Next Article in Journal
Diagnostic and Prognostic Value of IL-10, FABP2 and LPS Levels in HCC Patients
Previous Article in Journal
Gut Microbiota’s Oxalate-Degrading Activity and Its Implications on Cardiovascular Health in Patients with Kidney Failure: A Pilot Prospective Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring the Multifaceted Potential of Sildenafil in Medicine

by
Ciprian Pușcașu
,
Anca Zanfirescu
*,
Simona Negreș
and
Oana Cristina Șeremet
Faculty of Pharmacy, “Carol Davila” University of Medicine and Pharmacy, Traian Vuia 6, 020956 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Medicina 2023, 59(12), 2190; https://doi.org/10.3390/medicina59122190
Submission received: 30 November 2023 / Revised: 11 December 2023 / Accepted: 14 December 2023 / Published: 17 December 2023
(This article belongs to the Section Pharmacology)

Abstract

:
Phosphodiesterase type 5 (PDE5) is pivotal in cellular signalling, regulating cyclic guanosine monophosphate (cGMP) levels crucial for smooth muscle relaxation and vasodilation. By targeting cGMP for degradation, PDE5 inhibits sustained vasodilation. PDE5 operates in diverse anatomical regions, with its upregulation linked to various pathologies, including cancer and neurodegenerative diseases. Sildenafil, a selective PDE5 inhibitor, is prescribed for erectile dysfunction and pulmonary arterial hypertension. However, considering the extensive roles of PDE5, sildenafil might be useful in other pathologies. This review aims to comprehensively explore sildenafil’s therapeutic potential across medicine, addressing a gap in the current literature. Recognising sildenafil’s broader potential may unveil new treatment avenues, optimising existing approaches and broadening its clinical application.

1. Introduction

Sildenafil is the first oral medication approved by the United States Food and Drug Administration (FDA) for the therapeutic management of erectile dysfunction (ED) [1,2].
Originally developed with the intention of treating hypertension and angina pectoris, it showed no efficacy in phase 1 clinical studies, including patients with these pathologies. However, it induced a distinct pharmacological response in these individuals, inducing significant penile erection [1,3]. This accidental discovery led to the patenting of sildenafil by Pfizer in 1996, and, two years later, sildenafil received approval for the treatment of ED. Subsequently, several other uses of this pharmaceutical agent have been discovered [4].
Sildenafil is a highly effective and specific inhibitor of phosphodiesterase type 5 (PDE5) [5,6]. There are several types of phosphodiesterases, but only three (PDE5, PDE6, and PDE9) specifically hydrolyse cyclic guanosine monophosphate (cGMP) over cyclic adenosine monophosphate (cAMP) [7,8]. It hydrolyses the phosphodiesterase linkage and facilitates the catalytic conversion of cGMP to inactive 5′-guanosine monophosphate (GMP), hence regulating several physiological processes within the body [9], such as neuroprotection, antinociception, synaptic plasticity, calcium homeostasis, and vasodilation (Figure 1) [10]. cGMP is involved in these processes through the activation of ion channels or cGMP-dependent protein kinases or through its interaction with PDE [9]. Nitric oxide (NO) promotes cGMP production in the target cell. NO and cGMP are integral constituents of a signalling transduction cascade that operates through autocrine, paracrine, and potentially endocrine mechanisms. The NO/cGMP/PDE5 axis has been implicated in various illnesses, including neurological disorders, pulmonary arterial hypertension, cardiomyopathy, cancer, ED, and lower urinary tract syndrome [8,11,12,13,14].
Additionally, sildenafil might possess other mechanisms of action and, consequently, supplementary therapeutic effects. Thus, sildenafil shows promising results in the treatment of neurodegenerative diseases through the activation of peroxisome proliferator-activated receptor-γ coactivator 1α (PGC1α), following the accumulation of cGMP, which increases mitochondrial biogenesis [15,16], upregulates antioxidant enzymes [17], and decreases β-site amyloid precursor protein-cleaving enzyme 1 (BACE1) expression [18].
Sildenafil also exhibits anti-inflammatory and neuroprotective properties, potentially mediated through the regulation of the AMP-activated protein kinase (AMPK)/nuclear factor of kappa light polypeptide gene enhancer in the B-cell inhibitor alpha (IKβα)/nuclear factor kappa B (NFκB) signalling pathway [19]. Furthermore, endothelial NO synthase (eNOS) contributes to the neuroprotective mechanism of sildenafil by promoting the activation of AMPK [19,20]. The potential of PDE5 inhibitors to treat chronic pain is supported by their ability to counteract the downregulation of angiopoietin 1 expression, a crucial factor involved in vascular stability and neurite development, within cultured dorsal root ganglion neurons [21,22].
Sildenafil might enhance the susceptibility of various types of tumour cells to the cytotoxic impact of chemotherapeutic agents [23]. This enhancement is achieved through the promotion of apoptosis, which is mediated by the downregulation of B-cell lymphoma-extralarge (Bcl-xL) and Fas-associated phosphatase-1 (FAP-1) expression, increased generation of reactive oxygen species (ROS), and upregulation of caspase-3, 8, and 9 activities [23,24,25].
The primary aim of this narrative review is to comprehensively investigate and synthesise the diverse therapeutic potentials of sildenafil within the field of medicine, as no such review currently exists. Understanding the broader potential of sildenafil can have implications for clinical practice. It may lead to the identification of new treatment options or the optimisation of existing ones, potentially expanding the range of conditions for which the drug could be considered. In summary, a review of the multifaceted potential of sildenafil in medicine serves as a valuable tool for consolidating existing knowledge, identifying new opportunities for research and clinical application, and potentially improving patient care.

2. Materials and Method

PUBMED was used to search the literature for the most relevant articles containing in vitro and in vivo preclinical and clinical findings on the effect of sildenafil in various conditions other than ED and PAH. We limited the search to articles published in English between 2015 and 2023. We did, however, investigate previous relevant research. We used the following keywords and MeSH terms: “sildenafil” AND “pain”, “sildenafil” AND “Alzheimer” OR “ Alzheimer disease”, “sildenafil” AND “ Raynaud’s” OR “Raynaud’s phenomenon”, “sildenafil” AND “digital ulcer”, “sildenafil” AND “wound healing”, “sildenafil” AND “cancer”, “sildenafil” AND “depression”, “sildenafil” AND “retinopathy”, “sildenafil” AND “renal” OR “renal disease” OR “kidney” OR “kidney disease” OR “nephropathy”, “sildenafil” AND “gastrointestinal” OR “gastrointestinal disease”, “sildenafil” AND “cardiovascular” OR “cardiovascular disease”, “sildenafil” AND “lung” OR “lung disease”. We selected the most appropriate studies after analysis and cross-checking. A literature study was also conducted to present the important characteristics of sildenafil pharmacokinetics, side effects, and current indications. We used the following keywords and MeSH terms: “sildenafil” AND “pharmacokinetics” OR “side effects” OR “indications” OR “erectile disfunction” OR “ED” OR “pulmonary arterial hypertension” OR “PAH”.

3. Sildenafil Pharmacokinetic Profile and Associated Adverse Reactions

Sildenafil citrate is rapidly absorbed from the gastrointestinal tract, reaching maximum plasma concentration approximately 30–120 min (with a median of 60 min) after oral administration. The absorption rate of sildenafil citrate is reduced when taken with food (particularly fatty foods) [26,27,28,29,30]. Along with its primary circulating N-desmethyl metabolite, sildenafil citrate exhibits a binding affinity of around 96% to plasma proteins [26,27,31]. It is mainly metabolised through the hepatic microsomal isoenzyme cytochrome P (CYP)3A4, with a slight contribution from the hepatic isoenzymes CYP2C9. Sildenafil citrate has 16 known metabolites. The primary metabolite, N-desmethyl sildenafil, exhibits a PDE selectivity profile comparable to that of sildenafil and an in vitro PDE5 potency roughly half that of the parent medication. The N-desmethyl metabolite is metabolised further, having a terminal half-life of about 4 h. Sildenafil citrate is a mild inhibitor of drug-metabolising enzymes, with CYP2C9 being the most inhibited [28,30,32]. Sildenafil citrate is mainly eliminated as metabolites in the faeces, accounting for approximately 80% of the orally taken dose. A smaller proportion is excreted in the urine, representing approximately 13% of the orally supplied dose. Its half-life ranges from 3 to 5 h [26,27,31].
The adverse effects of sildenafil are typically mild-to-moderate in severity and of a brief duration. Among the most frequent adverse reactions associated with the use of sildenafil are headaches, flushing, dyspepsia, and visual disturbances [31]. Mechanistically, these effects can be attributed to the vasodilatory properties of sildenafil. This vasodilation, although therapeutically beneficial, may cause headaches and flushing due to changes in blood flow. Additionally, altered retinal sensitivity to light may contribute to visual disturbances. Gastrointestinal effects, such as dyspepsia, may be linked to the presence of PDE5 in the smooth muscle of the gastrointestinal tract. Understanding these underlying mechanisms is crucial for both clinicians and researchers to optimise the therapeutic benefits of sildenafil while mitigating its associated adverse reactions [26,27,31]. Continued research in this area will undoubtedly refine our comprehension of the intricate pharmacological profile of sildenafil, further enhancing its clinical utility [26,27]. Sildenafil’s systemic vasodilation may result in a transient decrease in blood pressure, particularly when co-administered with nitrate-based medications. This hemodynamic effect underscores the importance of cautious prescribing, especially in patients with a history of myocardial infarction or unstable angina. It is crucial for healthcare providers to conduct a thorough cardiovascular assessment before prescribing sildenafil, ensuring a balanced consideration of its benefits against potential risks in individuals with underlying cardiovascular issues. Regular monitoring and personalised treatment plans remain integral in minimising the cardiovascular risks associated with sildenafil use [26,27].

4. Current Indications

Sildenafil is commonly utilised as a PDE5 inhibitor for two main indications.

4.1. Erectile Dysfunction (ED)

ED is a prevalent medical condition observed in males, and its frequency and occurrence rise with age. It is associated with diminished overall health and the coexistence of other medical conditions [33]. ED refers to the incapacity to attain or sustain an erection that is considered adequate for engaging in sexual activity [34]. The global occurrence of ED ranges from 4 to 66 cases per 1000 males per year [33,35,36,37,38]. In Europe, according to the European Male Ageing Study, ED prevalence ranges from 6% to 64% across various age groups. The prevalence of ED was found to increase with age, with an average prevalence of 30% [39]. Furthermore, the prevalence of ED appears to be higher in the United States, as well as in Eastern and Southeastern Asian countries, when compared to Europe or South America [40].
The NO–cGMP pathway serves as the principal mechanism for penile erection in primates, including humans. Sexual arousal activates neurological pathways, triggering the release of NO directly into the corpus cavernosum of the penis from both neurons and endothelial cells [11,41]. The biological stimulus responsible for initiating penile erection is cGMP, which then activates protein kinase G (PKG), responsible for the reduction of intracellular calcium levels. This reduction leads to the relaxation of arterial and trabecular smooth muscle, which subsequently causes arterial dilatation, venous constriction, and the attainment of penile erection stiffness [41,42]. PDE5 is the predominant PDE enzyme found in the corpus cavernosum, which facilitates the degradation of cGMP through the process of hydrolysis [9]. A reduction in NO levels may lead to ED [38].
The efficacy of sildenafil and other PDE5 inhibitors is widely recognised as the most successful approach for managing ED due to its high efficacy and remarkable lack of substantial adverse effects [43,44].

4.2. Pulmonary Arterial Hypertension (PAH)

PAH is characterised by elevated blood pressure in the pulmonary arteries. The walls of the pulmonary arteries become narrowed, thickened, or stiff, leading to increased resistance to blood flow. This elevated resistance eventually leads to right heart failure if left untreated. The Sixth World Symposium on PAH (2022) established a mean pulmonary artery pressure (mPAP) threshold exceeding 20 mmHg as the criterion for defining PAH [45]. The worldwide prevalence of PAH is significant, affecting around 1% of the global population [46].
There are various strategies for treating endothelial dysfunction. They usually aim to enhance the bioavailability of NO or the amounts of cGMP within cells. Subsequently, PDE5 inhibitors are frequently used for the management of PAH [42], being approved for its treatment in adults and children starting at the age of 1 year [8]. For PDE5 inhibitors to exhibit efficacy, it is imperative that there be an adequate level of activity within the NO/soluble guanylate cyclase (sGC)/cGMP pathway. In cases where NO is inadequately produced, an alternative approach should be used, such as sGC stimulators that are not dependent on NO [42].
The FDA specifically designates sildenafil for the therapeutic management of group 1 PAH [31], a chronic and progressive disorder characterised by the development of vascular abnormalities within the pulmonary circulation [46]. The addition of sildenafil to epoprostenol therapy can lead to a delay in clinical worsening in patients with New York Heart Association (NYHA) Functional Class II–III symptoms and idiopathic aetiology in 71% of cases and connective tissue disease in 25% of cases [47,48].
On the other hand, the European Medicines Agency (EMA) recommends the use of sildenafil for the treatment of adult patients diagnosed with PAH who are classed as WHO functional class II (slight limitation of physical activity) and III (marked limitation of physical activity) to enhance their exercise capacity [31]. Furthermore, EMA also designates sildenafil as a therapeutic option for the management of PAH in paediatric patients ranging from 1 to 17 years of age [31,49,50].

5. Prospective Indications

Drug repurposing continues to be a subject of great interest within the pharmaceutical and healthcare sectors, allowing the discovery of novel applications for medications licensed for other original indications [51]. Once a medicine was discovered to have an off-target or a newly recognised on-target effect, it was pushed forward for commercial exploitation [52].
As mentioned above, sildenafil is currently approved for the treatment of ED and PAH. Given the varying distribution of PDE5 across organs, recent research has extensively explored the potential therapeutic effect of sildenafil in various other conditions, such as pain, cancer, Alzheimer’s disease, depression, Raynaud’s phenomenon, digital ulcers, wound healing, retinopathy, gastric ulcer, colitis, nephropathy, lung injuries, and ischaemia, providing novel potential clinical applications for this molecule.

5.1. Pain

Local administration of L-arginine mitigates carrageenan-induced hyperalgesia, an effect prevented by NO inhibitors. Thus, cGMP was thought to be involved in the process of antinociception [53,54]. Subsequently, the analgesic effect of sildenafil was investigated using various animal models and dosages, showing a dose-dependent analgesic effect against diabetic and traumatic neuropathic pain, thermal hyperalgesia, and formalin-induced pain (Table 1). Intravenously administered sildenafil raises cGMP levels and, subsequently, modulates potassium channel function, promoting gamma-aminobutyric acid (GABA) release and resulting in an antinociceptive action. Thus, GABA receptors might mediate the pharmacological activity of sildenafil [55]. Additionally, PDE5 inhibitors seem to reduce chronic pain, mitigating the decrease in angiopoietin 1 expression, a critical regulator of blood vessel stability and nerve fibre growth in cultured dorsal root ganglion neurons [21,22,56].

5.2. Alzheimer Disease

Various sildenafil concentrations elicit the activation of PGC1α, triggering mitochondrial biogenesis [15,16], enhancing the production of antioxidant enzymes [17], and reducing BACE1 expression [18]. Thus, sildenafil might provide substantial advantages for individuals afflicted with Alzheimer’s disease. Furthermore, its vasodilatory effect might provide an additional benefit for individuals with Alzheimer’s disease since they exhibit cerebral hypoperfusion [63].
Sildenafil inhibits apoptosis in neurons experiencing hypoxia and facilitates neurogenesis [64,65,66,67,68]. Thus, it might decelerate the degeneration of neurons associated with Alzheimer’s disease and facilitate neurogenesis. Additionally, sildenafil enhances insulin sensitivity and reduces endothelial inflammation in individuals with diabetes. Therefore, it could potentially exert similar effects on insulin sensitivity and inflammation in the context of Alzheimer’s disease [69,70].
The in vitro administration of sildenafil protected brain mitochondria against β amyloid (Aβ) and advanced glycation end product (AGEPs)-induced injuries. In vivo, the administration of sildenafil induced an upregulation of brain-derived neurotrophic factor, a reduction in reactive astrocytes and microglia, a decrease in proinflammatory cytokines and neuronal apoptosis, and an increase in neurogenesis. Furthermore, in individuals with Alzheimer’s disease, sildenafil reduced aberrant spontaneous neural activity, increased cerebral blood flow, and enhanced the cerebral metabolic rate of oxygen (Table 2) [71,72].

5.3. Systemic Sclerosis-Associated Raynaud’s Disease and Digital Ulcer

The pathophysiology of systemic sclerosis-induced vasculopathy includes reduced levels of NO. Subsequently, topical preparations that result in NO release induce a notable vasodilation in patients with this pathology [89,90]. The utilisation of a specific inhibitor targeting PDE5, such as sildenafil, is also an effective therapeutic approach for managing vascular disease in individuals with systemic sclerosis, considering that they have the ability to enhance the levels of NO and sustain peripheral blood circulation [91,92]. Clinical trials demonstrated that administration of sildenafil positively impacts the healing process of ischaemic digital ulcers in individuals diagnosed with systemic sclerosis, while in patients with Raynaud’s phenomenon, it decreases the severity, frequency, and duration of the disease (Table 3) [72,93,94].

5.4. Wound Healing

The treatment of chronic wounds continues to be a challenge in the medical field [107,108]. Tissue injury initiates a series of reparative processes aimed at restoring the structural integrity and functionality of the affected tissue [109], including enhancement of the clotting process, mitigation of oxidative stress, formation of new blood vessels, endothelial cell growth, and tissue remodelling [92]. NO seems to regulate some of these actions, leading to its recognition and utilisation as a biomarker for wound healing [110]. Moreover, deficiencies in inducible nitric oxide synthase (iNOS) and endothelial nitric oxide synthase (eNOS) have been linked to impaired wound healing [111,112]. Subsequently, sildenafil has been extensively studied in the field of wound healing, particularly in relation to its role in preserving the viability of skin flaps (Table 4).

5.5. Retinopathy

Other than its effects on PDE5, sildenafil also inhibits, to a smaller extent (about 10% of the activity of PDE5), PDE6, which is exclusively found in photoreceptor cells, being involved in phototransduction. Hence, it is plausible that the ocular manifestations associated with sildenafil use result from the suppression of PDE-6 activity [124,125].
PDE5 is found in the blood arteries of the choroid and retina [126]. Subsequently, sildenafil induces NO-induced vasodilation, particularly of the choroidal blood vessels, enhancing eye blood flow. This effect is probably achieved by the vasodilation process mediated by NO [127]. Furthermore, NO inhibits retinal vascular obliteration and the consequent development of proliferative retinopathies [128].
Animal studies demonstrated that sildenafil also reduced retinal vascular obliteration and neovascularisation. Additionally, it reduced the concentration of vascular endothelial growth factor (VEGF) in the ocular tissues, inhibited the progression of ischaemia injury, and reduced the thickness of the outer plexiform layer (Table 5). Thus, PDE5 inhibitors have the potential to impact disorders characterised by augmented choroidal thickness or choroidal and/or retinal ischaemia [129].

5.6. Cancer

PDE5 expression is elevated in various types of human carcinomas, such as urinary bladder tumours, metastatic breast cancers, and non-small cell lung cancers [133,134,135], possibly being involved in tumourigenesis. Hence, the suppression of PDE5 activity might result in antineoplastic properties [136].
Cyclic nucleotide second messengers, namely cGMP and cAMP, exhibit substrate affinity in the low micromolar range for ATP-binding cassette sub-family C member 4/multidrug resistance-associated protein 4 (ABCC4/MRP4) and ATP-binding cassette sub-family C member 5/multidrug resistance-associated protein 5 (ABCC5/MRP5). Sildenafil effectively suppresses the efflux function of ABCC4 and ABCC5 transporters [137,138]. Shi et al. demonstrated that sildenafil has inhibitory effects on ABC (ATP-binding cassette) transporters, specifically ABCB1 (ATP-binding cassette sub-family B) and ABCG2 (ATP-binding cassette sub-family G), and reverses multidrug resistance in cancer cells [139].
Numerous preclinical studies have documented the utilisation of sildenafil in conjunction with chemotherapeutic drugs for the management of diverse types of cancer (Table 6). Sildenafil has exhibited its efficacy in augmenting the delivery of anticancer drugs by leveraging the enhanced permeation retention (EPR) effect, leading to a substantial increase in drug concentrations within tumours and consequent induction of cellular apoptosis [23]. Furthermore, sildenafil has been shown to have a role in the regulation and enhancement of chemotherapeutic drugs in several cancer types. This phenomenon has been demonstrated in numerous in vitro and in vivo studies, wherein the downregulation of Bcl-xL expression; the promotion of ROS generation; the phosphorylation of BCL2-associated agonist of cell death (BAD) and B-cell lymphoma protein 2 (Bcl-2); the upregulation of caspase-3, 8, and 9 activities; the arrest of cells at the G0/G1 cell cycle phase; the circumvention of cancer cell resistance by inhibiting various ABC transporters through the elevation of cGMP; and the augmentation of autophagosome and autolysosome levels collectively induce apoptotic cell death in tumour cells [23,25,140].
In addition, clinical trials have been conducted to investigate the efficacy of sildenafil in the treatment of various forms of cancer (Table 6). The predominant focus of clinical research has been on the utilisation of sildenafil as a chemoadjuvant to improve its pharmacokinetic properties and to mitigate the adverse effects commonly associated with chemotherapy, specifically ED and cardiotoxicity.

5.7. Depression

Despite being primarily intended for peripheral tissue activity, sildenafil has the ability to traverse the blood–brain barrier (BBB) and impact central PDE5 activity, exhibiting various effects within the central nervous system, including the promotion of neurogenesis, augmentation of memory, mitigation of learning impairment, and neuroprotection [154]. Unsurprisingly, the utilisation of sildenafil to target PDE5 has emerged as a novel therapeutic approach in the management of various neurological and neuropsychiatric disorders [44], including depression. There is a strong correlation between depression and ED, with each condition mutually influencing and exacerbating the other. Consequently, failure may occur if one disease is treated while neglecting the other [155]. Additionally, ED can occur as an adverse effect of some antidepressant drugs [156]. Sexual dysfunction is also a prevalent symptom of major depressive disorder itself and may be linked to decreased levels of testosterone [157,158,159].
Clinical and preclinical studies have demonstrated that PDE5 inhibitors are effective not just in the treatment of ED, but also in ameliorating ED-associated depression [160]. Preclinical studies indicate an antidepressant effect similar to that of clinically used antidepressants. This action is believed to be mediated by the activation of oxytocin signalling pathways. Additionally, sildenafil increases the central concentrations of serotonin and noradrenaline [161,162] (Table 7).

5.8. Renal Diseases

Cyclic nucleotide signal transduction pathways are a burgeoning area of study in the field of kidney disease, with ongoing emphasis on the targeted inhibition of PDE5 [167]. Modulation of the cGMP-dependent protein kinase 1-phosphodiesterase (cGMP-cGK1-PDE) signalling pathway may be renoprotective and could lead to the development of innovative therapeutic approaches for chronic kidney disease. Subsequently, PDE5 inhibitors have been recognised as a prospective therapeutic alternative for various forms of renal injuries [168].
Multiple mechanisms have been postulated to play a role in counteracting the cascade of alterations generated by renal damage. The most extensively documented mechanisms by which PDE5 provides protection include stimulation of NO production via NOS activation, improvement of medullary blood flow, reversal of the Bcl2/Bax ratio, phosphorylation of extracellular signal-regulated kinase (ERK), activation of mitochondrial biogenesis, upregulation of renal progenitor cells, and regulation of multiple signalling pathways, including insulin/insulin-like growth factor 1 (IGF1) and nuclear factor NF-kB [169].
An elevation in ERK phosphorylation stimulates the activity of NOS, leading to an immediate release of NO [170]. In the days following renal injury, energy is required for the repair process, which is supplied by the cellular mitochondria. Despite the fact that mitochondria undergo continuous regeneration, cellular damage such as sepsis and hypoxia stimulates rapid biogenesis, and PGC-1α facilitates this process. PGC-1α induces the activation of nuclear respiratory factors 1 and 2, which subsequently stimulate mitochondrial transcription factor A, activating DNA (deoxyribonucleic acid) transcription within the mitochondria [16,171].
However, the protective effect of PDE5 operates via an alternative pathway. PDE5 increases cGMP, which stimulates PKG, which, in turn, opens mitochondrial KATP channels (ATP-sensitive potassium channel), causing Mg2+ release and depolarisation of the mitochondrial inner membrane. A depolarised membrane leads to a decrease in Ca2+ influx, which subsequently inhibits cellular death. An elevated concentration of Mg2+ decreases ROS and p38 mitogen-activated protein kinase (MAPK) activation, both of which are implicated in apoptosis [170,172,173].
Preclinical studies indicate that the administration of sildenafil improved renal function, decreasing the levels of oxidative stress indicators, such as malondialdehyde (MDA) and pro-inflammatory cytokines [174,175]. Furthermore, clinical studies further support this effect, that the administration of sildenafil decreases serum creatinine levels and increases the glomerular filtration rate. Additionally, sildenafil considerably reduces albuminuria, as well as haemoglobin A1c (HbA1c) levels (Table 8) [176,177].

5.9. Gastrointestinal Diseases

Gastric ulcers arise due to the harmful effects of acid and pepsin [206]. The stomach typically maintains a delicate equilibrium involving various components (such as defensive agents, the mucus bicarbonate barrier, surface epithelial cells, mucosal regeneration, blood circulation, acid–base balance, and epidermal growth factor) to counteract the effects of acid and pepsin present in the luminal content [207]. The equilibrium is typically disrupted due to Helicobacter pylori infection or prolonged use of nonsteroidal anti-inflammatory medicines (NSAIDs) [208]. NO exerts a substantial impact on the generation of mucosal resistance and repair [209], effectively safeguarding against stomach mucosal injury [210]. Consequently, sildenafil facilitates ulcer repair.
Inflammatory bowel diseases, such as ulcerative colitis, are chronic conditions of unknown origin characterised by significant inflammation of the digestive tract [211]. Their pathogenesis comprises a complex interplay among genetic, immunological, and environmental variables. During the acute episodes of inflammation, there is an increase in inflammatory cells at the intestinal level, leading to excessive synthesis of proinflammatory mediators, including cytokines, eicosanoids, ROS, and nitrogen metabolites [212,213,214]. Sildenafil exerts anti-inflammatory effects by preventing lipid peroxidation, cytokine release, and oxidant generation [215,216].
Preclinical studies examined the therapeutic potential of sildenafil against various gastrointestinal diseases (Table 9), demonstrating that it effectively prevents ulceration by increasing the concentration of NO within the gastric tissue. Additionally, sildenafil enhances levels of antioxidant enzymes, reduces gastric acid secretion, decreases levels of proinflammatory cytokines, and increases the level of glutathione (GSH) [72,216,217].

5.10. Cardiovascular Diseases

Extensive studies have been conducted on the possible application of sildenafil in the field of cardiovascular disease. Seven of the eleven major PDE subtypes—PDE1, 2, 3, 4, 5, 8, and 9—are expressed in the heart, and their diverse effects on cAMP and cGMP signalling in distinct cell types, including cardiomyocytes, provide therapeutic prospects to combat heart disease [226]. PDE5 is abundantly expressed in isolated canine or murine ventricular cardiomyocytes, and it is strongly expressed in both experimental and human heart disease. PDE5 inhibition should have the potential to alter heart function, contributing to the treatment of a range of cardiovascular illnesses [42].
The vasodilatory effects of PDE5 inhibitors have positive effects on vascular coagulopathy and enhance endothelial functioning in various regions of the body [227]. The presence of PDE5 antibodies has been detected in coronary vascular smooth muscle cells, it seems that healthy myocardium does not exhibit significant expression of this enzyme [8]. Nonetheless, in cases of congestive heart failure and right ventricular hypertrophy, the activation of angiotensin II in vascular smooth muscle cells results in an increase in PDE5 levels, thereby leading to a decrease in cGMP/PKG reactivity [228].
A strong correlation has been established between ED and endothelial dysfunction, indicating that ED might serve as an external indicator for various underlying cardiovascular diseases. Sildenafil enhances endothelial function in medical disorders such as diabetes and congestive heart failure and enhances perfusion and oxygen tension within the vasculature holding potential for addressing ischaemia/reperfusion injury, including myocardial infarction and stroke [1,229,230,231].
Preclinical and clinical trials demonstrate the cardioprotective effects associated with the inhibition of PDE5 in individuals diagnosed with established cardiovascular disease. In animal models, sildenafil enhanced both systolic and diastolic cardiac function, reduced cardiac hypertrophy and apoptosis of cardiomyocytes, and lowered the susceptibility to post-ischaemic arrhythmia [72,232,233]. In individuals diagnosed with heart failure, the administration of sildenafil decreased pulmonary artery pressure and alleviated dyspnoea. Additionally, it improved brachial artery flow-mediated dilatation and boosted breathing during physical exertion [234]. Moreover, in patients diagnosed with chronic stable left ventricular failure, sildenafil has demonstrated improvements in left ventricular ejection fraction and performance on the 6 min walking test (Table 10) [235].

5.11. Lung Diseases

While sildenafil has received approval for the treatment of PAH, studies have demonstrated its positive effects on various respiratory conditions, such as bronchopulmonary dysplasia and cystic fibrosis, while other conditions, such as asthma, had a negative impact. Furthermore, research conducted on animals has demonstrated that the administration of sildenafil in the treatment of PAH is associated with a decrease in several inflammatory mediators and the regulation of several intracellular signalling molecules, including MAPK, NF-kβ, and ERK [267]. Preclinical and clinical research indicates that sildenafil could potentially serve as a novel therapeutic approach for lung disorders, particularly those characterised by an inflammatory reaction (Table 11).

6. Summary

Our narrative review focused on summarising the potential clinical applications of sildenafil. Sildenafil, recognised for its vasodilatory effects, has shown potential therapeutic applications beyond ED and PAH, extending into various medical domains. In the realm of pain management, studies suggest that sildenafil may modulate nociceptive pathways. Its ability to increase pain reaction latency in animal models with neuropathic pain, such as those induced by diabetes, indicates a potential role in alleviating chronic pain [22,58]. The underlying mechanism involves the activation of the cGMP pathway, which is implicated in pain signal processing. Sildenafil’s influence on this pathway may contribute to the attenuation of pain responses, providing a foundation for exploring its utility in clinical settings addressing chronic pain conditions [53,54].
Emerging evidence also points to the neuroprotective effects of sildenafil in neurodegenerative diseases such as Alzheimer’s. The drug’s impact on cerebral blood flow and its ability to reduce oxidative stress and inflammation in the brain suggest potential benefits in mitigating neurodegenerative processes. The modulation of cGMP-dependent pathways and enhancement of neurotrophic factors are proposed mechanisms through which sildenafil may exert neuroprotective effects. This opens avenues for research into the use of sildenafil as a potential adjunctive therapy for neurodegenerative disorders [69,77,88].
In the context of systemic sclerosis-associated Raynaud’s disease and digital ulcers, sildenafil’s vasodilatory properties become particularly relevant. By inhibiting PDE5, the enzyme responsible for cGMP degradation, sildenafil prolongs the vasodilatory effects of NO. This mechanism improves blood flow, offering potential relief for patients with systemic sclerosis where vascular complications are prominent [91,92]. The positive impact on digital ulcers, a common complication in systemic sclerosis, suggests a role for sildenafil in managing the vascular manifestations of autoimmune diseases. The precise mechanisms involve enhanced vasodilation and improved microcirculation, showcasing the multifaceted therapeutic potential of sildenafil in systemic sclerosis-associated complications [72,93,94].
In the realm of cancer, studies propose that sildenafil’s inhibition of PDE5 could impact tumour growth [136]. Modulation of the cGMP signalling pathway influences angiogenesis and immune responses, enhancing the efficacy of chemotherapy and presenting a potential avenue for adjunctive cancer therapy [23]. In depression, sildenafil’s neuroprotective effects and its influence on neurotransmitter systems, including serotonin and dopamine, have been explored [154,161]. By enhancing cGMP signalling in the brain, sildenafil may exhibit antidepressant effects, paving the way for further investigations into its role in mood disorders.
Renal diseases, marked by impaired blood flow and inflammation, may benefit from sildenafil’s vasodilatory and anti-inflammatory properties. The drug’s ability to enhance renal blood flow through the relaxation of vascular smooth muscle cells offers a potential therapeutic avenue for conditions such as chronic kidney disease [169,202]. Gastrointestinal diseases, particularly those involving impaired blood perfusion, could also be targeted by sildenafil. The drug’s vasodilatory effects may aid in improving blood flow to the gastrointestinal tract, potentially mitigating complications in conditions such as inflammatory bowel disease [215,216]. Additionally, sildenafil’s impact on ischaemia/reperfusion injury holds promise in cardiovascular events such as myocardial infarction and stroke. The vasodilation induced by sildenafil can enhance blood flow during reperfusion, potentially reducing tissue damage and inflammation [1,229,230,231,244]. These multifaceted applications of sildenafil underscore its potential as a versatile therapeutic agent beyond its well-established roles, necessitating further research to unravel the full extent of its clinical utility.
While preclinical studies have demonstrated promising outcomes, including mechanisms of action and potential benefits, several limitations impede a straightforward transition to clinical applications. Preclinical trials often involve animal models, which may not fully mirror the complexities of human physiology and pathology. Additionally, variations in dosages, administration methods, and experimental designs among preclinical studies contribute to a lack of standardised protocols. Furthermore, the diverse nature of human diseases requires careful consideration of factors such as patient variability, comorbidities, and long-term effects, which may not be fully captured in preclinical settings.
Supplementary limitations in the current body of research include a potential bias towards positive outcomes in published studies, which may not reflect the entire spectrum of findings and could skew the overall perception of sildenafil’s efficacy. Additionally, there is a need for more comprehensive investigations into potential adverse effects and interactions with other medications, especially in the context of chronic usage. Addressing these limitations is crucial for establishing the safety and effectiveness of sildenafil in diverse clinical scenarios.
Consequently, the current gap between preclinical evidence and clinical trials underscores the need for rigorous, well-designed clinical studies to validate the safety, efficacy, and optimal dosages of sildenafil in these diverse medical indications. Bridging this gap is crucial to ascertain the true therapeutic potential of sildenafil and to establish evidence-based guidelines for its application in clinical settings.

7. Conclusions

In conclusion, our comprehensive review highlights the diverse and promising therapeutic potential of sildenafil beyond its well-established uses. From its role in pain management, neurodegenerative diseases, and systemic sclerosis-associated complications to its impact on cancer, depression, and various organ-specific diseases, sildenafil emerges as a multifaceted therapeutic agent. However, rigorous, well-designed clinical studies are essential to guide evidence-based guidelines for the nuanced application of sildenafil in diverse clinical settings, ensuring its potential as a versatile therapeutic agent is realised to its fullest extent.

Author Contributions

Conceptualisation, S.N.; data curation, C.P.; writing—original draft preparation, C.P.; writing—review and editing, A.Z. and O.C.Ș. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data generated or analysed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jackson, G.; Gillies, H.; Osterloh, I. Past, present, and future: A 7-year update of Viagra® (sildenafil citrate). Int. J. Clin. Pract. 2005, 59, 680–691. [Google Scholar] [CrossRef] [PubMed]
  2. Uthayathas, S.; Karuppagounder, S.; Thrash, B.M.; Parameshwaran, K.; Suppiramaniam, V.; Dhanasekaran, M. Versatile effects of sildenafil: Recent pharmacological applications. Pharmacol. Rep. 2007, 59, 150–163. [Google Scholar]
  3. Kadhim, T.J.; Khalf, O.A.A. A review search of sildenafil uses in human and in the veterinary medicine. AIP Conf. Proc. 2023, 2475, 100010. [Google Scholar] [CrossRef]
  4. Simonca, L.; Tulloh, R. Sildenafil in Infants and Children. Children 2017, 4, 60. [Google Scholar] [CrossRef] [PubMed]
  5. Ghofrani, H.A.; Osterloh, I.H.; Grimminger, F. Sildenafil: From angina to erectile dysfunction to pulmonary hypertension and beyond. Nat. Rev. Drug Discov. 2006, 5, 689–702. [Google Scholar] [CrossRef] [PubMed]
  6. Gong, B.; Ma, M.; Xie, W.; Yang, X.; Huang, Y.; Sun, T.; Luo, Y.; Huang, J. Direct comparison of tadalafil with sildenafil for the treatment of erectile dysfunction: A systematic review and meta-analysis. Int. Urol. Nephrol. 2017, 49, 1731–1740. [Google Scholar] [CrossRef] [PubMed]
  7. Delhaye, S.; Bardoni, B. Role of phosphodiesterases in the pathophysiology of neurodevelopmental disorders. Mol. Psychiatry 2021, 26, 4570–4582. [Google Scholar] [CrossRef]
  8. Saikia, Q.; Hazarika, A.; Mishra, R. A Review on the Pharmacological Importance of PDE5 and Its Inhibition to Manage Biomedical Conditions. J. Pharmacol. Pharmacother. 2022, 13, 246–257. [Google Scholar] [CrossRef]
  9. Nandi, S.; Kumar, P.; Amin, S.; Jha, T.; Gayen, S. First molecular modelling report on tri-substituted pyrazolines as phosphodiesterase 5 (PDE5) inhibitors through classical and machine learning based multi-QSAR analysis. SAR QSAR Environ. Res. 2021, 32, 917–939. [Google Scholar] [CrossRef]
  10. Pasmanter, N.; Iheanacho, F.; Hashmi, M.F. Biochemistry, Cyclic GMP. 2023. Available online: https://pubmed.ncbi.nlm.nih.gov/31194391 (accessed on 12 November 2023).
  11. Corbin, J.D. Mechanisms of action of PDE5 inhibition in erectile dysfunction. Int. J. Impot. Res. 2004, 16, S4–S7. [Google Scholar] [CrossRef]
  12. Francis, S.H.; Blount, M.A.; Corbin, J.D. Mammalian Cyclic Nucleotide Phosphodiesterases: Molecular Mechanisms and Physiological Functions. Physiol. Rev. 2011, 91, 651–690. [Google Scholar] [CrossRef] [PubMed]
  13. Ückert, S.; Kuczyk, M.A. Cyclic Nucleotide Metabolism Including Nitric Oxide and Phosphodiesterase-Related Targets in the Lower Urinary Tract. In Urinary Tract; Springer: Berlin/Heidelberg, Germany, 2011; pp. 527–542. [Google Scholar] [CrossRef]
  14. Nelson, M.D.; Rader, F.; Tang, X.; Tavyev, J.; Nelson, S.F.; Miceli, M.C.; Elashoff, R.M.; Sweeney, H.L.; Victor, R.G. PDE5 inhibition alleviates functional muscle ischemia in boys with Duchenne muscular dystrophy. Neurology 2014, 82, 2085–2091. [Google Scholar] [CrossRef] [PubMed]
  15. Fernandez-Marcos, P.J.; Auwerx, J. Regulation of PGC-1α, a nodal regulator of mitochondrial biogenesis. Am. J. Clin. Nutr. 2011, 93, 884S–890S. [Google Scholar] [CrossRef] [PubMed]
  16. Whitaker, R.M.; Wills, L.P.; Stallons, L.J.; Schnellmann, R.G. cGMP-Selective Phosphodiesterase Inhibitors Stimulate Mitochondrial Biogenesis and Promote Recovery from Acute Kidney Injury. J. Pharmacol. Exp. Ther. 2013, 347, 626–634. [Google Scholar] [CrossRef]
  17. Olmos, Y.; Sanchez-Gomez, F.J.; Wild, B.; Garcia-Quintans, N.; Cabezudo, S.; Lamas, S.; Monsalve, M. SirT1 Regulation of Antioxidant Genes Is Dependent on the Formation of a FoxO3a/PGC-1α Complex. Antioxid. Redox Signal. 2013, 19, 1507–1521. [Google Scholar] [CrossRef]
  18. Kwak, Y.-D.; Wang, R.; Li, J.J.; Zhang, Y.-W.; Xu, H.; Liao, F.-F. Differential regulation of BACE1 expression by oxidative and nitrosative signals. Mol. Neurodegener. 2011, 6, 17. [Google Scholar] [CrossRef]
  19. Nunes, A.K.S.; Rapôso, C.; Rocha, S.W.S.; Barbosa, K.P.D.S.; Luna, R.L.D.A.; da Cruz-Höfling, M.A.; Peixoto, C.A. Involvement of AMPK, IKβα-NFκB and eNOS in the sildenafil anti-inflammatory mechanism in a demyelination model. Brain Res. 2015, 1627, 119–133. [Google Scholar] [CrossRef]
  20. Zhang, J.; Xie, Z.; Dong, Y.; Wang, S.; Liu, C.; Zou, M.-H. Identification of Nitric Oxide as an Endogenous Activator of the AMP-activated Protein Kinase in Vascular Endothelial Cells. J. Biol. Chem. 2008, 283, 27452–27461. [Google Scholar] [CrossRef]
  21. Wang, L.; Chopp, M.; Zhang, Z.G. PDE5 inhibitors promote recovery of peripheral neuropathy in diabetic mice. Neural Regen. Res. 2017, 12, 218–219. [Google Scholar] [CrossRef]
  22. Pușcașu, C.; Ungurianu, A.; Șeremet, O.C.; Andrei, C.; Mihai, D.P.; Negreș, S. The Influence of Sildenafil–Metformin Combination on Hyperalgesia and Biochemical Markers in Diabetic Neuropathy in Mice. Medicina 2023, 59, 1375. [Google Scholar] [CrossRef]
  23. Haider, M.; Elsherbeny, A.; Pittalà, V.; Fallica, A.N.; Alghamdi, M.A.; Greish, K. The Potential Role of Sildenafil in Cancer Management through EPR Augmentation. J. Pers. Med. 2021, 11, 585. [Google Scholar] [CrossRef] [PubMed]
  24. Urla, C.; Stagno, M.J.; Fuchs, J.; Warmann, S.W.; Schmid, E. Combination therapy of doxorubicin and Sildenafil inhibits the growth of pediatric rhabdomyosarcoma. J. Cancer Res. Clin. Oncol. 2022, 149, 2513–2522. [Google Scholar] [CrossRef] [PubMed]
  25. Das, A.; Durrant, D.; Mitchell, C.; Mayton, E.; Hoke, N.N.; Salloum, F.N.; Park, M.A.; Qureshi, I.; Lee, R.; Dent, P.; et al. Sildenafil increases chemotherapeutic efficacy of doxorubicin in prostate cancer and ameliorates cardiac dysfunction. Proc. Natl. Acad. Sci. USA 2010, 107, 18202–18207. [Google Scholar] [CrossRef] [PubMed]
  26. FDA. Label: VIAGRA (Sildenafil Citrate) Tablets. Available online: www.fda.gov/medwatch (accessed on 7 September 2023).
  27. Sildenafil 25 mg Film-Coated Tablets—Summary of Product Characteristics (SmPC)—(emc). Available online: https://www.medicines.org.uk/emc/product/3172/smpc (accessed on 7 September 2023).
  28. Chaumais, M.-C.; Perrin, S.; Sitbon, O.; Simonneau, G.; Humbert, M.; Montani, D. Pharmacokinetic evaluation of sildenafil as a pulmonary hypertension treatment. Expert Opin. Drug Metab. Toxicol. 2013, 9, 1193–1205. [Google Scholar] [CrossRef] [PubMed]
  29. Rezvanfar, M.A.; Rahimi, H.R.; Abdollahi, M. ADMET considerations for phosphodiesterase-5 inhibitors. Expert Opin. Drug Metab. Toxicol. 2012, 8, 1231–1245. [Google Scholar] [CrossRef]
  30. Wright, P.J. Comparison of phosphodiesterase type 5 (PDE5) inhibitors. Int. J. Clin. Pract. 2006, 60, 967–975. [Google Scholar] [CrossRef]
  31. Revatio, P. (Sildenafil Citrate) Product Monograph. Available online: https://pdf.hres.ca/dpd_pm/00056673.PDF (accessed on 23 October 2023).
  32. Hyland, R.; Roe, E.G.H.; Jones, B.C.; Smith, D.A. Identification of the cytochrome P450 enzymes involved in the N-demethylation of sildenafil. Br. J. Clin. Pharmacol. 2001, 51, 239–248. [Google Scholar] [CrossRef]
  33. Corona, G.; Cucinotta, D.; Di Lorenzo, G.; Ferlin, A.; Giagulli, V.A.; Gnessi, L.; Isidori, A.M.; Maiorino, M.I.; Miserendino, P.; Murrone, A.; et al. The Italian Society of Andrology and Sexual Medicine (SIAMS), along with ten other Italian Scientific Societies, guidelines on the diagnosis and management of erectile dysfunction. J. Endocrinol. Investig. 2023, 46, 1241–1274. [Google Scholar] [CrossRef]
  34. Virag, R.; Zwang, G.; Dermange, H.; Legman, M. Vasculogenic Impotence: A Review of 92 Cases with 54 Surgical Operations. Vasc. Surg. 1981, 15, 9–17. [Google Scholar] [CrossRef]
  35. Gades, N.M.; Jacobson, D.J.; McGree, M.E.; Sauver, J.L.S.; Lieber, M.M.; Nehra, A.; Girman, C.J.; Jacobsen, S.J. Longitudinal Evaluation of Sexual Function in a Male Cohort: The Olmsted County Study of Urinary Symptoms and Health Status among Men. J. Sex. Med. 2009, 6, 2455–2466. [Google Scholar] [CrossRef]
  36. McCabe, M.P.; Sharlip, I.D.; Lewis, R.; Atalla, E.; Balon, R.; Fisher, A.D.; Laumann, E.; Lee, S.W.; Segraves, R.T. Incidence and Prevalence of Sexual Dysfunction in Women and Men: A Consensus Statement from the Fourth International Consultation on Sexual Medicine 2015. J. Sex. Med. 2016, 13, 144–152. [Google Scholar] [CrossRef] [PubMed]
  37. Martin, S.A.; Atlantis, E.; Lange, K.; Taylor, A.W.; O’Loughlin, P.; Wittert, G.A. Predictors of Sexual Dysfunction Incidence and Remission in Men. J. Sex. Med. 2014, 11, 1136–1147. [Google Scholar] [CrossRef] [PubMed]
  38. Johannes, C.B.; Araujo, A.B.; Feldman, H.A.; Derby, C.A.; Kleinman, K.P.; McKinlay, J.B. Incidence of erectile dysfunction in men 40 to 69 years old: Longitudinal results from the Massachusetts male aging study. J. Urol. 2000, 163, 460–463. [Google Scholar] [CrossRef] [PubMed]
  39. Corona, G.; Lee, D.M.; Forti, G.; O’connor, D.B.; Maggi, M.; O’neill, T.W.; Pendleton, N.; Bartfai, G.; Boonen, S.; Casanueva, F.F.; et al. Age-Related Changes in General and Sexual Health in Middle-Aged and Older Men: Results from the European Male Ageing Study (EMAS). J. Sex. Med. 2010, 7, 1362–1380. [Google Scholar] [CrossRef] [PubMed]
  40. Yafi, F.A.; Jenkins, L.; Albersen, M.; Corona, G.; Isidori, A.M.; Goldfarb, S.; Maggi, M.; Nelson, C.J.; Parish, S.; Salonia, A.; et al. Erectile dysfunction. Nat. Rev. Dis. Prim. 2016, 2, 16003. [Google Scholar] [CrossRef] [PubMed]
  41. Ojo, O.A.; Ojo, A.B.; Maimako, R.F.; Rotimi, D.; Iyobhebhe, M.; Alejolowo, O.O.; Nwonuma, C.O.; Elebiyo, T.C. Exploring the potentials of some compounds from Garcinia kola seeds towards identification of novel PDE-5 inhibitors in erectile dysfunction therapy. Andrologia 2021, 53, e14092. [Google Scholar] [CrossRef] [PubMed]
  42. Andersson, K. PDE5 inhibitors—Pharmacology and clinical applications 20 years after sildenafil discovery. Br. J. Pharmacol. 2018, 175, 2554–2565. [Google Scholar] [CrossRef]
  43. Argiolas, A.; Argiolas, F.M.; Argiolas, G.; Melis, M.R. Erectile Dysfunction: Treatments, Advances and New Therapeutic Strategies. Brain Sci. 2023, 13, 802. [Google Scholar] [CrossRef]
  44. Boyce, E. Sildenafil citrate: A therapeutic update. Clin. Ther. 2001, 23, 2–23. [Google Scholar] [CrossRef]
  45. Humbert, M.; Kovacs, G.; Hoeper, M.M.; Badagliacca, R.; Berger, R.M.; Brida, M.; Carlsen, J.; Coats, A.J.; Escribano-Subias, P.; Ferrari, P.; et al. 2022 ESC/ERS Guidelines for the diagnosis and treatment of pulmonary hypertension. Eur. Respir. J. 2022, 61, 2200879. [Google Scholar] [CrossRef]
  46. Cullivan, S.; Gaine, S.; Sitbon, O. New trends in pulmonary hypertension. Eur. Respir. Rev. 2023, 32, 220211. [Google Scholar] [CrossRef] [PubMed]
  47. Sastry, B.; Narasimhan, C.; Reddy, N.; Raju, B. Clinical efficacy of sildenafil in primary pulmonary hypertension: A randomized, placebo-controlled, double-blind, crossover study. J. Am. Coll. Cardiol. 2004, 43, 1149–1153. [Google Scholar] [CrossRef] [PubMed]
  48. Simonneau, G.; Rubin, L.J.; Galiè, N.; Barst, R.J.; Fleming, T.R.; Frost, A.; Engel, P.; Kramer, M.R.; Serdarevic-Pehar, M.; Layton, G.R.; et al. Long-term sildenafil added to intravenous epoprostenol in patients with pulmonary arterial hypertension. J. Heart Lung Transplant. 2014, 33, 689–697. [Google Scholar] [CrossRef] [PubMed]
  49. Barst, R.J.; Ivy, D.D.; Gaitan, G.; Szatmari, A.; Rudzinski, A.; Garcia, A.E.; Sastry, B.; Pulido, T.; Layton, G.R.; Serdarevic-Pehar, M.; et al. A Randomized, Double-Blind, Placebo-Controlled, Dose-Ranging Study of Oral Sildenafil Citrate in Treatment-Naive Children with Pulmonary Arterial Hypertension. Circulation 2012, 125, 324–334. [Google Scholar] [CrossRef] [PubMed]
  50. Barst, R.J.; Beghetti, M.; Pulido, T.; Layton, G.; Konourina, I.; Zhang, M.; Ivy, D.D. STARTS-2. Circulation 2014, 129, 1914–1923. [Google Scholar] [CrossRef] [PubMed]
  51. Ashburn, T.T.; Thor, K.B. Drug repositioning: Identifying and developing new uses for existing drugs. Nat. Rev. Drug Discov. 2004, 3, 673–683. [Google Scholar] [CrossRef] [PubMed]
  52. Pushpakom, S.; Iorio, F.; Eyers, P.A.; Escott, K.J.; Hopper, S.; Wells, A.; Doig, A.; Guilliams, T.; Latimer, J.; McNamee, C.; et al. Drug repurposing: Progress, challenges and recommendations. Nat. Rev. Drug Discov. 2019, 18, 41–58. [Google Scholar] [CrossRef]
  53. Mixcoatl-Zecuatl, T.; Aguirre-Bañuelos, P.; Granados-Soto, V. Sildenafil produces antinociception and increases morphine antinociception in the formalin test. Eur. J. Pharmacol. 2000, 400, 81–87. [Google Scholar] [CrossRef]
  54. Ferreira, S.H.; Nakamura, M. I—Prostaglandin hyperalgesia, a cAMP/Ca2+ dependent process. Prostaglandins 1979, 18, 179–190. [Google Scholar] [CrossRef]
  55. Huang, L.J.; Yoon, M.H.; Choi, J.I.; Kim, W.M.; Lee, H.G.; Kim, Y.O. Effect of Sildenafil on Neuropathic Pain and Hemodynamics in Rats. Yonsei Med. J. 2010, 51, 82–87. [Google Scholar] [CrossRef]
  56. Swiecicka, A. The efficacy of PDE5 inhibitors in diabetic patients. Andrology 2022, 11, 245–256. [Google Scholar] [CrossRef]
  57. Wang, L.; Chopp, M.; Szalad, A.; Jia, L.; Lu, X.; Lu, M.; Zhang, L.; Zhang, Y.; Zhang, R.; Zhang, Z.G. Sildenafil Ameliorates Long Term Peripheral Neuropathy in Type II Diabetic Mice. PLoS ONE 2015, 10, e0118134. [Google Scholar] [CrossRef]
  58. Pușcașu, C. Investigation of antihyperalgesic effects of different doses of sildenafil and metformin in alloxan-induced diabetic neuropathy in mice. Farmacia 2023, 71, 598–611. [Google Scholar] [CrossRef]
  59. Ambriz-Tututi, M.; Velázquez-Zamora, D.A.; Urquiza-Marín, H.; Granados-Soto, V. Analysis of the mechanism underlying the peripheral antinociceptive action of sildenafil in the formalin test. Eur. J. Pharmacol. 2005, 512, 121–127. [Google Scholar] [CrossRef]
  60. Jain, N.K.; Patil, C.; Singh, A.; Kulkarni, S.K. Sildenafil-induced peripheral analgesia and activation of the nitric oxide–cyclic GMP pathway. Brain Res. 2001, 909, 170–178. [Google Scholar] [CrossRef]
  61. Patil, C.S.; Singh, V.P.; Singh, S.; Kulkarni, S.K. Modulatory Effect of the PDE-5 Inhibitor Sildenafil in Diabetic Neuropathy. Pharmacology 2004, 72, 190–195. [Google Scholar] [CrossRef]
  62. Schäfer, A.; Fraccarollo, D.; Pförtsch, S.; Flierl, U.; Vogt, C.; Pfrang, J.; Kobsar, A.; Renné, T.; Eigenthaler, M.; Ertl, G.; et al. Improvement of vascular function by acute and chronic treatment with the PDE-5 inhibitor sildenafil in experimental diabetes mellitus. Br. J. Pharmacol. 2008, 153, 886–893. [Google Scholar] [CrossRef]
  63. Domek-Łopacińska, K.U.; Strosznajder, J.B. Cyclic GMP and Nitric Oxide Synthase in Aging and Alzheimer’s Disease. Mol. Neurobiol. 2010, 41, 129–137. [Google Scholar] [CrossRef]
  64. Caretti, A.; Bianciardi, P.; Ronchi, R.; Fantacci, M.; Guazzi, M.; Samaja, M. Phosphodiesterase-5 Inhibition Abolishes Neuron Apoptosis Induced by Chronic Hypoxia Independently of Hypoxia-Inducible Factor-1α Signaling. Exp. Biol. Med. 2008, 233, 1222–1230. [Google Scholar] [CrossRef]
  65. Wang, L.; Zhang, Z.G.; Zhang, R.L.; Chopp, M. Activation of the PI3-K/Akt Pathway Mediates cGMP Enhanced-Neurogenesis in the Adult Progenitor Cells Derived from the Subventricular Zone. J. Cereb. Blood Flow Metab. 2005, 25, 1150–1158. [Google Scholar] [CrossRef]
  66. Ferreira, E.D.F.; Romanini, C.V.; Cypriano, P.E.; de Oliveira, R.M.W.; Milani, H. Sildenafil provides sustained neuroprotection in the absence of learning recovery following the 4-vessel occlusion/internal carotid artery model of chronic cerebral hypoperfusion in middle-aged rats. Brain Res. Bull. 2013, 90, 58–65. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, R.; Wang, Y.; Zhang, L.; Zhang, Z.; Tsang, W.; Lu, M.; Zhang, L.; Chopp, M. Sildenafil (Viagra) Induces Neurogenesis and Promotes Functional Recovery After Stroke in Rats. Stroke 2002, 33, 2675–2680. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, R.L.; Chopp, M.; Roberts, C.; Wei, M.; Wang, X.; Liu, X.; Lu, M.; Zhang, Z.G. Sildenafil Enhances Neurogenesis and Oligodendrogenesis in Ischemic Brain of Middle-Aged Mouse. PLoS ONE 2012, 7, e48141. [Google Scholar] [CrossRef] [PubMed]
  69. Zemel, M.B.; Kolterman, O.; Rinella, M.; Vuppalanchi, R.; Flores, O.; Barritt, A.S.; Siddiqui, M.; Chalasani, N. Randomized Controlled Trial of a Leucine-Metformin-Sildenafil Combination (NS-0200) on Weight and Metabolic Parameters. Obesity 2018, 27, 59–67. [Google Scholar] [CrossRef] [PubMed]
  70. Aversa, A.; Vitale, C.; Volterrani, M.; Fabbri, A.; Spera, G.; Fini, M.; Rosano, G.M.C. Chronic administration of Sildenafil improves markers of endothelial function in men with Type 2 diabetes. Diabet. Med. 2008, 25, 37–44. [Google Scholar] [CrossRef] [PubMed]
  71. Sanders, O. Sildenafil for the Treatment of Alzheimer’s Disease: A Systematic Review. J. Alzheimer’s Dis. Rep. 2020, 4, 91–106. [Google Scholar] [CrossRef] [PubMed]
  72. Ala, M.; Jafari, R.M.; Dehpour, A.R. Sildenafil beyond erectile dysfunction and pulmonary arterial hypertension: Thinking about new indications. Fundam. Clin. Pharmacol. 2020, 35, 235–259. [Google Scholar] [CrossRef]
  73. Son, Y.; Kim, K.; Cho, H.-R. Sildenafil protects neuronal cells from mitochondrial toxicity induced by β-amyloid peptide via ATP-sensitive K+ channels. Biochem. Biophys. Res. Commun. 2018, 500, 504–510. [Google Scholar] [CrossRef]
  74. Sung, S.K.; Woo, J.S.; Kim, Y.H.; Son, D.W.; Lee, S.W.; Song, G.S. Sildenafil Ameliorates Advanced Glycation End Products-Induced Mitochondrial Dysfunction in HT-22 Hippocampal Neuronal Cells. J. Korean Neurosurg. Soc. 2016, 59, 259–268. [Google Scholar] [CrossRef]
  75. Puzzo, D.; Staniszewski, A.; Deng, S.X.; Privitera, L.; Leznik, E.; Liu, S.; Zhang, H.; Feng, Y.; Palmeri, A.; Landry, D.W.; et al. Phosphodiesterase 5 Inhibition Improves Synaptic Function, Memory, and Amyloid- Load in an Alzheimer’s Disease Mouse Model. J. Neurosci. 2009, 29, 8075–8086. [Google Scholar] [CrossRef]
  76. Zhang, J.; Guo, J.; Zhao, X.; Chen, Z.; Wang, G.; Liu, A.; Wang, Q.; Zhou, W.; Xu, Y.; Wang, C. Phosphodiesterase-5 inhibitor sildenafil prevents neuroinflammation, lowers beta-amyloid levels and improves cognitive performance in APP/PS1 transgenic mice. Behav. Brain Res. 2013, 250, 230–237. [Google Scholar] [CrossRef] [PubMed]
  77. Zhu, L.; Yang, J.-Y.; Xue, X.; Dong, Y.-X.; Liu, Y.; Miao, F.-R.; Wang, Y.-F.; Xue, H.; Wu, C.-F. A novel phosphodiesterase-5 Inhibitor: Yonkenafil modulates neurogenesis, gliosis to improve cognitive function and ameliorates amyloid burden in an APP/PS1 transgenic mice model. Mech. Ageing Dev. 2015, 150, 34–45. [Google Scholar] [CrossRef] [PubMed]
  78. Jin, F.; Gong, Q.-H.; Xu, Y.-S.; Wang, L.-N.; Jin, H.; Li, F.; Li, L.-S.; Ma, Y.-M.; Shi, J.-S. Icariin, a phoshphodiesterase-5 inhibitor, improves learning and memory in APP/PS1 transgenic mice by stimulation of NO/cGMP signalling. Int. J. Neuropsychopharmacol. 2014, 17, 871–881. [Google Scholar] [CrossRef] [PubMed]
  79. Devan, B.D.; Sierramercadojr, D.; Jimenez, M.; Bowker, J.L.; Duffy, K.B.; Spangler, E.L.; Ingram, D.K. Phosphodiesterase inhibition by sildenafil citrate attenuates the learning impairment induced by blockade of cholinergic muscarinic receptors in rats. Pharmacol. Biochem. Behav. 2004, 79, 691–699. [Google Scholar] [CrossRef] [PubMed]
  80. Cuadrado-Tejedor, M.; Hervias, I.; Ricobaraza, A.; Puerta, E.; Pérez-Roldán, J.; García-Barroso, C.; Franco, R.; Aguirre, N.; García-Osta, A. Sildenafil restores cognitive function without affecting β-amyloid burden in a mouse model of Alzheimer’s disease. Br. J. Pharmacol. 2011, 164, 2029–2041. [Google Scholar] [CrossRef]
  81. García-Barroso, C.; Ricobaraza, A.; Pascual-Lucas, M.; Unceta, N.; Rico, A.J.; Goicolea, M.A.; Sallés, J.; Lanciego, J.L.; Oyarzabal, J.; Franco, R.; et al. Tadalafil crosses the blood–brain barrier and reverses cognitive dysfunction in a mouse model of AD. Neuropharmacology 2013, 64, 114–123. [Google Scholar] [CrossRef]
  82. Puzzo, D.; Loreto, C.; Giunta, S.; Musumeci, G.; Frasca, G.; Podda, M.V.; Arancio, O.; Palmeri, A. Effect of phosphodiesterase-5 inhibition on apoptosis and beta amyloid load in aged mice. Neurobiol. Aging 2014, 35, 520–531. [Google Scholar] [CrossRef]
  83. Devan, B.D.; Pistell, P.J.; Duffy, K.B.; Kelley-Bell, B.; Spangler, E.L.; Ingram, D.K. Phosphodiesterase inhibition facilitates cognitive restoration in rodent models of age-related memory decline. NeuroRehabilitation 2014, 34, 101–111. [Google Scholar] [CrossRef]
  84. Orejana, L.; Barros-Miñones, L.; Jordan, J.; Cedazo-Minguez, A.; Tordera, R.M.; Aguirre, N.; Puerta, E. Sildenafil Decreases BACE1 and Cathepsin B Levels and Reduces APP Amyloidogenic Processing in the SAMP8 Mouse. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2014, 70, 675–685. [Google Scholar] [CrossRef]
  85. Orejana, L.; Barros-Miñones, L.; Aguirre, N.; Puerta, E. Implication of JNK pathway on tau pathology and cognitive decline in a senescence-accelerated mouse model. Exp. Gerontol. 2013, 48, 565–571. [Google Scholar] [CrossRef]
  86. Ibrahim, M.; Haleem, M.; AbdelWahab, S.; Abdel-Aziz, A. Sildenafil ameliorates Alzheimer disease via the modulation of vascular endothelial growth factor and vascular cell adhesion molecule-1 in rats. Hum. Exp. Toxicol. 2020, 40, 596–607. [Google Scholar] [CrossRef] [PubMed]
  87. Samudra, N.; Motes, M.; Lu, H.; Sheng, M.; Diaz-Arrastia, R.; Devous, M.; Hart, J.; Womack, K.B. A Pilot Study of Changes in Medial Temporal Lobe Fractional Amplitude of Low Frequency Fluctuations after Sildenafil Administration in Patients with Alzheimer’s Disease. J. Alzheimer’s Dis. 2019, 70, 163–170. [Google Scholar] [CrossRef]
  88. Sheng, M.; Lu, H.; Liu, P.; Li, Y.; Ravi, H.; Peng, S.-L.; Diaz-Arrastia, R.; Devous, M.D.; Womack, K.B. Sildenafil Improves Vascular and Metabolic Function in Patients with Alzheimer’s Disease. J. Alzheimer’s Dis. 2017, 60, 1351–1364. [Google Scholar] [CrossRef] [PubMed]
  89. Tucker, A.; Pearson, R.; Cooke, E.; Benjamin, N. Effect of nitric-oxide-generating system on microcirculatory blood flow in skin of patients with severe Raynaud’s syndrome: A randomised trial. Lancet 1999, 354, 1670–1675. [Google Scholar] [CrossRef]
  90. Anderson, M.E.; Moore, T.L.; Hollis, S.; Jayson, M.I.V.; King, T.A.; Herrick, A.L.; Anderson, M.E.; Moore, T.L.; Hollis, S.; Jayson, M.I.V.; et al. Digital vascular response to topical glyceryl trinitrate, as measured by laser Doppler imaging, in primary Raynaud’s phenomenon and systemic sclerosis. Rheumatology 2002, 41, 324–328. [Google Scholar] [CrossRef] [PubMed]
  91. Placzek, M.; Degitz, K.; Schmidt, H.; Plewig, G. Acne fulminans in late-onset congenital adrenal hyperplasia. Lancet 1999, 354, 739–740. [Google Scholar] [CrossRef] [PubMed]
  92. Farsaie, S.; Khalili, H.; Karimzadeh, I.; Dashti-Khavidaki, S. An Old Drug for a New Application: Potential Benefits of Sildenafil in Wound Healing. J. Pharm. Pharm. Sci. 2012, 15, 483–498. [Google Scholar] [CrossRef]
  93. Fries, R.; Shariat, K.; von Wilmowsky, H.; Böhm, M. Sildenafil in the Treatment of Raynaud’s Phenomenon Resistant to Vasodilatory Therapy. Circulation 2005, 112, 2980–2985. [Google Scholar] [CrossRef]
  94. Herrick, A.L.; Hoogen, F.V.D.; Gabrielli, A.; Tamimi, N.; Reid, C.; O’Connell, D.; Vázquez-Abad, M.-D.; Denton, C.P. Modified-release sildenafil reduces Raynaud’s phenomenon attack frequency in limited cutaneous systemic sclerosis. Arthritis Rheum. 2011, 63, 775–782. [Google Scholar] [CrossRef]
  95. Brueckner, C.S.; Becker, M.O.; Kroencke, T.; Huscher, D.; Scherer, H.U.; Worm, M.; Burmester, G.; Riemekasten, G. Effect of sildenafil on digital ulcers in systemic sclerosis: Analysis from a single centre pilot study. Ann. Rheum. Dis. 2009, 69, 1475–1478. [Google Scholar] [CrossRef]
  96. Rademacher, J.-G.; Wincup, C.; Tampe, B.; Korsten, P. Combination therapy with bosentan and sildenafil for refractory digital ulcers and Raynaud’s phenomenon in a 30-year-old woman with systemic sclerosis: Case report and literature review. J. Scleroderma Relat. Disord. 2019, 5, 159–164. [Google Scholar] [CrossRef] [PubMed]
  97. Gore, J.; Silver, R. Oral sildenafil for the treatment of Raynaud’s phenomenon and digital ulcers secondary to systemic sclerosis. Ann. Rheum. Dis. 2005, 64, 1387. [Google Scholar] [CrossRef] [PubMed]
  98. Hachulla, E.; Hatron, P.-Y.; Carpentier, P.; Agard, C.; Chatelus, E.; Jego, P.; Mouthon, L.; Queyrel, V.; Fauchais, A.-L.; Michon-Pasturel, U.; et al. Efficacy of sildenafil on ischaemic digital ulcer healing in systemic sclerosis: The placebo-controlled SEDUCE study. Ann. Rheum. Dis. 2015, 75, 1009–1015. [Google Scholar] [CrossRef] [PubMed]
  99. Bellando-Randone, S.; Lepri, G.; Bruni, C.; Blagojevic, J.; Radicati, A.; Cometi, L.; De Paulis, A.; Matucci-Cerinic, M.; Guiducci, S. Combination therapy with Bosentan and Sildenafil improves Raynaud’s phenomenon and fosters the recovery of microvascular involvement in systemic sclerosis. Clin. Rheumatol. 2015, 35, 127–132. [Google Scholar] [CrossRef] [PubMed]
  100. Kamata, Y.; Kamimura, T.; Iwamoto, M.; Minota, S. Comparable effects of sildenafil citrate and alprostadil on severe Raynaud’s phenomenon in a patient with systemic sclerosis. Clin. Exp. Dermatol. 2005, 30, 451. [Google Scholar] [CrossRef] [PubMed]
  101. Rosenkranz, S.; Diet, F.; Karasch, T.; Weihrauch, J.; Wassermann, K.; Erdmann, E. Sildenafil improved pulmonary hypertension and peripheral blood flow in a patient with scleroderma-associated lung fibrosis and the raynaud phenomenon. Ann. Intern. Med. 2003, 139, 871–873. [Google Scholar] [CrossRef] [PubMed]
  102. Roustit, M.; Hellmann, M.; Cracowski, C.; Blaise, S.; Cracowski, J.L. Sildenafil increases digital skin blood flow during all phases of local cooling in primary Raynaud’s phenomenon. Clin. Pharmacol. Ther. 2012, 91, 813–819. [Google Scholar] [CrossRef]
  103. Rosenkranz, S.; Caglayan, E.; Diet, F.; Karasch, T.; Weihrauch, J.; Wassermann, K.; Erdmann, E. Langzeiteffekte von Sildenafil bei Sklerodermie-assoziierter pulmonaler Hypertonie und Raynaud-Syndrom. Dtsch. Med. Wochenschr. 2004, 129, 1736–1740. [Google Scholar] [CrossRef]
  104. Cordova, L.Z.; Schrale, R.; Castley, A. Sildenafil Therapy in Patients with Digital Burns and Raynaud’s Syndrome. J. Burn. Care Res. 2018, 40, 136–139. [Google Scholar] [CrossRef]
  105. Wortsman, X.; Del Barrio-Díaz, P.; Meza-Romero, R.; Poehls-Risco, C.; Vera-Kellet, C. Nifedipine cream versus sildenafil cream for patients with secondary Raynaud phenomenon: A randomized, double-blind, controlled pilot study. J. Am. Acad. Dermatol. 2018, 78, 189–190. [Google Scholar] [CrossRef]
  106. Andrigueti, F.V.; Ebbing, P.C.C.; Arismendi, M.I.; Kayser, C. Evaluation of the effect of sildenafil on the microvascular blood flow in patients with systemic sclerosis: A randomised, double-blind, placebo-controlled study. Clin. Exp. Rheumatol. 2017, 35 (Suppl. S1), 151–158. Available online: http://www.ncbi.nlm.nih.gov/pubmed/28281457 (accessed on 23 October 2023). [PubMed]
  107. McDermott-Scales, L.; Cowman, S.; Gethin, G. Prevalence of wounds in a community care setting in Ireland. J. Wound Care 2009, 18, 405–417. [Google Scholar] [CrossRef] [PubMed]
  108. Bours, G.J.; Halfens, R.J.; Abu-Saad, H.H.; Grol, R.T. Prevalence, prevention, and treatment of pressure ulcers: Descriptive study in 89 institutions in The Netherlands. Res. Nurs. Health 2002, 25, 99–110. [Google Scholar] [CrossRef] [PubMed]
  109. Taş, A.; Atasoy, N.; Özbek, H.; Aslan, L.; Yüksel, H.; Ceylan, E.; Dagoglu, G. Effects of Sildenafil Citrate (Viagra) in the Early Phase of Healing Process in Open Wounds Induced Experimentally in Dogs. Acta Vet. Brno 2003, 72, 273–277. [Google Scholar] [CrossRef]
  110. Weller, R. Nitric oxide donors and the skin: Useful therapeutic agents? Clin. Sci. 2003, 105, 533–535. [Google Scholar] [CrossRef] [PubMed]
  111. Yamasaki, K.; Edington, H.D.; McClosky, C.; Tzeng, E.; Lizonova, A.; Kovesdi, I.; Steed, D.L.; Billiar, T.R. Reversal of impaired wound repair in iNOS-deficient mice by topical adenoviral-mediated iNOS gene transfer. J. Clin. Investig. 1998, 101, 967–971. [Google Scholar] [CrossRef]
  112. Lee, P.C.; Salyapongse, A.N.; Bragdon, G.A.; Shears, L.L.; Watkins, S.C.; Edington, H.D.J.; Billiar, T.R. Impaired wound healing and angiogenesis in eNOS-deficient mice. Am. J. Physiol. Circ. Physiol. 1999, 277, H1600–H1608. [Google Scholar] [CrossRef]
  113. Kaya, B.; Çerkez, C.; Işılgan, S.E.; Göktürk, H.; Yığman, Z.; Serel, S.; Can, B.; Ergün, H. Comparison of the effects of systemic sildenafil, tadalafil, and vardenafil treatments on skin flap survival in rats. J. Plast. Surg. Hand Surg. 2015, 49, 358–362. [Google Scholar] [CrossRef]
  114. Hart, K.; Baur, D.; Hodam, J.; Lesoon-Wood, L.; Parham, M.; Keith, K.; Vazquez, R.; Ager, E.; Pizarro, J. Short- and Long-Term Effects of Sildenafil on Skin Flap Survival in Rats. Laryngoscope 2006, 116, 522–528. [Google Scholar] [CrossRef]
  115. Souza, R.A.C.; Martinelli-Kläy, C.P.; D’acampora, A.J.; Bernardes, G.J.S.; Sgrott, S.M.; Souza, L.A.C.; Lombardi, T.; Sudbrack, T.R. Effects of sildenafil and tadalafil on skin flap viability. Arch. Dermatol. Res. 2021, 314, 151–157. [Google Scholar] [CrossRef]
  116. Abu Dayyih, W.; Abu Rayyan, W.; Al-Matubsi, H.Y. Impact of sildenafil-containing ointment on wound healing in healthy and experimental diabetic rats. Acta Diabetol. 2020, 57, 1351–1358. [Google Scholar] [CrossRef] [PubMed]
  117. Derici, H.; Kamer, E.; Ünalp, H.R.; Diniz, G.; Bozdag, A.D.; Tansug, T.; Ortac, R.; Erbil, Y. Effect of sildenafil on wound healing: An experimental study. Langenbeck’s Arch. Surg. 2009, 395, 713–718. [Google Scholar] [CrossRef] [PubMed]
  118. Gürsoy, K.; Oruç, M.; Kankaya, Y.; Ulusoy, M.G.; Koçer, U.; Kankaya, D.; Gürsoy, R.N.; Çevik, Ö.; Öğüş, E.; Fidanci, V. Effect of topically applied sildenafil citrate on wound healing: Experimental study. Biomol. Biomed. 2014, 14, 125–131. [Google Scholar] [CrossRef] [PubMed]
  119. Gad, S.B.; Hafez, M.H.; El-Sayed, Y.S. Platelet-rich plasma and/or sildenafil topical applications accelerate and better repair wound healing in rats through regulation of proinflammatory cytokines and collagen/TGF-β1 pathway. Environ. Sci. Pollut. Res. 2020, 27, 40757–40768. [Google Scholar] [CrossRef]
  120. Cakmak, E.; Yesilada, A.K.; Sevim, K.Z.; Sumer, O.; Tatlidede, H.S.; Sakiz, D. Effect of sildenafil citrate on secondary healing in full thickness skin defects in experiment. Bratisl. Med. J. 2014, 115, 267–271. [Google Scholar] [CrossRef]
  121. Kulshrestha, S.; Chawla, R.; Alam, T.; Adhikari, J.; Basu, M. Efficacy and dermal toxicity analysis of Sildenafil citrate based topical hydrogel formulation against traumatic wounds. Biomed. Pharmacother. 2019, 112, 108571. [Google Scholar] [CrossRef]
  122. Tas, A.; Karasu, A.; Comba, B.; Aksu, D.; Duz, E.; Tanritanir, P. Effects of Sildenafil Citrate on the Hematological Parameters in the Early Phase of Wound Healing in Diabetic Rats. Asian J. Anim. Vet. Adv. 2011, 6, 290–296. [Google Scholar] [CrossRef]
  123. Barakat, W.; Anwar, H.; Fouad, M. Sildenafil and Vinpocetine Promote Wound Healing in Diabetic Rats. J. Adv. Pharm. Res. 2020, 5, 221. [Google Scholar] [CrossRef]
  124. Sorour, O.A.; Nassar, E.; Sarhan, N.; El-Anwar, N.; ElKholy, R.A.; Tahoon, D.M.; Sweilam, A.; Tadros, D. Chronic sildenafil citrate use decreases retinal vascular endothelial growth factor expression in diabetic rats: A pilot study. Int. J. Retin. Vitr. 2023, 9, 42. [Google Scholar] [CrossRef]
  125. Arora, S.; Surakiatchanukul, T.; Arora, T.; Cagini, C.; Lupidi, M.; Chhablani, J. Sildenafil in ophthalmology: An update. Surv. Ophthalmol. 2021, 67, 463–487. [Google Scholar] [CrossRef]
  126. Koksal, M.; Ozdemir, H.; Kargi, S.; Yesilli, C.; Tomaç, S.; Mahmutyazicioglu, K.; Mungan, A. The effects of sildenafil on ocular blood flow. Acta Ophthalmol. Scand. 2005, 83, 355–359. [Google Scholar] [CrossRef] [PubMed]
  127. Foresta, C.; Caretta, N.; Zuccarello, D.; Poletti, A.; Biagioli, A.; Caretti, L.; Galan, A. Expression of the PDE5 enzyme on human retinal tissue: New aspects of PDE5 inhibitors ocular side effects. Eye 2007, 22, 144–149. [Google Scholar] [CrossRef] [PubMed]
  128. Cavallaro, G.; Filippi, L.; Bagnoli, P.; La Marca, G.; Cristofori, G.; Raffaeli, G.; Padrini, L.; Araimo, G.; Fumagalli, M.; Groppo, M.; et al. The pathophysiology of retinopathy of prematurity: An update of previous and recent knowledge. Acta Ophthalmol. 2014, 92, 2–20. [Google Scholar] [CrossRef] [PubMed]
  129. Da Cruz, N.F.S.; Polizelli, M.U.; Cezar, L.M.; Cardoso, E.B.; Penha, F.; Farah, M.E.; Rodrigues, E.B.; Novais, E.A. Effects of phosphodiesterase type 5 inhibitors on choroid and ocular vasculature: A literature review. Int. J. Retin. Vitr. 2020, 6, 38. [Google Scholar] [CrossRef] [PubMed]
  130. Fawzi, A.A.; Chou, J.C.; Kim, G.A.; Rollins, S.D.; Taylor, J.M.; Farrow, K.N. Sildenafil Attenuates Vaso-Obliteration and Neovascularization in a Mouse Model of Retinopathy of Prematurity. Investig. Ophthalmol. Vis. Sci. 2014, 55, 1493–1501. [Google Scholar] [CrossRef] [PubMed]
  131. Peresypkina, A.; Belgorod State National Research University; Gubareva, V.; Levkova, E.; Shabelnikova, A. Correction of Retinal Angiopathy of Hypertensive Type by Minoxidil, Sildenafil in Experiment. Res. Result Pharmacol. Clin. Pharmacol. 2016, 2, 34–44. [Google Scholar] [CrossRef]
  132. Bélanger, A.; Li, S.; Poon, A.; Jung, S.; Balian, P.; Khoja, Z.; Polosa, A.; Lachapelle, P.; Wintermark, P. AB014. What is the role of sildenafil in repairing retinopathy of prematurity? Ann. Eye Sci. 2018, 3, AB014. [Google Scholar] [CrossRef]
  133. Ding, P.-R.; Tiwari, A.K.; Ohnuma, S.; Lee, J.W.K.K.; An, X.; Dai, C.-L.; Lu, Q.-S.; Singh, S.; Yang, D.-H.; Talele, T.T.; et al. The Phosphodiesterase-5 Inhibitor Vardenafil Is a Potent Inhibitor of ABCB1/P-Glycoprotein Transporter. PLoS ONE 2011, 6, e19329. [Google Scholar] [CrossRef]
  134. Piazza, G.A.; Thompson, W.J.; Pamukcu, R.; Alila, H.W.; Whitehead, C.M.; Liu, L.; Fetter, J.R.; E Gresh, W.; Klein-Szanto, A.J.; Farnell, D.R.; et al. Exisulind, a novel proapoptotic drug, inhibits rat urinary bladder tumorigenesis. Cancer Res. 2001, 61, 3961–3968. Available online: http://www.ncbi.nlm.nih.gov/pubmed/11358813 (accessed on 1 November 2023).
  135. Pusztai, L.; Zhen, J.H.; Arun, B.; Rivera, E.; Whitehead, C.; Thompson, W.J.; Nealy, K.M.; Gibbs, A.; Symmans, W.F.; Esteva, F.J.; et al. Phase I and II Study of Exisulind in Combination with Capecitabine in Patients with Metastatic Breast Cancer. J. Clin. Oncol. 2003, 21, 3454–3461. [Google Scholar] [CrossRef]
  136. Shi, Z.; Tiwari, A.K.; Patel, A.S.; Fu, L.-W.; Chen, Z.-S. Roles of Sildenafil in Enhancing Drug Sensitivity in Cancer. Cancer Res. 2011, 71, 3735–3738. [Google Scholar] [CrossRef] [PubMed]
  137. Jedlitschky, G.; Burchell, B.; Keppler, D. The Multidrug Resistance Protein 5 Functions as an ATP-dependent Export Pump for Cyclic Nucleotides. J. Biol. Chem. 2000, 275, 30069–30074. [Google Scholar] [CrossRef] [PubMed]
  138. Chen, Z.-S.; Lee, K.; Kruh, G.D. Transport of Cyclic Nucleotides and Estradiol 17-β-d-Glucuronide by Multidrug Resistance Protein 4. J. Biol. Chem. 2001, 276, 33747–33754. [Google Scholar] [CrossRef] [PubMed]
  139. Shi, Z.; Tiwari, A.K.; Shukla, S.; Robey, R.W.; Singh, S.; Kim, I.-W.; Bates, S.E.; Peng, X.; Abraham, I.; Ambudkar, S.V.; et al. Sildenafil Reverses ABCB1- and ABCG2-Mediated Chemotherapeutic Drug Resistance. Cancer Res. 2011, 71, 3029–3041. [Google Scholar] [CrossRef] [PubMed]
  140. Hajipour, H.; Ghorbani, M.; Kahroba, H.; Mahmoodzadeh, F.; Emameh, R.Z.; Taheri, R.A. Arginyl-glycyl-aspartic acid (RGD) containing nanostructured lipid carrier co-loaded with doxorubicin and sildenafil citrate enhanced anti-cancer effects and overcomes drug resistance. Process. Biochem. 2019, 84, 172–179. [Google Scholar] [CrossRef]
  141. Greish, K.; Fateel, M.; Abdelghany, S.; Rachel, N.; Alimoradi, H.; Bakhiet, M.; Alsaie, A. Sildenafil citrate improves the delivery and anticancer activity of doxorubicin formulations in a mouse model of breast cancer. J. Drug Target. 2017, 26, 610–615. [Google Scholar] [CrossRef]
  142. El-Naa, M.M.; Othman, M.; Younes, S. Sildenafil potentiates the antitumor activity of cisplatin by induction of apoptosis and inhibition of proliferation and angiogenesis. Drug Des. Dev. Ther. 2016, 10, 3661–3672. [Google Scholar] [CrossRef]
  143. Jan, S.; Reddy, G.L.; Wani, R.; Syed, M.; Dar, M.J.; Sawant, S.D.; Vishwakarma, R.A.; Syed, S.H. Differentiation of human neuroblastoma cell line IMR-32 by sildenafil and its newly discovered analogue IS00384. Cell. Signal. 2019, 65, 109425. [Google Scholar] [CrossRef]
  144. Marques, J.G.; Gaspar, V.M.; Markl, D.; Costa, E.C.; Gallardo, E.; Correia, I.J. Co-delivery of Sildenafil (Viagra®) and Crizotinib for Synergistic and Improved Anti-tumoral Therapy. Pharm. Res. 2014, 31, 2516–2528. [Google Scholar] [CrossRef]
  145. Mei, X.-L.; Yang, Y.; Zhang, Y.-J.; Li, Y.; Zhao, J.-M.; Qiu, J.-G.; Zhang, W.-J.; Jiang, Q.-W.; Xue, Y.-Q.; Zheng, D.-W.; et al. Sildenafil inhibits the growth of human colorectal cancer in vitro and in vivo. Am. J. Cancer Res. 2015, 5, 3311–3324. [Google Scholar]
  146. Hsu, J.-L.; Leu, W.-J.; Hsu, L.-C.; Ho, C.-H.; Liu, S.-P.; Guh, J.-H. Phosphodiesterase Type 5 Inhibitors Synergize Vincristine in Killing Castration-Resistant Prostate Cancer Through Amplifying Mitotic Arrest Signaling. Front. Oncol. 2020, 10, 1274. [Google Scholar] [CrossRef] [PubMed]
  147. Morsi, D.S.; Barnawi, I.O.; Ibrahim, H.M.; El-Morsy, A.M.; El Hassab, M.A.; El Latif, H.M.A. Immunomodulatory, apoptotic and anti-proliferative potentials of sildenafil in Ehrlich ascites carcinoma murine model: In vivo and in silico insights. Int. Immunopharmacol. 2023, 119, 110135. [Google Scholar] [CrossRef] [PubMed]
  148. Islam, B.N.; Sharman, S.K.; Hou, Y.; Bridges, A.E.; Singh, N.; Kim, S.; Kolhe, R.; Trillo-Tinoco, J.; Rodriguez, P.C.; Berger, F.G.; et al. Sildenafil Suppresses Inflammation-Driven Colorectal Cancer in Mice. Cancer Prev. Res. 2017, 10, 377–388. [Google Scholar] [CrossRef] [PubMed]
  149. Meyer, C.; Sevko, A.; Ramacher, M.; Bazhin, A.V.; Falk, C.S.; Osen, W.; Borrello, I.; Kato, M.; Schadendorf, D.; Baniyash, M.; et al. Chronic inflammation promotes myeloid-derived suppressor cell activation blocking antitumor immunity in transgenic mouse melanoma model. Proc. Natl. Acad. Sci. USA 2011, 108, 17111–17116. [Google Scholar] [CrossRef] [PubMed]
  150. Chavez, A.H.; Coffield, K.S.; Rajab, M.H.; Jo, C. Incidence rate of prostate cancer in men treated for erectile dysfunction with phosphodiesterase type 5 inhibitors: Retrospective analysis. Asian J. Androl. 2013, 15, 246–248. [Google Scholar] [CrossRef] [PubMed]
  151. Danial, C.; Tichy, A.L.; Tariq, U.; Swetman, G.L.; Khuu, P.; Leung, T.H.; Benjamin, L.; Teng, J.; Vasanawala, S.S.; Lane, A.T. An open-label study to evaluate sildenafil for the treatment of lymphatic malformations. J. Am. Acad. Dermatol. 2014, 70, 1050–1057. [Google Scholar] [CrossRef]
  152. Poklepovic, A.; Qu, Y.; Dickinson, M.; Kontos, M.C.; Kmieciak, M.; Schultz, E.; Bandopadhyay, D.; Deng, X.; Kukreja, R.C. Randomized study of doxorubicin-based chemotherapy regimens, with and without sildenafil, with analysis of intermediate cardiac markers. Cardio-Oncology 2018, 4, 7. [Google Scholar] [CrossRef]
  153. Study Details|Sorafenib, Valproic Acid, and Sildenafil in Treating Patients wth Recurrent High-Grade Glioma|ClinicalTrials.gov. Available online: https://clinicaltrials.gov/study/NCT01817751?cond=Glioblastoma&intr=sildenafil&rank=1 (accessed on 2 November 2023).
  154. Socała, K.; Nieoczym, D.; Pieróg, M.; Szuster-Ciesielska, A.; Wyska, E.; Wlaź, P. Antidepressant-like activity of sildenafil following acute and subchronic treatment in the forced swim test in mice: Effects of restraint stress and monoamine depletion. Metab. Brain Dis. 2016, 31, 1095–1104. [Google Scholar] [CrossRef]
  155. Solak, Y.; Atalay, H.; Kan, S.; Kaynar, M.; Bodur, S.; Yeksan, M.; Turk, S. Effects of sildenafil and vardenafil treatments on sleep quality and depression in hemodialysis patients with erectile dysfunction. Int. J. Impot. Res. 2011, 23, 27–31. [Google Scholar] [CrossRef]
  156. Kennedy, S.H.; Dugré, H.; Defoy, I. A multicenter, double-blind, placebo-controlled study of sildenafil citrate in Canadian men with erectile dysfunction and untreated symptoms of depression, in the absence of major depressive disorder. Int. Clin. Psychopharmacol. 2011, 26, 151–158. [Google Scholar] [CrossRef]
  157. Bonierbale, M.; Tignol, J. The ELIXIR study: Evaluation of sexual dysfunction in 4557 depressed patients in France. Curr. Med. Res. Opin. 2003, 19, 114–124. [Google Scholar] [CrossRef] [PubMed]
  158. Angst, J. Sexual problems in healthy and depressed persons. Int. Clin. Psychopharmacol. 1998, 13, S1–S4. [Google Scholar] [CrossRef] [PubMed]
  159. Rizvi, S.J.; Kennedy, S.H.; Ravindran, L.N.; Giacobbe, P.; Eisfeld, B.S.; Mancini, D.; McIntyre, R.S. The Relationship between Testosterone and Sexual Function in Depressed and Healthy Men. J. Sex. Med. 2010, 7, 816–825. [Google Scholar] [CrossRef] [PubMed]
  160. Raffaele, R. Efficacy and Safety of Fixed-Dose Oral Sildenafil in the Treatment of Sexual Dysfunction in Depressed Patients with Idiopathic Parkinson’s Disease. Eur. Urol. 2002, 41, 382–386. [Google Scholar] [CrossRef] [PubMed]
  161. Dadomo, H.; Ponzi, D.; Nicolini, Y.; Vignali, A.; Ablondi, F.; Ceresini, G.; Maggio, M.; Palanza, P.; Govoni, P.; Volpi, R.; et al. Behavioral and hormonal effects of prolonged Sildenafil treatment in a mouse model of chronic social stress. Behav. Brain Res. 2020, 392, 112707. [Google Scholar] [CrossRef] [PubMed]
  162. Matsushita, H.; Matsuzaki, M.; Han, X.-J.; Nishiki, T.-I.; Ohmori, I.; Michiue, H.; Matsui, H.; Tomizawa, K. Antidepressant-like effect of sildenafil through oxytocin-dependent cyclic AMP response element-binding protein phosphorylation. Neuroscience 2012, 200, 13–18. [Google Scholar] [CrossRef]
  163. Tomaz, V.; Cordeiro, R.; Costa, A.; de Lucena, D.; Júnior, H.N.; de Sousa, F.; Vasconcelos, S.; Vale, M.; Quevedo, J.; Macêdo, D. Antidepressant-like effect of nitric oxide synthase inhibitors and sildenafil against lipopolysaccharide-induced depressive-like behavior in mice. Neuroscience 2014, 268, 236–246. [Google Scholar] [CrossRef]
  164. Brink, C.B.; Clapton, J.D.; Eagar, B.E.; Harvey, B.H. Appearance of antidepressant-like effect by sildenafil in rats after central muscarinic receptor blockade: Evidence from behavioural and neuro-receptor studies. J. Neural Transm. 2007, 115, 117–125. [Google Scholar] [CrossRef]
  165. Seidman, S.N.; Roose, S.P.; Menza, M.A.; Shabsigh, R.; Rosen, R.C. Treatment of Erectile Dysfunction in Men with Depressive Symptoms: Results of a Placebo-Controlled Trial with Sildenafil Citrate. Am. J. Psychiatry 2001, 158, 1623–1630. [Google Scholar] [CrossRef]
  166. Müller, M.J.; Benkert, O. Lower self-reported depression in patients with erectile dysfunction after treatment with sildenafil. J. Affect. Disord. 2001, 66, 255–261. [Google Scholar] [CrossRef]
  167. Afsar, B.; Ortiz, A.; Covic, A.; Gaipov, A.; Esen, T.; Goldsmith, D.; Kanbay, M. Phosphodiesterase type 5 inhibitors and kidney disease. Int. Urol. Nephrol. 2015, 47, 1521–1528. [Google Scholar] [CrossRef] [PubMed]
  168. Georgiadis, G.; Zisis, I.-E.; Docea, A.O.; Tsarouhas, K.; Fragkiadoulaki, I.; Mavridis, C.; Karavitakis, M.; Stratakis, S.; Stylianou, K.; Tsitsimpikou, C.; et al. Current Concepts on the Reno-Protective Effects of Phosphodiesterase 5 Inhibitors in Acute Kidney Injury: Systematic Search and Review. J. Clin. Med. 2020, 9, 1284. [Google Scholar] [CrossRef] [PubMed]
  169. Ma, P.; Zhang, S.; Su, X.; Qiu, G.; Wu, Z. Protective effects of icariin on cisplatin-induced acute renal injury in mice. Am. J. Transl. Res. 2015, 7, 2105–2114. Available online: http://www.ncbi.nlm.nih.gov/pubmed/26692955 (accessed on 15 November 2023). [PubMed]
  170. Choi, D.E.; Jeong, J.Y.; Lim, B.J.; Chung, S.; Chang, Y.K.; Lee, S.J.; Na, K.R.; Kim, S.Y.; Shin, Y.T.; Lee, K.W.; et al. Pretreatment of sildenafil attenuates ischemia-reperfusion renal injury in rats. Am. J. Physiol. Physiol. 2009, 297, F362–F370. [Google Scholar] [CrossRef] [PubMed]
  171. Zhu, C.-Y.; Liu, M.; Liu, Y.-Z.; Li, W.; Zhai, W.; Che, J.-P.; Yan, Y.; Wang, G.-C.; Zheng, J.-H. Preventive effect of phosphodiesterase 5 inhibitor Tadalafil on experimental post-pyelonephritic renal injury in rats. J. Surg. Res. 2013, 186, 253–261. [Google Scholar] [CrossRef] [PubMed]
  172. Hosgood, S.A.; Randle, L.V.; Patel, M.; Watson, C.J.; Bradley, J.A.; Nicholson, M.L. Sildenafil Citrate in a Donation After Circulatory Death Experimental Model of Renal Ischemia-Reperfusion Injury. Transplantation 2014, 98, 612–617. [Google Scholar] [CrossRef]
  173. Yamanaka, R.; Shindo, Y.; Hotta, K.; Suzuki, K.; Oka, K. NO/cGMP/PKG signaling pathway induces magnesium release mediated by mitoKATP channel opening in rat hippocampal neurons. FEBS Lett. 2013, 587, 2643–2648. [Google Scholar] [CrossRef]
  174. El-Azab, M.F.; Al-Karmalawy, A.A.; Antar, S.A.; Hanna, P.A.; Tawfik, K.M.; Hazem, R.M. A novel role of Nano selenium and sildenafil on streptozotocin-induced diabetic nephropathy in rats by modulation of inflammatory, oxidative, and apoptotic pathways. Life Sci. 2022, 303, 120691. [Google Scholar] [CrossRef]
  175. Mehanna, O.M.; El Askary, A.; Al-Shehri, S.; El-Esawy, B. Effect of phosphodiesterase inhibitors on renal functions and oxidant/antioxidant parameters in streptozocin-induced diabetic rats. Arch. Physiol. Biochem. 2017, 124, 424–429. [Google Scholar] [CrossRef]
  176. Webb, D.J.; Vachiery, J.-L.; Hwang, L.-J.; Maurey, J.O. Sildenafil improves renal function in patients with pulmonary arterial hypertension. Br. J. Clin. Pharmacol. 2015, 80, 235–241. [Google Scholar] [CrossRef]
  177. Grover-Páez, F.; Rivera, G.V.; Ortíz, R.G. Sildenafil citrate diminishes microalbuminuria and the percentage of A1c in male patients with type 2 diabetes. Diabetes Res. Clin. Pract. 2007, 78, 136–140. [Google Scholar] [CrossRef] [PubMed]
  178. Lauver, D.A.; Carey, E.G.; Bergin, I.L.; Lucchesi, B.R.; Gurm, H.S. Sildenafil Citrate for Prophylaxis of Nephropathy in an Animal Model of Contrast-Induced Acute Kidney Injury. PLoS ONE 2014, 9, e113598. [Google Scholar] [CrossRef] [PubMed]
  179. Iordache, A.M.; Buga, A.M.; Albulescu, D.; Vasile, R.C.; Mitrut, R.; Georgiadis, G.; Zisis, I.-E.; Mamoulakis, C.; Tsatsakis, A.; Docea, A.O.; et al. Phosphodiesterase-5 inhibitors ameliorate structural kidney damage in a rat model of contrast-induced nephropathy. Food Chem. Toxicol. 2020, 143, 111535. [Google Scholar] [CrossRef] [PubMed]
  180. Jeong, K.-H.; Lee, T.-W.; Ihm, C.-G.; Lee, S.-H.; Moon, J.-Y.; Lim, S.-J. Effects of Sildenafil on Oxidative and Inflammatory Injuries of the Kidney in Streptozotocin-Induced Diabetic Rats. Am. J. Nephrol. 2008, 29, 274–282. [Google Scholar] [CrossRef] [PubMed]
  181. El-Mahdy, N.A.; El-Sayad, M.E.-S.; El-Kadem, A.H. Combination of telmisartan with sildenafil ameliorate progression of diabetic nephropathy in streptozotocin-induced diabetic model. Biomed. Pharmacother. 2016, 81, 136–144. [Google Scholar] [CrossRef] [PubMed]
  182. Kuno, Y.; Iyoda, M.; Shibata, T.; Hirai, Y.; Akizawa, T. Sildenafil, a phosphodiesterase type 5 inhibitor, attenuates diabetic nephropathy in non-insulin-dependent Otsuka Long-Evans Tokushima Fatty rats. Br. J. Pharmacol. 2011, 162, 1389–1400. [Google Scholar] [CrossRef] [PubMed]
  183. de Almeida, L.S.; Barboza, J.R.; Freitas, F.P.S.; Porto, M.L.; Vasquez, E.C.; Meyrelles, S.S.; Gava, A.L.; Pereira, T.M.C. Sildenafil prevents renal dysfunction in contrast media-induced nephropathy in Wistar rats. Hum. Exp. Toxicol. 2016, 35, 1194–1202. [Google Scholar] [CrossRef]
  184. Altintop, I.; Tatli, M.; Karakukcu, C.; Sarica, Z.S.; Yay, A.H.; Balcioglu, E.; Ozturk, A. Serum and Tissue HIF-2 Alpha Expression in CIN, N-Acetyl Cysteine, and Sildenafil-Treated Rat Models: An Experimental Study. Medicina 2018, 54, 54. [Google Scholar] [CrossRef]
  185. Tripathi, A.S.; Mazumder, P.M.; Chandewar, A.V. Sildenafil, a phosphodiesterase type 5 inhibitor, attenuates diabetic nephropathy in STZ-induced diabetic rats. J. Basic Clin. Physiol. Pharmacol. 2015, 27, 57–62. [Google Scholar] [CrossRef]
  186. Ali, B.H.; Al Za’Abi, M.; Adham, S.A.; Al Suleimani, Y.; Karaca, T.; Manoj, P.; Al Kalbani, J.; Yasin, J.; Nemmar, A. The effect of sildenafil on rats with adenine—Induced chronic kidney disease. Biomed. Pharmacother. 2018, 108, 391–402. [Google Scholar] [CrossRef]
  187. Ali, B.H.; Abdelrahman, A.M.; Al-Salam, S.; Sudhadevi, M.; AlMahruqi, A.S.; Al-Husseni, I.S.; Beegam, S.; Dhanasekaran, S.; Nemmar, A.; Al-Moundhri, M. The Effect of Sildenafil on Cisplatin Nephrotoxicity in Rats. Basic Clin. Pharmacol. Toxicol. 2011, 109, 300–308. [Google Scholar] [CrossRef] [PubMed]
  188. Helmy, M.W.; Allah, D.M.A.; Zaid, A.M.A.; El-Din, M.M.M. Role of nitrergic and endothelin pathways modulations in cisplatin-induced nephrotoxicity in male rats. J. Physiol. Pharmacol. 2014, 65, 393–399. Available online: http://www.ncbi.nlm.nih.gov/pubmed/24930511 (accessed on 14 November 2023). [PubMed]
  189. Abdel-Latif, R.G.; Morsy, M.A.; El-Moselhy, M.A.; Khalifa, M.A. Sildenafil protects against nitric oxide deficiency-related nephrotoxicity in cyclosporine A treated rats. Eur. J. Pharmacol. 2013, 705, 126–134. [Google Scholar] [CrossRef] [PubMed]
  190. Khames, A.; Khalaf, M.M.; Gad, A.M.; El-Raouf, O.M.A. Ameliorative effects of sildenafil and/or febuxostat on doxorubicin-induced nephrotoxicity in rats. Eur. J. Pharmacol. 2017, 805, 118–124. [Google Scholar] [CrossRef] [PubMed]
  191. Mohey, V.; Singh, M.; Puri, N.; Kaur, T.; Pathak, D.; Singh, A.P. Sildenafil obviates ischemia-reperfusion injury–induced acute kidney injury through peroxisome proliferator–activated receptor γ agonism in rats. J. Surg. Res. 2015, 201, 69–75. [Google Scholar] [CrossRef] [PubMed]
  192. Tapia, E.; Sanchez-Lozada, L.G.; Soto, V.; Manrique, A.M.; Ortiz-Vega, K.M.; Santamaría, J.; Medina-Campos, O.N.; Cristóbal, M.; Avila-Casado, C.; Pedraza-Chaverri, J.; et al. Sildenafil Treatment Prevents Glomerular Hypertension and Hyperfiltration in Rats with Renal Ablation. Kidney Blood Press. Res. 2012, 35, 273–280. [Google Scholar] [CrossRef]
  193. Rodragguez-Iturbe, B.; Ferrebuz, A.; Vanegas, V.; Quiroz, Y.; Espinoza, F.; Pons, H.; Vaziri, N.D. Early treatment with cGMP phosphodiesterase inhibitor ameliorates progression of renal damage. Kidney Int. 2005, 68, 2131–2142. [Google Scholar] [CrossRef]
  194. Yousry, M.M.; Farag, E.A.; Omar, A.I. Histological Study on the Potential Effect of Sildenafil on the Kidney and Testosterone Level in Experimentally Induced Diabetes in Male Rats. J. Cytol. Histol. 2016, 7, 4. [Google Scholar] [CrossRef]
  195. Bae, E.H.; Kim, I.J.; Joo, S.Y.; Kim, E.Y.; Kim, C.S.; Choi, J.S.; Ma, S.K.; Kim, S.H.; Lee, J.U.; Kim, S.W. Renoprotective Effects of Sildenafil in DOCA-Salt Hypertensive Rats. Kidney Blood Press. Res. 2012, 36, 248–257. [Google Scholar] [CrossRef]
  196. Akgül, T.; Huri, E.; Yagmurdur, H.; Ayyıldız, A.; Üstün, H.; Germiyanoğlu, C. Phosphodiesterase 5 inhibitors attenuate renal tubular apoptosis after partial unilateral ureteral obstruction: An experimental study. Kaohsiung J. Med. Sci. 2011, 27, 15–19. [Google Scholar] [CrossRef]
  197. Küçük, A.; Yucel, M.; Erkasap, N.; Tosun, M.; Koken, T.; Ozkurt, M.; Erkasap, S. The effects of PDE5 inhibitory drugs on renal ischemia/reperfusion injury in rats. Mol. Biol. Rep. 2012, 39, 9775–9782. [Google Scholar] [CrossRef] [PubMed]
  198. Stegbauer, J.; Friedrich, S.; Potthoff, S.A.; Broekmans, K.; Cortese-Krott, M.M.; Quack, I.; Rump, L.C.; Koesling, D.; Mergia, E. Phosphodiesterase 5 Attenuates the Vasodilatory Response in Renovascular Hypertension. PLoS ONE 2013, 8, e80674. [Google Scholar] [CrossRef] [PubMed]
  199. Dias, A.T.; Rodrigues, B.P.; Porto, M.L.; Gava, A.L.; Balarini, C.M.; Freitas, F.P.S.; Palomino, Z.; Casarini, D.E.; Campagnaro, B.P.; Pereira, T.M.C.; et al. Sildenafil ameliorates oxidative stress and DNA damage in the stenotic kidneys in mice with renovascular hypertension. J. Transl. Med. 2014, 12, 35. [Google Scholar] [CrossRef] [PubMed]
  200. Sonneveld, R.; Hoenderop, J.G.; Isidori, A.M.; Henique, C.; Dijkman, H.B.; Berden, J.H.; Tharaux, P.-L.; van der Vlag, J.; Nijenhuis, T. Sildenafil Prevents Podocyte Injury via PPAR-γ–Mediated TRPC6 Inhibition. J. Am. Soc. Nephrol. 2016, 28, 1491–1505. [Google Scholar] [CrossRef] [PubMed]
  201. Lledo-Garcia, E.; Subira-Rios, D.; Ogaya-Pinies, G.; Tejedor-Jorge, A.; Del Cañizo-Lopez, J.F.; Hernandez-Fernandez, C. Intravenous Sildenafil as a Preconditioning Drug Against Hemodynamic Consequences of Warm Ischemia-Reperfusion on the Kidney. J. Urol. 2011, 186, 331–333. [Google Scholar] [CrossRef] [PubMed]
  202. Patel, N.N.; Lin, H.; Toth, T.; Jones, C.; Ray, P.; Welsh, G.I.; Satchell, S.C.; Sleeman, P.; Angelini, G.D.; Murphy, G.J. Phosphodiesterase-5 Inhibition Prevents Postcardiopulmonary Bypass Acute Kidney Injury in Swine. Ann. Thorac. Surg. 2011, 92, 2168–2176. [Google Scholar] [CrossRef] [PubMed]
  203. Lledo-Garcia, E.; Rodriguez-Martinez, D.; Cabello-Benavente, R.; Moncada-Iribarren, I.; Tejedor-Jorge, A.; Dulin, E.; Hernandez-Fernandez, C.; Del Canizo-Lopez, J. Sildenafil Improves Immediate Posttransplant Parameters in Warm-Ischemic Kidney Transplants: Experimental Study. Transplant. Proc. 2007, 39, 1354–1356. [Google Scholar] [CrossRef]
  204. Lledó-García, E.; Subirá-Ríos, D.; Rodríguez–Martínez, D.; Dulín, E.; Alvarez-Fernández, E.; Hernández-Fernández, C.; del Cañizo-López, J.F. Sildenafil as a Protecting Drug for Warm Ischemic Kidney Transplants: Experimental Results. J. Urol. 2009, 182, 1222–1225. [Google Scholar] [CrossRef]
  205. Zahran, M.H.; Barakat, N.; Khater, S.; Awadalla, A.; Mosbah, A.; Nabeeh, A.; Hussein, A.M.; Shokeir, A.A. Renoprotective effect of local sildenafil administration in renal ischaemia–reperfusion injury: A randomised controlled canine study. Arab. J. Urol. 2019, 17, 150–159. [Google Scholar] [CrossRef]
  206. Malfertheiner, P.; Chan, F.K.; EL McColl, K. Peptic ulcer disease. Lancet 2009, 374, 1449–1461. [Google Scholar] [CrossRef]
  207. Tarnawski, A.S.; Ahluwalia, A.; Jones, M.K. Angiogenesis in gastric mucosa: An important component of gastric erosion and ulcer healing and its impairment in aging. J. Gastroenterol. Hepatol. 2014, 29, 112–123. [Google Scholar] [CrossRef] [PubMed]
  208. Kalayci, M.; Kocdor, M.A.; Kuloglu, T.; Sahin, I.; Sarac, M.; Aksoy, A.; Yardim, M.; Dalkilic, S.; Gursu, O.; Aydin, S.; et al. Comparison of the therapeutic effects of sildenafil citrate, heparin and neuropeptides in a rat model of acetic acid-induced gastric ulcer. Life Sci. 2017, 186, 102–110. [Google Scholar] [CrossRef] [PubMed]
  209. Lanas, A. Role of nitric oxide in the gastrointestinal tract. Arthritis Res. Ther. 2008, 10 (Suppl. S2), S4. [Google Scholar] [CrossRef] [PubMed]
  210. Moustafa, Y.M.; Khoder, D.M.; El-Awady, E.E.; Zaitone, S.A. Sildenafil citrate protects against gastric mucosal damage induced by indomethacin in rats. Eur. Rev. Med. Pharmacol. Sci. 2013, 17, 179–188. Available online: http://www.ncbi.nlm.nih.gov/pubmed/23377805 (accessed on 14 November 2023). [PubMed]
  211. Karakoyun, B.; Uslu, U.; Ercan, F.; Aydin, M.S.; Yuksel, M.; Ogunc, A.V.; Alican, I. The effect of phosphodiesterase-5 inhibition by sildenafil citrate on inflammation and apoptosis in rat experimental colitis. Life Sci. 2011, 89, 402–407. [Google Scholar] [CrossRef] [PubMed]
  212. Elson, C.O.; Sartor, R.; Tennyson, G.S.; Riddell, R.H. Experimental models of inflammatory bowel disease. Gastroenterology 1995, 109, 1344–1367. [Google Scholar] [CrossRef]
  213. Podolsky, D.K. Pride and prejudice: Inflammatory bowel disease models and drug development. Curr. Opin. Gastroenterol. 2000, 16, 295–296. [Google Scholar] [CrossRef]
  214. Strober, W. Interactions between Epithelial Cells and Immune Cells in the Intestine. Ann. New York Acad. Sci. 1998, 859, 37–45. [Google Scholar] [CrossRef]
  215. Yildirim, A.; Ersoy, Y.; Ercan, F.; Atukeren, P.; Gumustas, K.; Uslu, U.; Alican, I. Phosphodiesterase-5 inhibition by sildenafil citrate in a rat model of bleomycin-induced lung fibrosis. Pulm. Pharmacol. Ther. 2010, 23, 215–221. [Google Scholar] [CrossRef]
  216. Iseri, S.O.; Ersoy, Y.; Ercan, F.; Yuksel, M.; Atukeren, P.; Gumustas, K.; Alican, I. The effect of sildenafil, a phosphodiesterase-5 inhibitor, on acetic acid-induced colonic inflammation in the rat. J. Gastroenterol. Hepatol. 2009, 24, 1142–1148. [Google Scholar] [CrossRef]
  217. Medeiros, J.V.R.; Gadelha, G.G.; Lima, S.J.; Garcia, J.A.; Soares, P.M.G.; Santos, A.A.; Brito, G.A.C.; Ribeiro, R.A.; Souza, M.H.L.P. Role of the NO/cGMP/KATP pathway in the protective effects of sildenafil against ethanol-induced gastric damage in rats. Br. J. Pharmacol. 2008, 153, 721–727. [Google Scholar] [CrossRef] [PubMed]
  218. Aydinli, B.; Yildirgan, M.I.; Oztürk, G.; Atamanalap, S.S.; Polat, K.Y.; Başoğlu, M.; Gündoğdu, C.; Süleyman, H.; Kiziltunç, A.; Gürsan, N.; et al. The role of sildenafil citrate in the protection of gastric mucosa from nonsteroidal anti-inflammatory drug-induced damage. Ulus. Travma Acil Cerrahi Derg. 2007, 13, 268–273. Available online: http://www.ncbi.nlm.nih.gov/pubmed/17978907 (accessed on 14 November 2023). [PubMed]
  219. Elberry, A.A. Protective effect of sildenafil against cysteamine induced duodenal ulcer in Wistar rats. Afr. J. Pharm. Pharmacol. 2013, 7, 2352–2357. [Google Scholar] [CrossRef]
  220. Alves, G.M.; Aires, R.; Santos, V.D.S.; Côco, L.Z.; Peters, B.; Assis, A.D.L.E.M.; Athaydes, B.R.; Amorim, F.G.; Nogueira, B.V.; Gonçalves, R.C.d.R.; et al. Sildenafil attenuates nonsteroidal anti-inflammatory-induced gastric ulceration in mice via antioxidant and antigenotoxic mechanisms. Clin. Exp. Pharmacol. Physiol. 2020, 48, 401–411. [Google Scholar] [CrossRef] [PubMed]
  221. El Mahdy, R.N.; Risha, S.; Sisi, A.; Sobhy, W. Potential Protective Effects of Sildenafil and Moringa on Experimentally-induced Gastric Ulcer in Rats. Int. J. Cancer Biomed. Res. 2020, 4, 43–55. [Google Scholar] [CrossRef]
  222. El-Sisi, A.E.; Sokar, S.S.; Abu-Risha, S.E.; Khira, D.Y. The potential beneficial effects of sildenafil and diosmin in experimentally-induced gastric ulcer in rats. Heliyon 2020, 6, e04761. [Google Scholar] [CrossRef]
  223. Margonis, G.A.; Christoloukas, N.; Antoniou, E.; Arkadopoulos, N.; Theodoropoulos, G.; Agrogiannis, G.; Pikoulis, E.; Patsouris, E.S.; Zografos, G.C.; Papalois, A.E. Effectiveness of sildenafil and U-74389G in a rat model of colitis. J. Surg. Res. 2015, 193, 667–674. [Google Scholar] [CrossRef]
  224. Fakhfouri, G.; Rahimian, R.; Hashemi, S.; Rasouli, M.R.; Bahremand, A.; Mehr, S.E.; Khorramizadeh, M.R.; Dehpour, A.R. Sildenafil attenuates TNBS-induced colitis in rats: Possible involvement of cGMP and KATP channels. Fundam. Clin. Pharmacol. 2012, 26, 190–193. [Google Scholar] [CrossRef]
  225. Kato, N.; Mashita, Y.; Kato, S.; Mitsufuji, S.; Yoshikawa, T.; Takeuchi, K. Sildenafil, an Inhibitor of Phosphodiesterase Subtype 5, Prevents Indomethacin-Induced Small-Intestinal Ulceration in Rats via a NO/cGMP-Dependent Mechanism. Dig. Dis. Sci. 2009, 54, 2346–2356. [Google Scholar] [CrossRef]
  226. Kim, G.E.; Kass, D.A. Cardiac Phosphodiesterases and Their Modulation for Treating Heart Disease. In Heart Failure; Springer: Berlin/Heidelberg, Germany, 2016; pp. 249–269. [Google Scholar] [CrossRef]
  227. Roy, S.; Kloner, R.A.; Salloum, F.N.; Jovin, I.S. Cardiac Effects of Phosphodiesterase-5 Inhibitors: Efficacy and Safety. Cardiovasc. Drugs Ther. 2023, 37, 793–806. [Google Scholar] [CrossRef]
  228. Zhang, M.; Kass, D.A. Phosphodiesterases and cardiac cGMP: Evolving roles and controversies. Trends Pharmacol. Sci. 2011, 32, 360–365. [Google Scholar] [CrossRef] [PubMed]
  229. Stuckey, B.G.; Jadzinsky, M.N.; Murphy, L.J.; Montorsi, F.; Kadioglu, A.; Fraige, F.; Manzano, P.; Deerochanawong, C. Sildenafil Citrate for Treatment of Erectile Dysfunction in Men with Type 1 Diabetes. Diabetes Care 2003, 26, 279–284. [Google Scholar] [CrossRef] [PubMed]
  230. Bocchi, E.A.; Guimarães, G.; Mocelin, A.; Bacal, F.; Bellotti, G.; Ramires, J.F. Sildenafil Effects on Exercise, Neurohormonal Activation, and Erectile Dysfunction in Congestive Heart Failure. Circulation 2002, 106, 1097–1103. [Google Scholar] [CrossRef] [PubMed]
  231. Korkmaz-Icöz, S.; Radovits, T.; Szabó, G. Targeting phosphodiesterase 5 as a therapeutic option against myocardial ischaemia/reperfusion injury and for treating heart failure. Br. J. Pharmacol. 2017, 175, 223–231. [Google Scholar] [CrossRef] [PubMed]
  232. Westermann, D.; Becher, P.M.; Lindner, D.; Savvatis, K.; Xia, Y.; Fröhlich, M.; Hoffmann, S.; Schultheiss, H.-P.; Tschöpe, C. Selective PDE5A inhibition with sildenafil rescues left ventricular dysfunction, inflammatory immune response and cardiac remodeling in angiotensin II-induced heart failure in vivo. Basic Res. Cardiol. 2012, 107, 308. [Google Scholar] [CrossRef]
  233. Nagy, O.; Hajnal, Á.; Parratt, J.R.; Végh, Á. Sildenafil (Viagra) reduces arrhythmia severity during ischaemia 24 h after oral administration in dogs. Br. J. Pharmacol. 2004, 141, 549–551. [Google Scholar] [CrossRef] [PubMed]
  234. Guazzi, M.; Samaja, M.; Arena, R.; Vicenzi, M.; Guazzi, M.D. Long-Term Use of Sildenafil in the Therapeutic Management of Heart Failure. J. Am. Coll. Cardiol. 2007, 50, 2136–2144. [Google Scholar] [CrossRef]
  235. Elhakeem, W.A.; Khairy, H. Effect of sildenafil on stable chronic heart failure: A prospective randomized-controlled clinical trial. J. Med. Sci. Res. 2019, 2, 118. [Google Scholar] [CrossRef]
  236. Salloum, F.N.; Das, A.; Thomas, C.S.; Yin, C.; Kukreja, R.C. Adenosine A1 receptor mediates delayed cardioprotective effect of sildenafil in mouse. J. Mol. Cell. Cardiol. 2007, 43, 545–551. [Google Scholar] [CrossRef]
  237. Madhani, M.; Hall, A.R.; Cuello, F.; Charles, R.L.; Burgoyne, J.R.; Fuller, W.; Hobbs, A.J.; Shattock, M.J.; Eaton, P. Phospholemman Ser69 phosphorylation contributes to sildenafil-induced cardioprotection against reperfusion injury. Am. J. Physiol. Circ. Physiol. 2010, 299, H827–H836. [Google Scholar] [CrossRef]
  238. Behmenburg, F.; Dorsch, M.; Huhn, R.; Mally, D.; Heinen, A.; Hollmann, M.W.; Berger, M.M. Impact of Mitochondrial Ca2+-Sensitive Potassium (mBKCa) Channels in Sildenafil-Induced Cardioprotection in Rats. PLoS ONE 2015, 10, e0144737. [Google Scholar] [CrossRef] [PubMed]
  239. Banjac, N.; Vasović, V.; Stilinović, N.; Tomas, A.; Vasović, L.; Martić, N.; Prodanović, D.; Jakovljević, V. The Effects of Different Doses of Sildenafil on Coronary Blood Flow and Oxidative Stress in Isolated Rat Hearts. Pharmaceuticals 2023, 16, 118. [Google Scholar] [CrossRef] [PubMed]
  240. Skeffington, K.L.; Ahmed, E.M.; Rapetto, F.; Chanoit, G.; Bond, A.R.; Vardeu, A.; Ghorbel, M.T.; Suleiman, M.-S.; Caputo, M. The effect of cardioplegic supplementation with sildenafil on cardiac energetics in a piglet model of cardiopulmonary bypass and cardioplegic arrest with warm or cold cardioplegia. Front. Cardiovasc. Med. 2023, 10, 1194645. [Google Scholar] [CrossRef] [PubMed]
  241. Imai, Y.; Kariya, T.; Iwakiri, M.; Yamada, Y.; Takimoto, E. Sildenafil ameliorates right ventricular early molecular derangement during left ventricular pressure overload. PLoS ONE 2018, 13, e0195528. [Google Scholar] [CrossRef] [PubMed]
  242. Blanton, R.M.; Takimoto, E.; Lane, A.M.; Aronovitz, M.; Piotrowski, R.; Karas, R.H.; Kass, D.A.; Mendelsohn, M.E. Protein Kinase G Iα Inhibits Pressure Overload–Induced Cardiac Remodeling and Is Required for the Cardioprotective Effect of Sildenafil In Vivo. J. Am. Heart Assoc. 2012, 1, e003731. [Google Scholar] [CrossRef]
  243. Wang, X.; Fisher, P.W.; Xi, L.; Kukreja, R.C. Essential role of mitochondrial Ca2+-activated and ATP-sensitive K+ channels in sildenafil-induced late cardioprotection. J. Mol. Cell. Cardiol. 2008, 44, 105–113. [Google Scholar] [CrossRef]
  244. Salloum, F.; Yin, C.; Xi, L.; Kukreja, R.C. Sildenafil Induces Delayed Preconditioning Through Inducible Nitric Oxide Synthase–Dependent Pathway in Mouse Heart. Circ. Res. 2003, 92, 595–597. [Google Scholar] [CrossRef]
  245. Shalwala, M. A Novel Role of SIRT1 in Sildenafil Induced Cardioprotection in Mice. Master’s Thesis, Virginia Commonwealth University, Richmond, VA, USA, May 2010. [Google Scholar] [CrossRef]
  246. Chau, V.Q.; Salloum, F.N.; Hoke, N.N.; Abbate, A.; Kukreja, R.C. Mitigation of the progression of heart failure with sildenafil involves inhibition of RhoA/Rho-kinase pathway. Am. J. Physiol. Circ. Physiol. 2011, 300, H2272–H2279. [Google Scholar] [CrossRef]
  247. Salloum, F.N.; Abbate, A.; Das, A.; Houser, J.-E.; Mudrick, C.A.; Qureshi, I.Z.; Hoke, N.N.; Roy, S.K.; Brown, W.R.; Prabhakar, S.; et al. Sildenafil (Viagra) attenuates ischemic cardiomyopathy and improves left ventricular function in mice. Am. J. Physiol. Circ. Physiol. 2008, 294, H1398–H1406. [Google Scholar] [CrossRef]
  248. Adamo, C.M.; Dai, D.-F.; Percival, J.M.; Minami, E.; Willis, M.S.; Patrucco, E.; Froehner, S.C.; Beavo, J.A. Sildenafil reverses cardiac dysfunction in the mdx mouse model of Duchenne muscular dystrophy. Proc. Natl. Acad. Sci. USA 2010, 107, 19079–19083. [Google Scholar] [CrossRef]
  249. Elrod, J.W.; Greer, J.J.M.; Lefer, D.J. Sildenafil-mediated acute cardioprotection is independent of the NO/cGMP pathway. Am. J. Physiol. Circ. Physiol. 2007, 292, H342–H347. [Google Scholar] [CrossRef] [PubMed]
  250. Zhu, G.; Ueda, K.; Hashimoto, M.; Zhang, M.; Sasaki, M.; Kariya, T.; Sasaki, H.; Kaludercic, N.; Lee, D.; Bedja, D.; et al. The mitochondrial regulator PGC1α is induced by cGMP–PKG signaling and mediates the protective effects of phosphodiesterase 5 inhibition in heart failure. FEBS Lett. 2021, 596, 17–28. [Google Scholar] [CrossRef] [PubMed]
  251. Lee, K.H.; Kwon, S.J.; Woo, J.-S.; Lee, G.-J.; Lee, S.-R.; Jang, H.-H.; Kim, H.S.; Kim, J.W.; Park, H.K.; Cho, K.S.; et al. Effects of sildenafil on nanostructural and nanomechanical changes in mitochondria in an ischaemia-reperfusion rat model. Clin. Exp. Pharmacol. Physiol. 2014, 41, 763–768. [Google Scholar] [CrossRef] [PubMed]
  252. Botha, P.; MacGowan, G.A.; Dark, J.H. Sildenafil Citrate Augments Myocardial Protection in Heart Transplantation. Transplantation 2010, 89, 169–177. [Google Scholar] [CrossRef] [PubMed]
  253. Ekinci, A.; Aşır, F.; Dursun, R. Protective Effect of Sildenafil on the Heart in Hepatic Ischemia/Reperfusion Injury. Anal. Quant. Cytopathol. Histopathol. 2021, 43, 116–120. Available online: https://www.researchgate.net/publication/354144820 (accessed on 11 October 2023).
  254. Koneru, S.; Penumathsa, S.V.; Thirunavukkarasu, M.; Vidavalur, R.; Zhan, L.; Singal, P.K.; Engelman, R.M.; Das, D.K.; Maulik, N. Sildenafil-mediated neovascularization and protection against myocardial ischaemia reperfusion injury in rats: Role of VEGF/angiopoietin-1. J. Cell. Mol. Med. 2008, 12, 2651–2664. [Google Scholar] [CrossRef]
  255. Zhang, Q.; Yuan, W.; Wang, G.; Wu, J.; Wang, M.; Li, C. The protective effects of a phosphodiesterase 5 inhibitor, sildenafil, on postresuscitation cardiac dysfunction of cardiac arrest: Metabolic evidence from microdialysis. Crit. Care 2014, 18, 641. [Google Scholar] [CrossRef]
  256. Wang, G.; Zhang, Q.; Yuan, W.; Wu, J.; Li, C. Sildenafil Protects against Myocardial Ischemia-Reperfusion Injury Following Cardiac Arrest in a Porcine Model: Possible Role of the Renin-Angiotensin System. Int. J. Mol. Sci. 2015, 16, 27015–27031. [Google Scholar] [CrossRef]
  257. He, Y.; Wang, G.; Li, C.; Wang, Y.; Zhang, Q. The protective effects of phosphodiesterase-5 inhibitor, sildenafil on post-resuscitation cardiac dysfunction of cardiac arrest: By regulating the miR-155-5p and miR-145-5p. Scand. J. Trauma Resusc. Emerg. Med. 2021, 29, 2. [Google Scholar] [CrossRef]
  258. Santiago-Vacas, E.; García-Lunar, I.; Solanes, N.; Dantas, A.P.; Ascaso, M.; Jimenez-Trinidad, F.R.; Ramirez, J.; Fernández-Friera, L.; Galán, C.; Sánchez, J.; et al. Effect of sildenafil on right ventricular performance in an experimental large-animal model of postcapillary pulmonary hypertension. Transl. Res. 2020, 228, 64–75. [Google Scholar] [CrossRef]
  259. Ockaili, R.; Salloum, F.; Hawkins, J.; Kukreja, R.C. Sildenafil (Viagra) induces powerful cardioprotective effect via opening of mitochondrial KATPchannels in rabbits. Am. J. Physiol. Circ. Physiol. 2002, 283, H1263–H1269. [Google Scholar] [CrossRef] [PubMed]
  260. Salloum, F.N.; Takenoshita, Y.; Ockaili, R.A.; Daoud, V.P.; Chou, E.; Yoshida, K.-I.; Kukreja, R.C. Sildenafil and vardenafil but not nitroglycerin limit myocardial infarction through opening of mitochondrial KATP channels when administered at reperfusion following ischemia in rabbits. J. Mol. Cell. Cardiol. 2007, 42, 453–458. [Google Scholar] [CrossRef] [PubMed]
  261. Halcox, J.P.; Nour, K.R.; Zalos, G.; Mincemoyer, R.; Waclawiw, M.A.; Rivera, C.E.; Willie, G.; Ellahham, S.; Quyyumi, A.A. The effect of sildenafil on human vascular function, platelet activation, and myocardial ischemia. J. Am. Coll. Cardiol. 2002, 40, 1232–1240. [Google Scholar] [CrossRef] [PubMed]
  262. Lewis, G.D.; Shah, R.; Shahzad, K.; Camuso, J.M.; Pappagianopoulos, P.P.; Hung, J.; Tawakol, A.; Gerszten, R.E.; Systrom, D.M.; Bloch, K.D.; et al. Sildenafil Improves Exercise Capacity and Quality of Life in Patients with Systolic Heart Failure and Secondary Pulmonary Hypertension. Circulation 2007, 116, 1555–1562. [Google Scholar] [CrossRef] [PubMed]
  263. Giannetta, E.; Isidori, A.M.; Galea, N.; Carbone, I.; Mandosi, E.; Vizza, C.D.; Naro, F.; Morano, S.; Fedele, F.; Lenzi, A. Chronic Inhibition of cGMP Phosphodiesterase 5A Improves Diabetic Cardiomyopathy. Circulation 2012, 125, 2323–2333. [Google Scholar] [CrossRef] [PubMed]
  264. Redfield, M.M.; Chen, H.H.; Borlaug, B.A.; Semigran, M.J.; Lee, K.L.; Lewis, G.; LeWinter, M.M.; Rouleau, J.L.; Bull, D.A.; Mann, D.L.; et al. Effect of Phosphodiesterase-5 Inhibition on Exercise Capacity and Clinical Status in Heart Failure with Preserved Ejection Fraction. JAMA 2013, 309, 1268–1277. [Google Scholar] [CrossRef]
  265. Fox, K.M.; Thadani, U.; Ma, P.T.; Nash, S.D.; Keating, Z.; Czorniak, M.A.; Gillies, H.; Keltai, M. Sildenafil citrate does not reduce exercise tolerance in men with erectile dysfunction and chronic stable angina. Eur. Heart J. 2003, 24, 2206–2212. [Google Scholar] [CrossRef]
  266. Denardo, S.J.; Wen, X.; Handberg, E.M.; Merz, C.N.B.; Sopko, G.S.; Cooper-DeHoff, R.M.; Pepine, C.J. Effect of Phosphodiesterase Type 5 Inhibition on Microvascular Coronary Dysfunction in Women: A Women’s Ischemia Syndrome Evaluation (WISE) Ancillary Study. Clin. Cardiol. 2011, 34, 483–487. [Google Scholar] [CrossRef]
  267. Kiss, T.; Kovacs, K.; Komocsi, A.; Tornyos, A.; Zalan, P.; Sumegi, B.; Jr, F.G. Novel Mechanisms of Sildenafil in Pulmonary Hypertension Involving Cytokines/Chemokines, MAP Kinases and Akt. PLoS ONE 2014, 9, e104890. [Google Scholar] [CrossRef]
  268. Al Qadi-Nassar, B.; Bichon-Laurent, F.; Portet, K.; Tramini, P.; Arnoux, B.; Michel, A. Effects of l-arginine and phosphodiesterase-5 inhibitor, sildenafil, on inflammation and airway responsiveness of sensitized BP2 mice. Fundam. Clin. Pharmacol. 2007, 21, 611–620. [Google Scholar] [CrossRef]
  269. de Visser, Y.P.; Walther, F.J.; Laghmani, E.H.; Boersma, H.; van der Laarse, A.; Wagenaar, G.T. Sildenafil attenuates pulmonary inflammation and fibrin deposition, mortality and right ventricular hypertrophy in neonatal hyperoxic lung injury. Respir. Res. 2009, 10, 30. [Google Scholar] [CrossRef] [PubMed]
  270. Park, H.-S.; Park, J.-W.; Kim, H.-J.; Choi, C.W.; Lee, H.-J.; Kim, B.I.; Chun, Y.-S. Sildenafil Alleviates Bronchopulmonary Dysplasia in Neonatal Rats by Activating the Hypoxia-Inducible Factor Signaling Pathway. Am. J. Respir. Cell Mol. Biol. 2013, 48, 105–113. [Google Scholar] [CrossRef] [PubMed]
  271. Wang, T.; Liu, Y.; Chen, L.; Wang, X.; Hu, X.-R.; Feng, Y.-L.; Liu, D.-S.; Xu, D.; Duan, Y.-P.; Lin, J.; et al. Effect of sildenafil on acrolein-induced airway inflammation and mucus production in rats. Eur. Respir. J. 2009, 33, 1122–1132. [Google Scholar] [CrossRef] [PubMed]
  272. Song, J.; Wang, Y.-Y.; Song, P.; Li, L.-P.; Fan, Y.-F.; Wang, Y.-L. Original Article Sildenafil ameliorated meconium-induced acute lung injury in a neonatal rat model. Int. J. Clin. Exp. Med. 2016, 9, 10238–10246. Available online: https://e-century.us/files/ijcem/9/6/ijcem0022567.pdf (accessed on 10 October 2023).
  273. Kosutova, P.; Mikolka, P.; Balentova, S.; Kolomaznik, M.; Adamkov, M.; Mokry, J.; Calkovska, A.; Mokra, D. Effects of phosphodiesterase 5 inhibitor sildenafil on the respiratory parameters, inflammation and apoptosis in a saline lavage-induced model of acute lung injury. JPP 2018, 5, 15. [Google Scholar]
  274. Rodriguez-Miguelez, P.; Ishii, H.; Seigler, N.; Crandall, R.; Thomas, J.; Forseen, C.; McKie, K.T.; Harris, R.A. Sildenafil improves exercise capacity in patients with cystic fibrosis: A proof-of-concept clinical trial. Ther. Adv. Chronic Dis. 2019, 10, 1–13. [Google Scholar] [CrossRef] [PubMed]
  275. Taylor-Cousar, J.; Wiley, C.; Felton, L.; Clair, C.S.; Jones, M.; Curran-Everett, D.; Poch, K.; Nichols, D.; Solomon, G.; Saavedra, M.; et al. Pharmacokinetics and tolerability of oral sildenafil in adults with cystic fibrosis lung disease. J. Cyst. Fibros. 2014, 14, 228–236. [Google Scholar] [CrossRef]
  276. Jackson, R.M.; Glassberg, M.K.; Ramos, C.F.; Bejarano, P.A.; Butrous, G.; Gómez-Marín, O. Sildenafil Therapy and Exercise Tolerance in Idiopathic Pulmonary Fibrosis. Lung 2009, 188, 115–123. [Google Scholar] [CrossRef]
  277. Moon, R.E.; Martina, S.D.; Peacher, D.F.; Potter, J.F.; Wester, T.E.; Cherry, A.D.; Natoli, M.J.; Otteni, C.E.; Kernagis, D.N.; White, W.D.; et al. Swimming-Induced Pulmonary Edema. Circulation 2016, 133, 988–996. [Google Scholar] [CrossRef]
  278. Andersen, A.; Waziri, F.; Schultz, J.G.; Holmboe, S.; Becker, S.W.; Jensen, T.; Søndergaard, H.M.; Dodt, K.K.; May, O.; Mortensen, U.M.; et al. Pulmonary vasodilation by sildenafil in acute intermediate-high risk pulmonary embolism: A randomized explorative trial. BMC Pulm. Med. 2021, 21, 72. [Google Scholar] [CrossRef]
  279. Reisi, M.; Modaresi, M.R.; Aghaii, Z.; Mirlohi, S.H.; Rafiemanesh, H.; Azizi, G.; Sayedi, S.J. Efficacy and safety of oral sildenafil in cystic fibrosis children with mild to moderate lung disease. Pediatr. Pulmonol. 2019, 55, 156–160. [Google Scholar] [CrossRef] [PubMed]
  280. Collard, H.R.; Anstrom, K.J.; Schwarz, M.I.; Zisman, D.A. Sildenafil Improves Walk Distance in Idiopathic Pulmonary Fibrosis. Chest 2007, 131, 897–899. [Google Scholar] [CrossRef] [PubMed]
  281. Behr, J.; Nathan, S.D.; Wuyts, W.A.; Bishop, N.M.; Bouros, D.E.; Antoniou, K.; Guiot, J.; Kramer, M.R.; Kirchgaessler, K.-U.; Bengus, M.; et al. Efficacy and safety of sildenafil added to pirfenidone in patients with advanced idiopathic pulmonary fibrosis and risk of pulmonary hypertension: A double-blind, randomised, placebo-controlled, phase 2b trial. Lancet Respir. Med. 2020, 9, 85–95. [Google Scholar] [CrossRef] [PubMed]
  282. Kolb, M.; Raghu, G.; Wells, A.U.; Behr, J.; Richeldi, L.; Schinzel, B.; Quaresma, M.; Stowasser, S.; Martinez, F.J. Nintedanib plus Sildenafil in Patients with Idiopathic Pulmonary Fibrosis. N. Engl. J. Med. 2018, 379, 1722–1731. [Google Scholar] [CrossRef]
Figure 1. The mechanism of action of sildenafil and its subsequent potential indications (X-inhibition; Medicina 59 02190 i001 increased; created with BioRender version, Bio Rad 2023).
Figure 1. The mechanism of action of sildenafil and its subsequent potential indications (X-inhibition; Medicina 59 02190 i001 increased; created with BioRender version, Bio Rad 2023).
Medicina 59 02190 g001
Table 1. Preclinical studies that evaluated the effect of sildenafil on pain.
Table 1. Preclinical studies that evaluated the effect of sildenafil on pain.
Pain
AnimalsAnimal ModelDosageResults
db/db miceDiabetic mice (type 2 diabetes)10 mg/kg bw,
orally
Sildenafil increased
blood vessel functionality, regional blood flow in the sciatic nerve, and
motor and sensory
conduction velocities in the sciatic nerve, as well as ameliorated the
sensitivity to peripheral heat stimuli in the tail flick test.
[57]
NMRI male miceAlloxan-induced diabetic neuropathy1.5, 2.5, 3 mg/kg bw,
orally
Sildenafil increased the pain reaction latency in both hot-plate tests and tail-withdrawal tests from cold water
[58].
NMRI male miceAlloxan-induced diabetic neuropathySildenafil 1.5, 2.5, 3 mg/kg bw
+
Metformin 150, 250, 500 mg/kg bw,
orally
The combination reversed thermal hyperalgesia in hot-plate and tail-withdrawal tests.
It reduced the
activity of iNOS in the brain and liver and IL-6 brain concentration.
[22]
Wistar female ratsFormalin-induced pain50, 100, and 200 µg into the ipsilateral paw
20 min before formalin
Locally administered sildenafil resulted in dose-dependent
antinociception through the activation of the cyclic GMP-PKG-K+ channel pathway
[59].
Wistar female ratsFormalin-induced pain50, 100, and 200 µg,
i.pl.
Sildenafil exhibited a dose-dependent antinociceptive effect and enhanced morphine-induced analgesia
[53].
Wistar rats

Albino mice
Carrageenan-induced hyperalgesia

Acetic acid-induced pain
1, 2, 5, and 10 mg/kg,
i.p.

50–200 µg,
i.pl.
Administered i.p., sildenafil reduced acetic acid-induced writhing
in mice in a dose-dependent
manner.
Local administration of sildenafil reduced the intensity of
hyperalgesia induced by
carrageenan
[60].
Wistar rats

Albino mice
Streptozotocin-induced diabetic neuropathy1–10 mg/kg, i.p

50–400 µg/paw, i.pl.
Sildenafil increased the pain reaction latency in writhing tests (mice) and paw hyperalgesia tests (rats).
[61].
Wistar male ratsStreptozotocin-induced diabetic neuropathy5 mg/kg bw, orallySildenafil effectively and
consistently enhanced vascular
function both in terms of immediate and long-term effects
[62].
Sprague–Dawley male ratsLigation of L5/6 spinal nerves induced neuropathic pain1, 3, 10, and 30 mg/kg bw
i.p.
Sildenafil increased the withdrawal threshold in von Frey tests in a dose-dependent manner
[55].
bw, body weight; GMP-PKG, cGMP-dependent protein kinase G; i.pl., intraplantar; IL6, interleukin 6; iNOS, inducible nitric oxide synthase.
Table 2. Preclinical and clinical studies that evaluated the effect of sildenafil on Alzheimer’s disease.
Table 2. Preclinical and clinical studies that evaluated the effect of sildenafil on Alzheimer’s disease.
Alzheimer’s Disease
In Vitro Studies
PopulationDosageResults
HT-22 mouse hippocampal neuronal cells treated with Aβ peptide10–100 μMProtects the mitochondria of neuronal cells from Aβ-induced injury. This effect is
dependent on
mitochondrial KATP channels.
[73]
HT-22 mouse hippocampal neuronal cells exposed to advanced glycation end products20 μMReduced the opening of mitochondrial permeability transition pores and HO1 induced
apoptosis.
[74]
Preclinical studies
PS1/APP mice3 mg/kg bw,
i.p.
Enhanced immediate and prolonged synaptic function, phosphorylation of CREB, and memory; decreased Aß concentrations
[75].
PS1/APP mice10 mg/kg bw,
i.p
Effectively mitigated memory impairments, restored the proper functioning of the cGMP/PKG/pCREB signalling pathway, and decreased Aβ1-40, Aβ1-42, IL-1β, IL-6, and TNF-α
[76].
PS1/APP mice6 mg/kg bw,
i.p.
Enhanced memory, reduced amyloid plaque accumulation, mitigated inflammatory processes, and promoted neurogenesis
[77].
PS1/APP mice2 mg/kg bw,
orally
Ameliorated memory deficits, reduced amyloid pathology, and increased the NOS, NO, and cGMP
[78].
Sprague–Dawley rats with scopolamine-induced cholinergic dysfunction 1.5, 3.0, 4.5 mg/kg bw,
i.p.
Demonstrated beneficial effect on memory
Retrieval.
[79].
Tg2576 transgenic mice15 mg/kg bw,
i.p.
Enhanced memory, reduced tau protein levels, inhibited the activity of GSK3β, lowered the CDK5 p25/35 ratio, and upregulated the expression of BDNF and Arc proteins
[80].
J20 mice15 mg/kg bw,
in drinking water
Enhanced memory function, reduced tau
hyperphosphorylation, and increased GSK3β-mediated phosphorylation of Akt
[81].
C57Bl/6J WT mice3 mg/kg bw,
i.p.
Reduced double-stranded DNA breaks, downregulated pro-apoptotic factors caspase-3 and Bax, and upregulated anti-apoptotic factors Bcl2 and BDNF
[82].
Aged male Fisher 344 rats1.5, 3.0, 4.5, 10.0 mg/kg bw,
i.p.
Enhanced spatial memory
[83].
SAMP8 mice7.5 mg/kg bw,
i.p.
Enhanced amyloid and tau pathology, memory function, and gliosis
[84].
SAMP8 mice7.5 mg/kg bw,
i.p.
Reduced hippocampal JNK phosphorylation and tau phosphorylation, as well as ameliorated memory impairments
[85].
Sprague–Dawley male rats with aluminium-induced Alzheimer’s15 mg/kg bw,
orally
Reduced the expression of VCAM-1, TNF-α, oxidative stress markers, and α-synuclein
immunostaining, while increasing the levels of VEGF-A and nestin
[86].
Clinical studies
Alzheimer’s disease (n = 10)50 mg,
orally
Reduced the intrinsic neuronal activity in the right hippocampus
[87].
Alzheimer’s disease (n = 14)50 mg,
orally
Enhanced the cerebral metabolic rate of oxygen and cerebral blood flow (n = 12). Reduced cerebral vascular reactivity (n = 8)
[88].
Arc protein, neuronal activity-induced expression of the immediate early protein; Aβ1-40, amyloid β-protein 1-40; Aβ1-42, amyloid β-protein 1-42; Bax, B-cell lymphoma protein 2 (Bcl-2)-associated X; Bcl2, B-cell lymphoma 2; BDNF, brain-derived neurotrophic factor; bw, body weight, CDK5 p25/35, cyclin-dependent kinase 5; cGMP, cyclic, guanosine monophosphate; CREB, AMP response element-binding protein; GSK3β, glycogen synthase kinase-3 beta; HO1, heme oxygenase 1; i.p., intraperitoneally; IL-1β, interleukin 1ß; IL-6, interleukin 6; JNK, Jun N-terminal kinases; NO, nitric oxyde; NOS, nitric oxide synthase; pCREB, phosphorylated transcription factor cAMP-response element-binding protein; PKG, protein kinase G; TNF-α, tumour necrosis factor α; VCAM-1, vascular cell adhesion molecule 1; VEGF-A, vascular endothelial growth factor A.
Table 3. Clinical studies evaluating the effect of sildenafil in digital ulcers and Raynaud’s syndrome.
Table 3. Clinical studies evaluating the effect of sildenafil in digital ulcers and Raynaud’s syndrome.
Digital Ulcer
Clinical Studies
PopulationSildenafil Dosage (Oral Administration)Results/Reference
Patients with systemic sclerosis and digital
ulcers (n = 19)
114 mgReduces the quantity of digital ulcers (initial: 49, final: 17)
[95].
Females with systemic sclerosis and digital ulcers (n = 1)Sildenafil 20 mg
+
Bosentan 125 mg
Successfully heals
pre-existing digital ulcers
[96].
Patients with scleroderma (n = 10)12.5 to 100 mgDetermines the complete resolution of digital ulcerations for eight
individuals [97].
Patients with systemic sclerosis and digital
ulcers (n = 6)
50 mgFull healing (n = 2) or gradual improvement over time (n = 4)
[93].
Patients with systemic sclerosis (n = 83) and digital ulcers (n = 192)
Phase 3 study
20 mgSignificant reduction in the incidence of digital ulcers vs. placebo
[98].
Raynaud’s disease
Clinical studies
PopulationDosageResults
Symptomatic secondary Raynaud’s phenomenon (n = 16)50 mg, orallyDecreases the average number and cumulative duration of Raynaud’s attacks.
Decreases in the mean Raynaud’s Condition Score were observed. Significantly increased capillary blood flow
[93].
Raynaud’s phenomenon secondary to systemic sclerosis (n = 57)100 mg, orallyReduced the frequency of attacks among patients, while exhibiting a high level of tolerability
[94].
Raynaud’s phenomenon secondary to systemic sclerosis (n = 10)50 mg, orallyRapidly decreased the frequency and intensity of Raynaud’s phenomenon symptoms (n = 8)
[97].
Raynaud’s phenomenon secondary to systemic sclerosis (n = 123)Sildenafil 20 mg
+
Bosentan 125 mg, orally
Significantly enhanced Raynaud’s
Condition Score and improved nailfold
videocapillaroscopy
[99].
Raynaud’s disease and joint pain (n = 1)50 mg, orallyInduces vasodilation by upregulating cGMP and increasing NO synthesis
[100].
Severe Raynaud’s phenomenon and systemic scleroderma50 mg, orallySignificantly increased peripheral blood circulation and reduced the
intensity and occurrence of Raynaud’s syndrome
[101].
Primary Raynaud’ss disease (n = 15)100 mg, orallyIncreased the cutaneous vascular conductance and skin temperature [102].
Scleroderma-associated Raynaud’s phenomenon (n = 1)50 mg, orallyRaised the peripheral blood perfusion and ameliorated the manifestation of Raynaud’s syndrome symptoms
[103].
Severe Raynaud’s phenomenon and systemic sclerosis (n = 30)Sildenafil 20 mg
+
Bosentan 125 mg, orally
Reduced the intensity, frequency, and duration of Raynaud’s phenomenon
[96].
Severe Raynaud’s phenomenon associated with scleroderma (n = 1)20 mg, orallyEnhanced the blood supply of microvasculature to the extremities and effectively mitigated the vascular alterations observed in
Raynaud’s syndrome
[104].
Secondary Raynaud’s phenomenon associated with connective tissue disease (n = 10)
Phase 1 study
5 g of 5% creamSignificantly enhanced the digital artery blood flow
[105].
Raynaud’s phenomenon secondary to systemic sclerosis (n = 41)
Phase 3 study
100 mg, orallyEnhanced peripheral blood circulation and alleviation of symptoms associated with Raynaud’s phenomenon
[106].
cGMP, cyclic guanosine monophosphate; NO, nitric oxide.
Table 4. Preclinical studies evaluating the effect of sildenafil in wound healing.
Table 4. Preclinical studies evaluating the effect of sildenafil in wound healing.
Wound Healing
Preclinical Studies
PopulationDosageResults
Wistar rats with incisions of the
dorsal backs
10 mg/kg bw, orallyThe extent of skin flap necrosis was smaller in the sildenafil-treated groups than in the control group, albeit without statistical significance
[113].
Sprague–Dawley rats with incisions of the dorsal backs9 mg/kg, i.p.A notable reduction in dead tissue and blood flow stagnation within the flap was observed in rats treated with sildenafil
[114].
Wistar rats with incisions of the dorsal backs10 mg/kg bw, orallyIncreased the vascularisation of the skin flap
[115].
Streptozotocin-induced diabetic Wistar rats with incisions of the dorsal backs5% sildenafil-containing ointmentsSignificantly decreased wound area in both non-diabetic and diabetic rats, especially during the first stages of wound healing
[116].
Wistar female rats with incisions of the abdominal wall10 mg/kg, orallyEnhanced both the breaking strength of the abdominal fascia and the process of neovascularisation
[117].
Sprague–Dawley female rats with incisions of the dorsal backs0.4 g (1% gel) and 2 g (5% gel)Enhanced vascularity, reduced inflammation, granulation tissue development, and maturation in a dose-dependent manner
[118].
Sprague–Dawley male rats with incisions of the dorsal backs10% hydrogelEnhanced wound healing, reducing proinflammatory cytokine (IL-6, TNF-α, and IL-1β) levels as well as CRP levels.
Increased the levels of tissue hydroxyproline,
collagen, nitrite, and total protein content
[119].
Wistar rats with incisions of the dorsal backs10 mg/kg bw,
orally
Decreased the size of full thickness defects
[120].
Sprague–Dawley female rats with non-splinted excisions dorsal wounds3, 5 and 10% hydrogelFacilitated the process of reepithelisation, collagen synthesis, deposition, and regeneration of skin appendages
[121].
Streptozotocin-induced diabetic Sprague–Dawley rats with incisions of the dorsal backs0.7 mg/kg bw, i.p.Regulated cellular activity during the initial stages of wound healing reducing lymphocyte numbers and increasing monocytes
[122].
Streptozotocin-induced diabetic Wistar rats with incisions of the dorsal backs10 mg/kg bw,
orally
Inhibited the diabetic ulcerative process,
reduced pain perception, and promoted wound healing
[123].
Cross-breed street dogs with incisions in all skin layers on the anterior brachial region25 mgIncreased the development of granulosa tissue and capillary network; enhanced fibroblast proliferation
[109].
bw, body weight; CRP, C-reactive protein; i.p., intraperitoneally; IL-1β, interleukin 1ß; IL-6, interleukin 6; TNF-α, tumour necrosis factor α.
Table 5. Preclinical studies evaluating the benefits of sildenafil in retinopathy.
Table 5. Preclinical studies evaluating the benefits of sildenafil in retinopathy.
Retinopathy
Preclinical Studies
AnimalsAnimal ModelDosageResults
C57/B16N mouse pupsOxygen-induced retinopathy3 mg/kg bw, s.c.Reduced retinal vascular obliteration and
neovascularisation due to the stabilisation of HIF-1α expression during exposure to hyperoxia [130].
Sprague–Dawley ratsStreptozotocin-induced
diabetic retinopathy
1 and 2.5 mg/kg bw,
orally
Reduced ocular VEGF levels in diabetic rats
[124].
Wistar ratsN-nitro-L-arginine methyl
ester-induced hypertensive retinopathy
0.5 mg/kg bw, i.p.Inhibited the progression of ischaemic
injury and alterations in retinal vascular morphology
[131].
Sprague–Dawley ratsOxygen-induced retinopathy50 mg/kg bw, orallySignificantly reduced the thickness of the outer plexiform layer
[132].
bw, body weight; HIF-1α, hypoxia-inducible factor 1-alpha; s.c., subcutaneous; VEGF, vascular endothelial growth factor.
Table 6. Preclinical and clinical studies that evaluated the benefits of sildenafil in cancer.
Table 6. Preclinical and clinical studies that evaluated the benefits of sildenafil in cancer.
Cancer
In Vitro and Animal Studies
Cancer ModelDosageResults
Prostate cancer

PC-3 and DU145 prostate cancer cells
Sildenafil 10 μM
+
Doxorubicin 1.5 μM (PC-3) or
0.5 μM (DU145)
Enhanced doxorubicin-induced apoptosis by increasing ROS production, caspase-3, and caspase-9 activity and decreased Bcl-x expression and Bad phosphorylation
[25].
Breast cancer

4T1 mammary carcinoma cells
Sildenafil 10, 30, 100 μM
+
Doxorubicin 1μM
The combination of sildenafil and doxorubicin had a synergistic effect inhibiting tumour cell growth
[141].
Breast cancer

MCF-7 breast cancer cells
5, 12.5, 25, 50 μg/mLDetermined the formation of necrotic tissue formation within the tumour and had cytotoxic effects
[142].
Breast cancer

MCF-7 breast cancer cells
Sildenafil 5, 12.5, 25, 50 μg/mL
+ cisplatin 5, 12.5, 25, 50 μg/mL
Enhanced the anticancer activity of cisplatin
[142].
Neuroblastoma

Human neuroblastoma cell line IMR-32
50 μMEnhanced the development of neurite outgrowths that exhibited the expression of neuronal markers, including NeuN, NF-H, and βIII tubulin
[143].
Breast cancer

MCF-7 breast cancer cells
Sildenafil 40.33 μM
+ crizotinib 55.25 μM
The combination elicited significant apoptosis in breast cancer cells
[144].
Colorectal cancer

SW620, HT-29, HCT116, SW480, and SW1116 colorectal cancer cells
10 mM concentrationSuppressed cellular proliferation and the cell cycle and induced tumour cell apoptosis [145].
Lung cancer

A549 human lung carcinoma cells
Sildenafil 15 μM
+
Doxorubicin 1.26 μM
The combination treatment resulted in tumour cell apoptosis through the
reversal of ABC-transporter-mediated multidrug resistance and the downregulation of Nrf2 and ABCC1
[140].
Cervical cancer

HeLa, HT-3, C33A, SiHa, U14 cervical cancer cells,
and human cervical epithelial cells (HCerEpiC)
0.5, 1.0, 1.5 and 2.0 μMDecreased cell viability and
the expression of EMT marker proteins and
p-Smad2/3 in HeLa cells was suppressed
[23].
Prostate cancer

Nude mice injected with PC-3 prostate cancer cells
Sildenafil 10 mg/kg bw,
i.p.
+
Vincristine 0.5 mg/kg bw,
i.p.
Enhanced vincristine-induced mitotic arrest and increased the susceptibility of apoptosis due to mitochondrial damage
[146].
Prostate cancer

BALB/cAnNCr-nu/nu male mice injected with PC-3 cells
Sildenafil 5 mg/kg bw, i.p.
+
Doxorubicin 1.5 mg/kg bw, i.p.
or
Sildenafil 10 mg/kg bw, orally
+
Doxorubicin 3 mg/kg bw, i.p.
Significantly inhibited tumour cell proliferation
[25].
Breast cancer

Balb/c female mice injected with 4T1 mammary carcinoma cells
Sildenafil 1 mg/kg bw
i.p.
+
Doxorubicin 5 mg/kg bw,
i.v.
The association treatment decreased tumour growth compared to the use of doxorubicin alone
[141].
Breast cancer

Albino female mice injected with Ehrlich ascites carcinoma cells
5 mg/kg, orallyReduced tumour volume; decreased the levels of angiogenin, TNF-α, and the expression of vascular endothelial growth factor; and increased caspase-3 levels
[142].
Breast cancer

Albino female mice that received Ehrlich ascites carcinoma cells
Sildenafil 5 mg/kg bw, orally
+
Cisplatin 7.5 mg/kg bw, i.p.
Enhanced the anticancer activity of cisplatin [142].
Colorectal cancer

Balb/c mice injected with SW480 or HCT116 colorectal cancer cells
50, 100 mg/kg bw, orallySuppressed tumour growth
[145].
Ascites tumour

Swiss CD1 female mice injected with Ehrlich ascites carcinoma cells
1, 5 mg/kg bw, i.p.Decreased the number of tumour cells;
reduced their viability, growth rate, and ability to
proliferate; and increased apoptosis
[147].
Colorectal cancer

Dextran-sulfate sodium (DSS)-induced colitis C57BI/6J male mice
5.7 mg/kg bw,
in drinking water
The administration of sildenafil resulted in a 50% reduction in the number of colon polyps
[148].
Malignant melanoma

Mice expressing the human ret transgene in melanocytes
20 mg/kg bw,
in drinking water
Decreased the levels of IL-1β, IL-6, VEGF, S100A9, and myeloid-derived suppressor cells.
Reduced the immunosuppressive capabilities of myeloid-derived suppressor cells
[149].
Clinical studies
Type of cancer/PopulationDosageResults
Prostate cancer

Patients with ED
N/AReduced likelihood of receiving a diagnosis of prostate cancer
[150].

Lymphangioma

Phase 1, 2 study
10–20 mg, orallyReduced both the volume and symptoms of lymphatic malformation in certain paediatric patients
[151].
Breast cancer

Phase 1 study
Sildenafil 100 mg, orally
+
Doxorubicin 75–360 mg/m2,
i.v.
Did not provide cardioprotection after doxorubicin treatment
[152].
Glioblastoma
Brain cancer

Phase 2 study
Sildenafil 50 mg
+
Sorafenib 400 mg
+
Valproic acid N/A orally
Enhanced the intracranial accumulation of anticancer agents and inhibited the
proliferation of tumour cells by blocking the ABCG2 drug efflux pump within the BBB [153].
ABCC1, ATP-binding cassette sub-family C member 1; ABCG2, ATP-binding cassette, sub-family G, isoform 2 protein; Bad, BCL2 associated agonist of cell death; BBB, blood–brain barrier; Bcl-x, B-cell lymphoma-extralarge; bw, body weight; ED, erectile dysfunction; EMT, epithelial-to-mesenchymal transition; i.p., intraperitoneally; i.v., intravenous; IL-1β, interleukin 1ß; IL-6, interleukin 6; NeuN, neuronal nuclear protein; NF-H, heavy neurofilaments; Nrf2, nuclear factor erythroid 2-related factor 2; p-Smad2/3, phospho-SMAD2/SMAD3; ROS, reactive species of oxygen; S100A9, S100 calcium-binding protein A9; TNF-α, tumour necrosis factor α; VEGF, vascular endothelial growth factor A.
Table 7. Preclinical and clinical studies that evaluated the effects of sildenafil in depression.
Table 7. Preclinical and clinical studies that evaluated the effects of sildenafil in depression.
Depression
Preclinical Studies
PopulationDosageResults
Intruder-resident paradigm
conducted on CD1 mice
10 mg/kg bw, i.p.Enhances the levels of key neurotransmitters in the brain, specifically serotonin and noradrenaline, mitigating the exacerbation of depressive symptoms
[161].
Swiss male mice with lipopolysaccharide-induced depression5 mg/kg bw, i.p.Reduced the duration of immobility observed during the forced swimming test.
Increased the sucrose preference and levels of prepulse inhibition.
Increased GSH levels and decreased lipid peroxidation and IL-1β levels
[163].
Oxytocin receptor knockout mice20 mg/kg bw, i.p.Activation of oxytocin signalling pathways
[162].
Male albino Swiss mice with restraint stress-induced depressive like behaviour60 mg/kg bw, i.p.Effectively restored the immobility generated by stress in the forced swim test
[154].
Sprague–Dawley male rats with central muscarinic receptor blockadeSildenafil 10 mg/kg bw, i.p.
+
Atropine 1 mg/kg bw, i.p.
The combination has notable antidepressant-like
effects in the forced swim test. It decreases the
density of cerebral β-adrenergic receptors
[164].
Clinical studies
ED associated with mild-to-moderate depressive disorder (n = 152)25–100 mg,
orally
Sildenafil significantly improved depressive symptoms and quality of life, lowering Hamilton depression scale scores
[165].
ED associated with depression (n = 54)N/AImproved symptoms associated with depression, according to the Centre of Epidemiologic Studies—Depression Scale
[166].
ED associated with depression in patients undergoing haemodialysis (n = 16)N/AImproved BDI scores compared to baseline
[155].
ED associated with mild-to-moderate untreated depressive symptoms (n = 202)25–100 mg,
orally
Improved BDI II scores compared to the baseline measurements
[156].
ED associated with depression and idiopathic Parkinson’s disease (n = 33)50 mg, orallySignificantly reduced depression symptoms in 75% of the
patients
[160].
BDI, Beck Depression Inventory; bw, body weight; GSH, glutathione; i.p., intraperitoneally; IL-1β, interleukin 1β.
Table 8. Preclinical and clinical studies evaluating the effects of sildenafil in renal diseases.
Table 8. Preclinical and clinical studies evaluating the effects of sildenafil in renal diseases.
Renal Disease
Preclinical Studies
PopulationDosageResults
Streptozotocin-induced diabetic albino rats3 mg/kg bw, orallyEnhanced kidney function by lowering
oxidative stress; inhibiting pro-inflammatory HMGB1, TNF-α, MCP1, NF-kB, and IL1; and reducing caspase-3 levels
[174].
Ioxilan-induced acute kidney injury in New Zealand white rabbits6 mg/kg bw, orallyReduced histological injury, acute kidney injury indicators (creatinine), and electrolyte disturbances (restores K+, Na+ levels)
[178].
Iopromide-induced nephropathy in Wistar male rats10 mg/kg bw, orallyImproved structural kidney damage, exhibiting a protective potential greater than that of
N-acetyl cysteine
[179].
Streptozotocin-induced diabetes in Sprague–Dawley male rats3 mg/kg bw,
in drinking water
Rats given sildenafil had a lower kidney-to-body weight ratio.
Reduced urinary albumin excretion, renal cortical 8-OHdG levels, renal nitrotyrosine protein expression, and positive iNOS and ED-1 staining in glomeruli and tubule
interstitium
[180].
Streptozotocin-induced diabetic albino ratsSildenafil 3 mg/kg bw, orally
+
Telmisartan 10 mg/kg bw,
orally
Significantly reduced BUN, S.Cr, LDL, TGF-1, IL-1, proteinuria, and AGEPs while increasing SOD and NO
[181].
Otsuka Long-Evans Tokushima fatty male rats2.5 mg/kg bw, in drinking waterSignificantly reduced albuminuria,
glomerular hyperfiltration, glomerulosclerosis score, and the amount of nuclear antigen-positive glomerular and tubulointerstitial
proliferating cells.
Decreased collagen types I and III mRNA levels in the renal cortex
[182].
Iohexol-induced nephropathy
Wistar male rats
50 mg/kg bw, orallySignificantly reduced GFR and RBF, plasma creatinine, uraemia, and proteinuria
[183].
Iohexol-induced nephropathy
Wistar female rats
50 mg/kg bw, orallyReduced serum and renal tissue levels of HIF-2α and sCr
[184].
Streptozotocin-induced diabetic nephropathy in Sprague–Dawley rats2.5 mg/kg bw, orallyEnhanced renal function by lowering
triglyceride levels and increasing the number of podocytes
[185].
Adenine-induced chronic kidney disease in Sprague–Dawley rats0.1, 0.5, or 2.5 mg/kg bw,
orally
Reduced the plasma concentration of the
antioxidant indices in a dose-dependent manner.
Reduced the levels of indoxyl sulfate, cytokine sclerostin, and neutrophil gelatinase-
associated lipocalin
[186].
Cisplatin-induced nephrotoxicity in Wistar male rats0.4 mg/kg bw, i.p.Decreased sCr, Bax/Bcl-2 ratio, caspase 3 expression, the number of TUNEL positive cells, and N-acetyl-b-D-glucosaminidase levels.
Increased eNOS and iNOS activity and elevated renal blood flow
[187].
Cisplatin-induced nephrotoxicity in Sprague–Dawley rats2 mg/kg, i.p.Decreased BUN, sCr, MDA, and TNF-α
levels.
Increased SOD levels and nitrite/nitrate
concentrations
[188].
Cyclosporine A-induced nephrotoxicity in Wistar male rats5 mg/kg bw, orallyReduced BUN, sCr, and MDA levels.
Decreased urine albumin/Cr ratio, iNOS, TNF-α, and caspase 3 activity.
Enhanced eNOS and GSH/NO/catalase activities
[189].
Doxorubicin-induced nephrotoxicity in Sprague–Dawley rats5 mg/kg bw, orallySildenafil reduced urea, sCr, uric acid, MDA, TNF-α, and caspase-3 levels, and raised GSH levels
[190].
Ischaemia reperfusion-induced acute kidney injury in
Wistar male rats
0.5, 1.0 mg/kg bw, i.p.Increased creatinine clearance; decreased blood urea nitrogen and uric acid levels.
Inhibited the elevation in thiobarbituric acid reactive substances and superoxide anion generation, while attenuating the decrease in GSH levels through the activation of PPAR-γ receptors
[191].
5/6 nephrectomy in Wistar rats5 mg/kg, orallyEffectively mitigated single nephron hyperfiltration and hypertension, inhibited the remodelling of renal arterioles, reduced
systemic hypertension and proteinuria, enhanced the excretion of cGMP and nitrite/nitrate in urine, reduced oxidative stress, and ameliorated histological damage in the remaining kidney
[192].
5/6 nephrectomy in Wistar rats2.5 mg/kg bw, orallyReduced sCr, systolic blood pressure, and
proteinuria, while increasing urinary levels of nitric NO and cGMP
[193].
Ischaemia-reperfusion renal injury
In Sprague–Dawley rats
0.5 mg/kg bw, i.p.Enhanced the recovery of renal injury by activating ERK, inducing the synthesis of iNOS and eNOS, and reducing the ratio of Bax to Bcl-2
[170].
Alloxan-induced diabetic nephropathy in Wistar male rats3 mg/kg bw, orallyReduced blood levels of urea and creatinine and decreased urinary albumin excretion. Increased levels of cGMP, antioxidant enzymes, and testosterone
[194].
Deoxycorticosterone acetate-salt induced hypertension in Sprague–Dawley rats50 mg/kg bw, orallyReduced creatinine clearance and increased albumin-to-creatinine ratio.
Reduced glomerulosclerosis and tubulointerstitial fibrosis.
Inhibited the upregulation of ED-1, TGF-β1, and Bax, and the downregulation of Bcl-2 in the renal tissue
[195].
Streptozotocin-induced diabetes in Sprague–Dawley rats3 mg/kg bw, in drinking waterDecreased plasma concentrations of urea,
creatinine, MDA, and NO.
Increased GSH, Gpx, SOD, and CAT levels, as well as total antioxidant capacity
[175].
Partial unilateral ureteral obstruction in Wistar rats1 mg/kg bw, orallySildenafil exhibited a protective effect against tubular apoptosis
[196].
Ischaemia-reperfusion renal injury
Sprague–Dawley rats
1 mg/kg bw, orallyReduced the levels of MDA, apoptotic cells, eNOS, and p53 positive cells
[197].
Renovascular hypertension induced by the two-kidney-one-clip-operation in NO-GC1 KO mice100 mg/kg bw, orallyElevated cGMP levels, enhanced sensitivity to NO, and reduced systolic blood pressure [198].
Left renal artery clamping in
C57BL/6 mice
40 mg/kg, orallyReduced left and right kidney hypertrophy, as well as systolic blood pressure, heart rate,
and intrarenal angiotensin I/II, while increasing plasma angiotensin 1–7 and NO levels
[199].
Podocyte-specific deletion
mice with streptozotocin-induced diabetes
5 mg/kg bw, orallyReduced TRPC6 expression, glomerular
desmin, urinary albumin, and increased nephrin
[200].
Ischaemia-reperfusion renal injury
in minipigs
0.7, 1.4 mg/kg bw, i.v.Reduced systemic mean arterial pressure (1.4 mg/kg) and increased right ventricular function (0.7 mg/kg)
[201].
Cardiopulmonary bypass in
White Landrace crossbred female pigs
10 mg/kg bw, i.v.Increased renal blood flow and NO production, while reducing proteinuria, IL-18 levels, cortical expression of endothelin-1, iNOS, and inflammatory cell infiltration.
Prevented phenotypic alterations in proximal tubular cells
[202].
Right–left nephrectomy
White pigs
100 mg/kg bw, orallyIncreased right ventricular function, NO levels, and decreased right ventricular resistance
[203].
Right–left nephrectomy
minipigs
100 mg/kg bw, orallyIncreased right ventricular function, NO levels, and lowered renal vascular resistance.
Reduced tubular oedema; improved endothelial cell integrity and mitochondrial ultrastructure
[204].
Warm ischaemia in porcine kidneys1.4 mg/kg bw, i.v.Enhanced RBF and urine cGMP levels and lowered intrarenal resistance and sCr
[172].
Folic acid induced acute renal injury in New Zealand white rabbits0.3 mg/kg bw, i.p.Upregulated the expression of COX1 and Tfam at mRNA level.
Increased mtDNA copy number and downregulated KIM-1 levels
[16].
Ischaemia-reperfusion renal injury in Mongrel dogs1 mg/kg bw prior to operation,
orally
OR
0.5 mg/kg bw during the operation i.v.
Reduced sCr and BUN levels.
Decreased the activity of caspase 3, TNF-α, IL-1β, and ICAM-1.
Increased the expression of eNOS, GFR, and Nrf2
[205].
Clinical studies
PopulationDosageResults
PAH associated with impaired renal function (n = 277)20, 40, 80 mg, orallyDecreased serum creatinine and increased glomerular filtration rate
[176].
Patients with type 2 diabetes-associated microalbuminuria (n = 40)50 mg, orallySignificantly reduced albuminuria and HbA1c levels
[177].
8-OHdG, 8-hydroxy-2′–deoxyguanosine; AGEPs, advanced glycosylation end-products; Bax, B-cell lymphoma protein 2 (Bcl-2)-associated X; Bcl-2, B-cell lymphoma 2; BUN, blood urea nitrogen; bw, body weight; CAT, catalase; COX1, cyclo-oxygenase 1; eNOS, endothelial nitric oxide synthase; ERK, extracellular signal-regulated kinase; GFR, glomerular filtration rate; Gpx, glutathione peroxidase; GSH, glutathione; HbA1c, hemoglobin A1C; HIF-2α, heterodimeric nuclear transcription factor-2 alpha; HMGB1, high-mobility group box 1; i.p. intraperitoneally; i.v., intravenous; ICAM-1, intercellular adhesion molecule; IL-1, interleukin 1; IL-18, interleukin-18; IL-1ß, interleukin 1ß; iNOS, inducible nitric oxide synthase; KIM-1, kidney injury molecule-1; LDL, low-density lipoproteins; MCP1, monocyte chemoattractant protein-1; MDA, malondialdehyde; NO, nitric oxide; Nrf2, nuclear factor erythroid 2-related factor 2; RBF, renal blood flow; S.Cr, serum creatinine; SOD, superoxide dismutase; TGF-1, transforming growth factor-beta; TGF-β1, transforming growth factor-β1; TNF-α, tumour necrosis factor α; TRPC6, transient receptor potential cation channel 6.
Table 9. Preclinical and clinical studies evaluating the effects of sildenafil in gastrointestinal diseases.
Table 9. Preclinical and clinical studies evaluating the effects of sildenafil in gastrointestinal diseases.
Gastrointestinal Diseases
Preclinical Studies
Animal ModelDosageResults/Reference
Ethanol-induced gastric damage in Wistar male rats1 mg/kg bw, orallyProtected against stomach damage by activating the NO/cGMP/K(ATP) pathway [217].
Indomethacin-induced gastric mucosal damage in Wistar male rats5, 10 mg/kg bw, orallyProtected the stomach mucosa against the aggressive impact of indomethacin by increasing NO levels and inhibiting lipid peroxidation
[217].
Indomethacin-induced gastric ulcer in Sprague–Dawley female rats50 mg/kg bw, orallyProtected the stomach mucosa, mitigating indomethacin-induced damage
[218].
Acetic acid-induced gastric ulcer in Sprague–Dawley rats5, 10 mg/kg bw, orallyReduced inflammation and enhanced the healing response in the stomach mucosa
[208].
Cysteamine-induced duodenal ulcer in Wistar male rats25 mg/kg bw, orallyReduced oxidative stress and ameliorated the lesions observed in the duodenal mucosa
[219].
Indomethacin-induced gastric ulcer in Swiss male rats5, 25, 50 mg/kg bw, orallyExhibited a dose-dependent gastroprotective effect, decreasing ROS levels, enhancing NO and antioxidant enzymes levels, improving stomach cellular viability, and
restoring variables associated with gastroprotection
[220].
Indomethacin-induced gastric ulcer in Wistar male rats50 mg/kg bw, orallySignificantly decreased gastric acid
secretion, ulcer score, tissue MDA, and TNF-α levels, while increasing NO levels
[221].
Indomethacin-induced gastric ulcer in Wistar male ratsSildenafil 10 mg/kg bw,
orally + ranitidine 50 mg/kg bw,
orally
The combination exhibited antiapoptotic action on the stomach mucosa
[222].
Acetic acid-induced colitis in Sprague–Dawley rats5 mg/kg bw, s.c.Resulted in the preservation of colon
microarchitecture and a decrease in lipid
peroxidation, MPO activity, TNF-α, and IL-1β levels; increased GSH levels [216].
Trinitrobenzenesulphonic acid-induced colitis in Sprague–Dawley rats25 mg/kg bw, orallyReduced the colonic levels of MDA, MPO, CL, and TNF-α, while concurrently increasing the amount of GSH
[211].
Trinitrobenzene sulphonic acid-induced colitis in Wistar male rats25 mg/kg bw, orallyDecreased tissue levels of TNF-α, ameliorating inflammation
[223].
Trinitrobenzene sulphonic acid-induced colitis in Sprague–Dawley male rats1 mg/kg bw, i.p.Reduced the levels of indicators associated with colonic damage (TNF-α, IL-1β, MPO, and MDA)
[224].
Indomethacin-induced intestinal ulceration in Sprague–Dawley male rats3–20 mg/kg bw, orallyDose-dependently reduced the severity of the lesions.
Inhibited MPO activity, iNOS production, and bacterial invasion
[225].
ATP, adenosine triphosphate; bw, body weight; cGMP, cyclic, guanosine monophosphate; CL, lucigenin chemiluminescence; GSH, glutathione; IL-1β, interleukin-1ß; MDA, malondialdehyde; MPO, myeloperoxidase; NO, nitric oxide; TNF-α, tumour necrosis factor α.
Table 10. Preclinical and clinical studies evaluating the effects of sildenafil in cardiovascular diseases.
Table 10. Preclinical and clinical studies evaluating the effects of sildenafil in cardiovascular diseases.
Cardiovascular Diseases
Preclinical Studies
PopulationDosageResults
Adult ischaemic cardiomyocytes derived from WT mice1 μMSignificantly decreased the number of trypan blue-positive necrotic cells
[236].
C57/BL6 mice exposed to global ischaemia followed by reperfusion
(Langendorff mode)
0.1 μMIncreased the activity of Na+/K+-ATPase, promoting the reperfusion process
[237].
Ischaemic hearts of Wistar rats3 μMElevated cGMP levels and concomitantly reduced the extent of the infarct
[238].
Ischaemic isolated hearts of Wistar rats (Langendorff mode)10, 20, 50, and 200 nMEnhanced the coronary flow at lower levels of coronary perfusion pressure across all doses
[239].
Piglet model of cardiopulmonary bypass and arrest10 nMRestored ATP levels, enhanced energy charge, and modified release of hypoxanthine and inosine
[240].
C57BL6/J mice with heart hypertrophy100 mg/kg bw, orallyImproved both systolic and diastolic
function, reducing cardiac hypertrophy and cardiomyocyte apoptosis
[232].
Constriction-induced left ventricular pressure overload in
C57BL/6J male mice
200 mg/kg bw,
in soft chow
Inhibited ERK and calcineurin activity in both the right and left ventricles
[241].
Transaortic constriction–induced left ventricular pressure overload in PKGIa LZM mice200 mg/kg bw,
in food
Significant inhibition of cardiac hypertrophy and left ventricular systolic dysfunction
[242].
WT mice exposed to left atrial incision (Langendorff mode)0.71 mg/kg bw, i.p.Decreased the extent of myocardial infarction
following an episode of ischaemia
[236].
ICR mice exposed to global ischaemia followed by reperfusion0.71 mg/kg bw, i.pDecreased the extent of the infarct through the
activation of mitoKCa and mitoKATP
[243].
ICR mice exposed to ischaemia followed by reperfusion0.7 mg/kg bw, i.p.Reduced the extent of the infarct, increasing both iNOS and eNOS levels
[244].
ICR mice exposed to descending coronary artery occlusion followed by reperfusion0.7 mg/kg bw,
i.p.
Elevated cardiac SIRT1 activity, leading to a
reduction in the extent of the myocardial infarction
[245].
ICR mice exposed to descending coronary artery occlusion21 mg/kg bw, i.p.Ensured the preservation of fractional shortening. Reduced the left ventricular end-diastolic dilatation, fibrosis, and apoptosis
[246].
ICR mice subjected to myocardial infarction through the closure of the left anterior descending coronary artery0.71 mg/kg bw, i.p.Reduced ischaemic injury score.
Increased the expression of eNOS and iNOS
proteins and the ratio of Bcl-2 to Bax.
Reduced apoptosis and the left ventricular
end-diastolic diameter
[247].
Dystrophin-deficient mice80 mg/kg bw, orallyMitigated impairments in heart function.
Improved both the myocardial
performance index and the ratio of early diastolic
filling velocity to late diastolic filling velocity
[248].
iNOS knockout and eNOS knockout C57BL6/J mice exposed to ischaemia followed by reperfusion0.06 mg/kg bw
injection into the LV lumen
The administration of sildenafil resulted in a notable decrease in the extent of myocardial infarction. This indicates that the acute
cardioprotective effects of low-dose sildenafil are not reliant on eNOS, iNOS, or cGMP
[249].
PGC1α−/−mice subjected to transverse aortic constriction-induced pressure overload200 mg/kg bw, orallyEnhanced cardiac function and remodelling.
Improved mitochondrial respiration and increased the expression of PGC1α mRNA in the myocardium
[250].
Sprague–Dawley rats exposed to ischaemia followed by reperfusion0.75 mg/kg bw,
i.p.
Significantly reduced the size of infarcted regions [251].
Wistar rats exposed to heterotopic cardiac transplantation0.7 mg/kg bw,
i.v.
Enhanced both systolic and diastolic
function of the myocardium following a three-hour
period of arrest.
Determined a notable translocation of CPK delta
[252].
Wistar rats exposed to ischaemia and reperfusion50 mg/kg bw,
orally
Suppressed the significant elevation of MDA
levels
[253].
Sprague–Dawley rats exposed to left anterior descending coronary artery occlusion followed by reperfusion0.7 mg/kg bw,
i.v.
Increased the density of capillaries and arterioles, enhancing blood flow.
Increased the expression of VEGF and Ang-1 at mRNA levels during the initial reperfusion period
[254].
Mongrel dogs exposed to blockage of the anterior descending coronary artery2 mg/kg bw, orallyProlonged the QT interval, particularly in the presence of ischaemic conditions
[233].
Piglets exposed to untreated ventricular fibrillation, which was afterwards followed by open-chest cardiopulmonary resuscitation0.5 mg/kg, i.p.The administration of sildenafil resulted in a shift towards aerobic metabolism in energy use.
Increased the levels of ATP and ADP, while decreasing the levels of lactate in the myocardial tissue
[255].
Pigs exposed to ventricular fibrillation and cardiopulmonary resuscitation0.5 mg/kg bw, i.p.Partially reduced the elevated levels of plasma Ang II and Ang (1–7).
Increased the expression of eNOS, cGMP, and iNOS
[256].
Cardiac arrest pigs0.5 mg/kg bw, i.p.Reduced the expression levels of miR-155-5p and miR-145-5p
[257].
Fixed banding of the venous pulmonary confluent-induced postcapillary PH pigs25–50 mg/kg bw, orallyReduced apoptotic cells.
Regulated gene expression decreasing oxidative stress and enhancing anti-inflammatory activity within the myocardium.
Improved ventricular function
[258].
New Zealand rabbits with left anterior descending artery occlusion0.7 mg/kg bw,
i.p.
Elicited both immediate and delayed protective effects against ischaemia-reperfusion injury by the activation of mitochondrial KATP channels
[259].
New Zealand rabbits exposed to ischaemia through the blockage of coronary arteries.0.71 mg/kg bw, i.v.Reduced the size of the infarct
[260].
Clinical studies
Patients with chronic heart failure (n = 46)50 mg, orallyDecreased pulmonary artery pressure and alleviated dyspnoea, while improving brachial artery flow-mediated dilatation and enhancing breathing during physical exertion
[234].
Patients with coronary artery disease (n = 24)100 mg, orallyInduced dilation of pericardial arteries while interfering with platelet activation
[261].
Patients with left ventricular failure (n = 100)50 mg, orallyEnhanced left ventricular ejection fraction, increased performance on the 6 min walking test, and positively impacted Doppler-derived variables related to left ventricular diastolic function
[235].
Patients with left ventricular systolic dysfunction (NYHA II-IV) (n = 34)25–75 mg, orallyEnhanced exercise capacity, decreasing pulmonary
vascular resistance and increasing the cardiac output during exercise
[262].
Patients with non-ischaemic diabetic cardiomyopathy (n = 59)100 mg,
orally
The administration of sildenafil was linked to an anti-remodelling impact, which subsequently led to enhanced cardiac kinetics
[263].
Patients with heart failure and preserved ejection fraction (n = 216)20, 60 mg, orallyThere was no change in exercise capacity, clinical
condition, or quality of life
[264].
Patients with coronary artery disease (n = 144)100 mg, orallyReduced the time elapsed until the onset of angina and the overall duration of activity
[265].
Patients with microvascular
coronary dysfunction (n = 23)
100 mg, orallyEnhanced the coronary flow reserve
[266].
ADP, adenosine diphosphate; ANG (1–7), angiotensin 1–7; ANG II, angiotensin II; ANG-1, angiotensin 1; ATP, adenosine triphosphate; Bax, B-cell lymphoma protein 2 (Bcl-2)-associated X; Bcl-2, B-cell lymphoma 2; bw, body weight; cGMP, cyclic, guanosine monophosphate; CPK, protein kinase C; eNOS, endothelial nitric oxide synthase; i.p., intraperitoneally; iNOS, inducible nitric oxide synthase; i.v., intravenous; KATP channel, ATP-sensitive potassium channel; MDA, malondialdehyde; miR-145-5p, microRNA-145-5p; miR-155-5p, microRNA-155-5p; mitoKATP, mitochondrial ATP-sensitive K+; mitoKCa, mitochondrial Ca2+-activated K+; mRNA, messenger RNA; PGC1α, Peroxisome proliferator-activated receptor-γ coactivator 1-α; RNA, ribonucleic acid; VEGF, vascular endothelial growth factor.
Table 11. Preclinical and clinical studies evaluating the effects of sildenafil in lung diseases.
Table 11. Preclinical and clinical studies evaluating the effects of sildenafil in lung diseases.
Lung Diseases
Preclinical Studies
PopulationDosageResults
Ovalbumin-sensitised BP2 mice
(asthma model)
Sildenafil 20 mg/kg bw,
+
L-arginine 50 mg/kg bw,
i.p.
Exacerbated airway hyper-responsiveness
[268].
Wistar rat pups with neonatal hyperoxic lung injury50 mg/kg bw, orallyExtended survival, elevated levels of pulmonary cGMP, diminished the pulmonary
inflammatory response, decreased fibrin deposition and right ventricular hypertrophy, and promoted alveolarisation
[269].
Neonatal Wistar rats exposed to hypoxia to induce bronchopulmonary dysplasia50, 100 mg/kg bw, orallyFacilitated the restoration of lung function by stimulating the activation of HIF-α and promoting the
overexpression of VEGF
[270].
Acrolein-induced airway inflammation in Sprague–Dawley rats25 mg/kg bw, orallyReduced the production of TNF-α, inhibited leukocyte migration, and decreased mucus hypersecretion.
Prevented epithelial hyperplasia and metaplasia
[271].
Bleomycin-induced lung fibrosis Sprague–Dawley rats10 mg/kg bw, s.c.Reduced lung fibrosis by decreasing the expression of TNF-α and IL-1β
[215].
Meconium-induced acute lung
injury in newborn Wistar rats
25 mg/kg bw, orallyPreserved lung tissue and mitigated the inflammatory burst comparable to dexamethasone
[272].
Repetitive lung gavage saline induced acute lung injury rabbits1 mg/kg bw, i.p.Decreased the migration of cells, specifically
neutrophils.
Reduced the release of TNF-α, IL-8, and IL-6, and decreased the levels of nitrite/nitrate,
3-nitrotyrosine, and MDA.
Prevented lung oedema development and reduced protein content in BAL and death of epithelial cells
[273].
Clinical studies
Patients with cystic fibrosis (n = 19)
Phase 2 study
20, 50 mg orallyEnhanced exercise capacity in individuals diagnosed with cystic fibrosis
[274].
Patients with mild-to-moderate cystic fibrosis (n = 36)
Phase 1, 2 study
20, 40 mg, orallyDecreased sputum elastase activity. Well-tolerated
[275].
Patients with idiopathic pulmonary fibrosis (n = 29)

Phase 2 study
20 mg, orallyThe administration of sildenafil did not yield a statistically significant increase in the distance
covered during the 6 min walk test, nor resulted in a significant decrease in the Borg dyspnoea index
[276].
Patients with swimming-induced pulmonary oedema (n = 10)50 mg, orallyDecreased the pulmonary vascular pressures without any negative impact on exercise
hemodynamics
[277].
Patients with pulmonary embolism (n = 20)
Early phase 1 study
50 mg, orallyThe administration of one oral dose of sildenafil did not provide any significant enhancement in cardiac index, but reduced systemic blood pressure
[278].
Children with mild-to-moderate lung disease (n = 20)1 mg/kg bw, orallyEnhanced the overall well-being and physical capabilities of children with cystic fibrosis, but also had a notable negative impact on lung function
[279].
Patients with idiopathic pulmonary fibrosis and PAH (n = 14)
Phase 2 study
20–50 mg, orallyEnhanced the 6 min walk distance
[280].
Patients with advanced idiopathic pulmonary fibrosis (n = 247)
Phase 2 study
Sildenafil 20 mg
+
Pirfenidone 801 mg,
orally
The co-administration of sildenafil with pirfenidone did not yield any therapeutic advantage
[281].
Patients with idiopathic pulmonary fibrosis (n = 274)
Phase 3 study
Sildenafil 20 mg
+
Nintedanib 150 mg, orally
The combination of nintedanib and sildenafil did not yield a statistically significant advantage when compared to the administration of nintedanib alone
[282].
BAL, bronchoalveolar lavage fluid; bw, body weight; cGMP, cyclic, guanosine monophosphate; i.p., intravenous; IL-1β, interleukin 1β; IL-6, interleukin 6; IL-8, interleukin 8; MDA, malondialdehyde; s.c., subcutaneous; TNF-α, tumour necrosis factor α; VEGF, vascular endothelial growth factor.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pușcașu, C.; Zanfirescu, A.; Negreș, S.; Șeremet, O.C. Exploring the Multifaceted Potential of Sildenafil in Medicine. Medicina 2023, 59, 2190. https://doi.org/10.3390/medicina59122190

AMA Style

Pușcașu C, Zanfirescu A, Negreș S, Șeremet OC. Exploring the Multifaceted Potential of Sildenafil in Medicine. Medicina. 2023; 59(12):2190. https://doi.org/10.3390/medicina59122190

Chicago/Turabian Style

Pușcașu, Ciprian, Anca Zanfirescu, Simona Negreș, and Oana Cristina Șeremet. 2023. "Exploring the Multifaceted Potential of Sildenafil in Medicine" Medicina 59, no. 12: 2190. https://doi.org/10.3390/medicina59122190

Article Metrics

Back to TopTop