoki H, Yamamura T. et al. Human astrocytes express aquaporin-1 and aquaporin-4 in vitro and in vivo. Neuropathology 2007;27:245–56. [DOI] [PubMed] [Google Scholar]
- 167. Manso M, Drake MJ, Fry CH. et al. Expression and localization of aquaporin water channels in adult pig urinary bladder. J Cell Mol Med 2019;23:3772–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Varadaraj K, Kumari SS, Mathias RT.. Functional expression of aquaporins in embryonic, postnatal, and adult mouse lenses. Dev Dyn 2007;236:1319–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169. Gletten RB, Cantrell LS, Bhattacharya S. et al. Lens aquaporin-5 inserts into bovine fiber cell plasma membranes via unconventional protein secretion. Invest Ophthalmol Vis Sci 2022;63:5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Zhang G, Hao Y, Chen L. et al. Expression of aquaporin 1, 3 and 5 in colorectal carcinoma: correlation with clinicopathological characteristics and prognosis. Pathol Oncol Res 2023;29:1611179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171. Tsang KW, Leung JC, Tipoe GL. et al. Down-regulation of aquaporin 3 in bronchiectatic airways in vivo. Respir Med 2003;97:59–64. [DOI] [PubMed] [Google Scholar]
- 172. Jensen HH, Login FH, Park JY. et al. Immunohistochemical evalulation of activated Ras and Rac1 as potential downstream effectors of aquaporin-5 in breast cancer in vivo. Biochem Biophys Res Commun 2017;493:1210–6. [DOI] [PubMed] [Google Scholar]
- 173. Deen PM, van Aubel RA, van Lieburg AF. et al. Urinary content of aquaporin 1 and 2 in nephrogenic diabetes insipidus. J Am Soc Nephrol 1996;7:836–41. [DOI] [PubMed] [Google Scholar]
- 174. Chen Y, Chen H, Wang C. et al. The correlation between the increased expression of aquaporins on the inner limiting membrane and the occurrence of diabetic macular edema. Oxid Med Cell Longev 2022;2022:e7412208. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 175. Musumeci G, Leonardi R, Carnazza ML. et al. Aquaporin 1 (AQP1) expression in experimentally induced osteoarthritic knee menisci: an in vivo and in vitro study. Tissue Cell 2013;45:145–52. [DOI] [PubMed] [Google Scholar]
- 176. Cai L, Lei C, Li R. et al. Overexpression of aquaporin 4 in articular chondrocytes exacerbates the severity of adjuvant-induced arthritis in rats: an in vivo and in vitro study. J Inflamm (Lond) 2017;14:6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177. Tomás-Camardiel M, Venero JL, De Pablos RM. et al. In vivo expression of aquaporin-4 by reactive microglia. J Neurochem 2004;91:891–9. [DOI] [PubMed] [Google Scholar]
- 178. Guttman JA, Samji FN, Li Y. et al. Aquaporins contribute to diarrhoea caused by attaching and effacing bacterial pathogens. Cell Microbiol 2007;9:131–41. [DOI] [PubMed] [Google Scholar]
- 179. Liu JY, Chen XX, Chen HY. et al. Downregulation of aquaporin 9 exacerbates beta-amyloid-induced neurotoxicity in Alzheimer’s disease models in vitro and in vivo. Neuroscience 2018;394:72–82. [DOI] [PubMed] [Google Scholar]
- 180. Yu Y, Wang M, Yu X. et al. Targeting Forkhead box O1-aquaporin 5 axis mitigates neuropathic pain in a CCI rat model through inhibiting astrocytic and microglial activation. Bioengineered 2022;13:8567–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181. Yang LY, Chen YR, Lee JE. et al. Dental pulp stem cell-derived conditioned medium alleviates subarachnoid hemorrhage-induced microcirculation impairment by promoting M2 microglia polarization and reducing astrocyte swelling. Transl Stroke Res 2023;14:688–703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182. Gleiser C, Wagner A, Fallier-Becker P. et al. Aquaporin-4 in astroglial cells in the CNS and supporting cells of sensory organs—A comparative perspective. Int J Mol Sci 2016;17:1411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183. McLennan R, Schumacher LJ, Morrison JA. et al. Neural crest migration is driven by a few trailblazer cells with a unique molecular signature narrowly confined to the invasive front. Development 2015;142:2014–25. [DOI] [PubMed] [Google Scholar]
- 184. Saadoun S, Papadopoulos MC, Watanabe H. et al. Involvement of aquaporin-4 in astroglial cell migration and glial scar formation. J Cell Sci 2005;118:5691–8. [DOI] [PubMed] [Google Scholar]
- 185. Ruiz-Ederra J, Verkman AS.. Aquaporin-1-facilitated keratocyte migration in cell culture and in vivo corneal wound healing models. Exp Eye Res 2009;89:159–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Hara-Chikuma M, Sohara E, Rai T. et al. Progressive adipocyte hypertrophy in aquaporin-7-deficient mice: adipocyte glycerol permeability as a novel regulator of fat accumulation. J Biol Chem 2005;280:15493–6. [DOI] [PubMed] [Google Scholar]
- 187. Yang B, Song Y, Zhao D, Verkman AS.. Phenotype analysis of aquaporin-8 null mice. Am J Physiol Cell Physiol 2005;288:C1161–1170. [DOI] [PubMed] [Google Scholar]
- 188. Ma T, Fukuda N, Song Y. et al. Lung fluid transport in aquaporin-5 knockout mice. J Clin Invest 2000;105:93–100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189. Varadaraj K, Kumari SS, Mathias RT.. Transgenic expression of AQP1 in the fiber cells of AQP0 knockout mouse: effects on lens transparency. Exp Eye Res 2010;91:393–404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190. Rojek A, Füchtbauer EM, Kwon TH. et al. Severe urinary concentrating defect in renal collecting duct-selective AQP2 conditional-knockout mice. Proc Natl Acad Sci USA 2006;103:6037–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191. Hara-Chikuma M, Verkman AS.. Aquaporin-3 facilitates epidermal cell migration and proliferation during wound healing. J Mol Med (Berl) 2008;86:221–31. [DOI] [PubMed] [Google Scholar]
- 192. Rojek AM, Skowronski MT, Füchtbauer EM. et al. Defective glycerol metabolism in aquaporin 9 (AQP9) knockout mice. Proc Natl Acad Sci USA 2007;104:3609–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Okada S, Misaka T, Tanaka Y. et al. Aquaporin-11 knockout mice and polycystic kidney disease animals share a common mechanism of cyst formation. FASEB J 2008;22:3672–84. [DOI] [PubMed] [Google Scholar]
- 194. Ohta E, Itoh T, Nemoto T. et al. Pancreas-specific aquaporin 12 null mice showed increased susceptibility to caerulein-induced acute pancreatitis. Am J Physiol Cell Physiol 2009;297:C1368–1378. [DOI] [PubMed] [Google Scholar]
- 195. Fan Y, Song TR, Wei Q. et al. Modulatory effect of aquaporin 5 on estrogen-induced epithelial-mesenchymal transition in prostate epithelial cells. Chin Med J (Engl) 2020;134:448–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Binder DK, Yao X, Zador Z. et al. Increased seizure duration and slowed potassium kinetics in mice lacking aquaporin-4 water channels. Glia 2006;53:631–6. [DOI] [PubMed] [Google Scholar]
- 197. Ding JH, Sha LL, Chang J. et al. Alterations of striatal neurotransmitter release in aquaporin-4 deficient mice: an in vivo microdialysis study. Neurosci Lett 2007;422:175–80. [DOI] [PubMed] [Google Scholar]
- 198. Auguste KI, Jin S, Uchida K. et al. Greatly impaired migration of implanted aquaporin-4-deficient astroglial cells in mouse brain toward a site of injury. FASEB J 2007;21:108–16. [DOI] [PubMed] [Google Scholar]
- 199. Hara Y, Ando F, Oikawa D. et al. LRBA is essential for urinary concentration and body water homeostasis. Proc Natl Acad Sci USA 2022;119:e2202125119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200. Schnermann J, Chou CL, Ma T. et al. Defective proximal tubular fluid reabsorption in transgenic aquaporin-1 null mice. Proc Natl Acad Sci USA 1998;95:9660–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201. Sindhu Kumari S, Varadaraj K.. Aquaporin 5 knockout mouse lens develops hyperglycemic cataract. Biochem Biophys Res Commun 2013;441:333–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202. Hibuse T, Maeda N, Funahashi T. et al. Aquaporin 7 deficiency is associated with development of obesity through activation of adipose glycerol kinase. Proc Natl Acad Sci USA 2005;102:10993–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203. Varadaraj K, Kumari SS.. Lens aquaporins function as peroxiporins to facilitate membrane transport of hydrogen peroxide. Biochem Biophys Res Commun 2020;524:1025–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204. Kannan A, Mariajoseph-Antony LF, Panneerselvam A. et al. Aquaporin 9 regulates Leydig cell steroidogenesis in diabetes. Syst Biol Reprod Med 2022;68:213–26. [DOI] [PubMed] [Google Scholar]
- 205. Clemens DM, Németh-Cahalan KL, Trinh L. et al. In vivo analysis of aquaporin 0 function in Zebrafish: permeability regulation is required for lens transparency. Invest Ophthalmol Vis Sci 2013;54:5136–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206. Sapkota D, Florian C, Doherty BM. et al. Aqp4 stop codon readthrough facilitates amyloid-β clearance from the brain. Brain 2022;145:2982–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207. Agbani EO, Williams CM, Li Y. et al. Aquaporin-1 regulates platelet procoagulant membrane dynamics and in vivo thrombosis. JCI Insight 2018;3:e99062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208. Takata T, Hamada S, Mae Y. et al. Uromodulin regulates murine aquaporin−2 activity via thick ascending limb–collecting duct cross−talk during water deprivation. Int J Mol Sci 2022;23:9410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Procino G, Barbieri C, Tamma G. et al. AQP2 exocytosis in the renal collecting duct—involvement of SNARE isoforms and the regulatory role of Munc18b. J Cell Sci 2008;121:2097–106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210. Adams ME, Mueller HA, Froehner SC.. In vivo requirement of the α-syntrophin PDZ domain for the sarcolemmal localization of nNOS and aquaporin-4. J Cell Biol 2001;155:113–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211. Szeto IYY, Chu DKH, Chen P. et al. SOX9 and SOX10 control fluid homeostasis in the inner ear for hearing through independent and cooperative mechanisms. Proc Natl Acad Sci USA 2022;119:e2122121119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212. Xue Y, Gursky Z, Monte B. et al. Sustained glymphatic transport and impaired drainage to the nasal cavity observed in multiciliated cell ciliopathies with hydrocephalus. Fluids Barriers CNS 2022;19:20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213. Arellano G, Loda E, Chen Y. et al. Interferon-gamma controls pathogenic T-helper 17 and B cells in a new Aquaporin-4 induced experimental autoimmune encephalomyelitis. J Immunol 2022;208:44.03–03. [Google Scholar]
- 214. Florio M, Engfors A, Gena P. et al. Characterization of the aquaporin-9 inhibitor RG100204 in vitro and in db/db mice. Cells 2022;11:3118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215. He X, Gao F, Hou J. et al. Metformin inhibits MAPK signaling and rescues pancreatic aquaporin 7 expression to induce insulin secretion in type 2 diabetes mellitus. J Biol Chem 2021;297:101002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216. Nomura N, Nunes P, Bouley R. et al. High-throughput chemical screening identifies AG-490 as a stimulator of aquaporin 2 membrane expression and urine concentration. Am J Physiol Cell Physiol 2014;307:C597–C605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Igarashi H, Tsujita M, Suzuki Y. et al. Inhibition of aquaporin-4 significantly increases regional cerebral blood flow. NeuroReport 2013;24:324–8. [DOI] [PubMed] [Google Scholar]
- 218. Klebe S, Griggs K, Cheng Y. et al. Blockade of aquaporin 1 inhibits proliferation, motility, and metastatic potential of mesothelioma In Vitro but not in an In Vivo model. Dis Markers 2015;2015:e286719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219. Höcherl K, Schmidt C, Kurt B. et al. Inhibition of NF-κB ameliorates sepsis-induced downregulation of aquaporin-2/V2 receptor expression and acute renal failure in vivo. Am J Physiol Renal Physiol 2010;298:F196–204. [DOI] [PubMed] [Google Scholar]
- 220. Kinoshita M, Nakatsuji Y, Kimura T. et al. Anti-aquaporin-4 antibody induces astrocytic cytotoxicity in the absence of CNS antigen-specific T cells. Biochem Biophys Res Commun 2010;394:205–10. [DOI] [PubMed] [Google Scholar]
- 221. Chen G, Yao C, Hasegawa T. et al. Effects of isoproterenol on aquaporin 5 levels in the parotid gland of mice in vivo. Am J Physiol Endocrinol Metab 2014;306:E100–E108. doi: [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Delporte C, O'Connell BC, He X. et al. Increased fluid secretion after adenoviral-mediated transfer of the aquaporin-1 cDNA to irradiated rat salivary glands. Proc Natl Acad Sci USA 1997;94:3268–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223. Mukherjee A, Wu D, Davis HC. et al. Non-invasive imaging using reporter genes altering cellular water permeability. Nat Commun 2016;7:13891. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224. Li M, Liu Z, Wu Y. et al. In vivo imaging of astrocytes in the whole brain with engineered AAVs and diffusion-weighted magnetic resonance imaging. Mol Psychiatry 2022;1–8. 10.1038/s41380-022-01580-0. [DOI] [PubMed] [Google Scholar]
- 225. Zheng N, Li M, Wu Y. et al. A novel technology for in vivo detection of cell type-specific neural connection with AQP1-encoding rAAV2-retro vector and metal-free MRI. NeuroImage 2022;258:119402. [DOI] [PubMed] [Google Scholar]
- 226. Urushihata T, Takuwa H, Takahashi M. et al. Exploring cell membrane water exchange in aquaporin-4-deficient ischemic mouse brain using diffusion-weighted MRI. Eur Radiol Exp 2021;5:44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227. Igarashi H, Tsujita M, Kwee IL, Nakada T.. Water influx into cerebrospinal fluid is primarily controlled by aquaporin-4, not by aquaporin-1: 17O JJVCPE MRI study in knockout mice. Neuroreport 2014;25:39–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228. Suzuki Y, Nakamura Y, Yamada K. et al. Aquaporin-4 positron emission tomography imaging of the human brain: first report. J Neuroimaging 2013;23:219–23. [DOI] [PubMed] [Google Scholar]
- 229. Vorontsova I, Vallmitjana A, Torrado B. et al. In vivo macromolecular crowding is differentially modulated by aquaporin 0 in zebrafish lens: insights from a nanoenvironment sensor and spectral imaging. Sci Adv 2022;8:eabj4833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230. Madeira A, de Almeida A, de Graaf C. et al. A gold coordination compound as a chemical probe to unravel aquaporin-7 function. ChemBioChem 2014;15:1487–94. [DOI] [PubMed] [Google Scholar]
- 231. Wacker SJ, Aponte-Santamaría C, Kjellbom P. et al. The identification of novel, high affinity AQP9 inhibitors in an intracellular binding site. Mol Membr Biol 2013;30:246–60. [DOI] [PubMed] [Google Scholar]
- 232. Calvanese L, Pellegrini-Calace M, Oliva R.. In silico study of human aquaporin AQP11 and AQP12 channels: study of Aquaporin AQP11 and AQP12 channels. Protein Sci 2013;22:455–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233. Chen Z, Jiao S, Zhao D. et al. The characterization of structure and prediction for aquaporin in tumour progression by machine learning. Front Cell Dev Biol 2022;10:845622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234. Varadi M, Anyango S, Deshpande M. et al. AlphaFold protein structure database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res 2022;50:D439–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235. Ilan B, Tajkhorshid E, Schulten K. et al. The mechanism of proton exclusion in aquaporin channels. Proteins Struct Funct Bioinforma 2004;55:223–8. [DOI] [PubMed] [Google Scholar]
- 236. De Groot BL, Frigato T, Helms V. et al. The mechanism of proton exclusion in the aquaporin-1 water channel. J Mol Biol 2003;333:279–93. [DOI] [PubMed] [Google Scholar]
- 237. Wei T, Zhou M, Gu L. et al. A novel gating mechanism of aquaporin-4 water channel mediated by blast shockwaves for brain edema. J Phys Chem Lett 2022;13:2486–92. [DOI] [PubMed] [Google Scholar]
- 238. Yu J, Yool AJ, Schulten K, Tajkhorshid E.. Mechanism of gating and ion conductivity of a possible tetrameric pore in aquaporin-1. Structure 2006;14:1411–23. [DOI] [PubMed] [Google Scholar]
- 239. Törnroth-Horsefield S, Wang Y, Hedfalk K. et al. Structural mechanism of plant aquaporin gating. Nature 2006;439:688–94. [DOI] [PubMed] [Google Scholar]
- 240. Bernardi M, Marracino P, Liberti M. et al. Controlling ionic conductivity through transprotein electropores in human aquaporin 4: a non-equilibrium molecular-dynamics study. Phys Chem Chem Phys 2019;21:3339–46. [DOI] [PubMed] [Google Scholar]
- 241. Hub JS, Aponte-Santamaría C, Grubmüller H, De Groot BL.. Voltage-regulated water flux through aquaporin channels in silico. Biophys J 2010;99:L97–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242. Mom R, Muries B, Benoit P. et al. Voltage‐gating of aquaporins, a putative conserved safety mechanism during ionic stresses. FEBS Lett 2021;595:41–57. [DOI] [PubMed] [Google Scholar]
- 243. Stansfeld PJ, Jefferys EE, Sansom MSP.. Multiscale simulations reveal conserved patterns of lipid interactions with aquaporins. Structure 2013;21:810–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244. Zhang YB, Chen LY.. In silico study of Aquaporin V: effects and affinity of the central pore-occluding lipid. Biophys Chem 2013;171:24–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245. Chan R, Falato M, Liang H, Chen LY.. In silico simulations of erythrocyte aquaporins with quantitative in vitro validation. RSC Adv 2020;10:21283–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246. Decker K, Page M, Boyd A. et al. Selective permeability of truncated aquaporin 1 in silico. ACS Biomater Sci Eng 2017;3:342–8. [DOI] [PubMed] [Google Scholar]
- 247. Wang Y, Cohen J, Boron WF. et al. Exploring gas permeability of cellular membranes and membrane channels with molecular dynamics. J Struct Biol 2007;157:534–44. [DOI] [PubMed] [Google Scholar]
- 248. Yusupov M, Yan D, Cordeiro RM, Bogaerts A.. Atomic scale simulation of H2O2 permeation through aquaporin: toward the understanding of plasma cancer treatment. J Phys D Appl Phys 2018;51:125401. [Google Scholar]
- 249. Hub JS, De Groot BL.. Mechanism of selectivity in aquaporins and aquaglyceroporins. Proc Natl Acad Sci USA 2008;105:1198–203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250. Zhu F, Tajkhorshid E, Schulten K.. Collective diffusion model for water permeation through microscopic channels. Phys Rev Lett 2004;93:224501. [DOI] [PubMed] [Google Scholar]
- 251. Hadidi H, Kamali R.. Molecular dynamics study of water transport through AQP5-R188C mutant causing palmoplantar keratoderma (PPK) using the gating mechanism concept. Biophys Chem 2021;277:106655. [DOI] [PubMed] [Google Scholar]
- 252. Pluhackova K, Schittny V, Bürkner P. et al. Multiple pore lining residues modulate water permeability of GlpF. Protein Sci 2022;31:e4431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253. Sachdeva R, Singh B.. Phosphorylation of Ser-180 of rat aquaporin-4 shows marginal affect on regulation of water permeability: molecular dynamics study. J Biomol Struct Dyn 2014;32:555–66. [DOI] [PubMed] [Google Scholar]
- 254. Mom R, Réty S, Mocquet V. et al. Plant aquaporin gating is reversed by phosphorylation on intracellular loop D—evidence from molecular dynamics simulations. Int J Mol Sci 2023;24:13798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255. de Almeida A, Mósca AF, Wragg D. et al. The mechanism of aquaporin inhibition by gold compounds elucidated by biophysical and computational methods. Chem Commun (Camb) 2017;53:3830–3. [DOI] [PubMed] [Google Scholar]
- 256. Mom R, Réty S, Auguin D.. Cortisol interaction with aquaporin-2 modulates its water permeability: perspectives for non-genomic effects of corticosteroids. Int J Mol Sci 2023;24:1499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257. Padhi S, Priyakumar UD.. Selectivity and transport in aquaporins from molecular simulation studies. Vitam Horm 2020;112:47–70. [DOI] [PubMed] [Google Scholar]
- 258. Hub JS, Grubmüller H, De Groot BL.. Dynamics and energetics of permeation through aquaporins. What do we learn from molecular dynamics simulations. Handb Exp Pharmacol 2009;190:57–76. [DOI] [PubMed] [Google Scholar]
- 259. Sutto L, Marsili S, Gervasio FL.. New advances in metadynamics. WIREs Comput Mol Sci 2012;2:771–9. [Google Scholar]
- 260. Wragg D, de Almeida A, Casini A, Leoni S.. Unveiling the mechanisms of aquaglyceroporin‐3 water and glycerol permeation by metadynamics. Chem A Eur J 2019;25:8713–8. [DOI] [PubMed] [Google Scholar]
- 261. Wragg D, Leoni S, Casini A.. Aquaporin-driven hydrogen peroxide transport: a case of molecular mimicry. RSC Chem Biol 2020;1:390–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262. Reale R, English NJ, Garate JA. et al. Human aquaporin 4 gating dynamics under and after nanosecond-scale static and alternating electric-field impulses: a molecular dynamics study of field effects and relaxation. J Chem Phys 2013;139:205101. [DOI] [PubMed] [Google Scholar]
- 263. English NJ, Garate JA.. Near-microsecond human aquaporin 4 gating dynamics in static and alternating external electric fields: non-equilibrium molecular dynamics. J Chem Phys 2016;145:085102. [DOI] [PubMed] [Google Scholar]
- 264. Seeliger D, Zapater C, Krenc D. et al. Discovery of novel human aquaporin-1 blockers. ACS Chem Biol 2013;8:249–56. [DOI] [PubMed] [Google Scholar]
- 265. Yadav DK, Kumar S, Choi EH. et al. Computational modeling on aquaporin-3 as skin cancer target: a virtual screening study. Front Chem 2020;8:250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266. Wenzel MN, Mósca AF, Graziani V. et al. Insights into the mechanisms of aquaporin-3 iby gold(III) complexes: the importance of non-coordinative adduct formation. Inorg Chem 2019;58:2140–8. [DOI] [PubMed] [Google Scholar]
- 267. Huber VJ, Tsujita M, Yamazaki M. et al. Identification of arylsulfonamides as Aquaporin 4 inhibitors. Bioorg Med Chem Lett 2007;17:1270–3. [DOI] [PubMed] [Google Scholar]
- 268. Brown D, Hasler U, Nunes P. et al. Phosphorylation events and the modulation of aquaporin 2 cell surface expression. Curr Opin Nephrol Hypertens 2008;17:491–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269. Canzler S, Fischer M, Ulbricht D. et al. ProteinPrompt : a webserver for predicting protein–protein interactions. Bioinform Adv 2022;2:vbac059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270. McLennan R, McKinney MC, Teddy JM. et al. Neural crest cells bulldoze through the microenvironment using Aquaporin-1 to stabilize filopodia. Development 2020;147:dev.185231. [DOI] [PubMed] [Google Scholar]
- 271. Leberecht C, Schroeder M, Labudde D.. A multiscale model of the regulation of aquaporin 2 recycling. NPJ Syst Biol Appl 2022;8:16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272. Fong S, Chiorini JA, Sneyd J. et al. Computational modeling of epithelial fluid and ion transport in the parotid duct after transfection of human aquaporin-1. Am J Physiol Gastrointest Liver Physiol 2017;312:G153–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273. Papadopoulos MC, Verkman AS.. Aquaporin-4 gene disruption in mice reduces brain swelling and mortality in pneumococcal meningitis. J Biol Chem 2005;280:13906–12. [DOI] [PubMed] [Google Scholar]
- 274. Jin B-J, Rossi A, Verkman AS.. Model of aquaporin-4 supramolecular assembly in orthogonal arrays based on heterotetrameric association of M1-M23 isoforms. Biophys J 2011;100:2936–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
No new data were generated or analysed in support of this research.

