Next Article in Journal
Helicate versus Mesocate in Quadruple-Stranded Lanthanide Cages: A Computational Insight
Next Article in Special Issue
Zrsr2 Is Essential for the Embryonic Development and Splicing of Minor Introns in RNA and Protein Processing Genes in Zebrafish
Previous Article in Journal
Mesenchymal Stem Cells and Formyl Peptide Receptor 2 Activity in Hyperoxia-Induced Lung Injury in Newborn Mice
Previous Article in Special Issue
In Vitro Characterization of the Physical Interactions between the Long Noncoding RNA TERRA and the Telomeric Proteins TRF1 and TRF2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Complexity and Multiplicity of the Specific cAMP Phosphodiesterase Family: PDE4, Open New Adapted Therapeutic Approaches

Section de Structures Biologiques, Pharmacologie et Enzymologie, CNRS/Unistra, CRBS, UR 3072, CEDEX, 67084 Strasbourg, France
Int. J. Mol. Sci. 2022, 23(18), 10616; https://doi.org/10.3390/ijms231810616
Submission received: 9 August 2022 / Revised: 4 September 2022 / Accepted: 7 September 2022 / Published: 13 September 2022
(This article belongs to the Collection Feature Papers in “Molecular Biology”)

Abstract

:
Cyclic nucleotides (cAMP, cGMP) play a major role in normal and pathologic signaling. Beyond receptors, cyclic nucleotide phosphodiesterases; (PDEs) rapidly convert the cyclic nucleotide in its respective 5′-nucleotide to control intracellular cAMP and/or cGMP levels to maintain a normal physiological state. However, in many pathologies, dysregulations of various PDEs (PDE1-PDE11) contribute mainly to organs and tissue failures related to uncontrolled phosphorylation cascade. Among these, PDE4 represents the greatest family, since it is constituted by 4 genes with multiple variants differently distributed at tissue, cellular and subcellular levels, allowing different fine-tuned regulations. Since the 1980s, pharmaceutical companies have developed PDE4 inhibitors (PDE4-I) to overcome cardiovascular diseases. Since, they have encountered many undesired problems, (emesis), they focused their research on other PDEs. Today, increases in the knowledge of complex PDE4 regulations in various tissues and pathologies, and the evolution in drug design, resulted in a renewal of PDE4-I development. The present review describes the recent PDE4-I development targeting cardiovascular diseases, obesity, diabetes, ulcerative colitis, and Crohn’s disease, malignancies, fatty liver disease, osteoporosis, depression, as well as COVID-19. Today, the direct therapeutic approach of PDE4 is extended by developing allosteric inhibitors and protein/protein interactions allowing to act on the PDE interactome.

1. Introduction

3′,5′- cyclic AMP (cAMP) plays a major role in the physiological control as well as in the pathophysiological control of tissue and cellular responses [1] (Figure 1) [2]. Its intracellular level depends:
-
Firstly, on its synthesis by particulate adenylyl cyclases (pAC) governed by membrane receptors (for review: [3] or either by a soluble AC (sAC), [4];
-
Secondly, its rapid degradation in 5′-AMP, by cyclic nucleotide phosphodiesterases (PDEs), allows the rapid restoration of the initial state of cellular cAMP level. Nevertheless, but interestingly, PDEs dysregulation plays a major role in many known, as well as unsolved, pathologies [5,6,7,8,9,10].
The PDEs represent a multigene superfamily, selectively hydrolyzing cAMP and /or cGMP (Figure 1). They are composed of 11 families representing more than 100 human variants. Among these, the biggest is the PDE4 family. It is constituted by four genes, giving rise to many proteins distributed differently cellularly and subcellularly, pointing out the diversity and the therapeutic potential of the PDE4 family [11].

1.1. The PDE4 Family

The PDE4 was first described as a PDE hydrolyzing selectively cAMP and thereafter was characterized by its specific sensitivity to the antidepressant compound rolipram (ZK 62711) and consequently named rolipram inhibited-cAMP-PDE (ROI-cAMP-PDE). Thus, differentiating it from the specific cAMP-PDE isolated from human platelets, named cGI-PDE, since this cAMP-PDE is inhibited by cGMP [12,13]. In another way, it was shown that rolipram is an antidepressant agent that binds with high affinity to the rat brain, opening the link between the high-affinity rolipram-binding site and the presence of PDE4 in the brain [14].
Consequently, together with Prof. C.G. Wermuth and Dr. J.J. Bourguignon, new PDE4 inhibitors [15] were conceived based on rolipram and trequinsin chemical structures opening a new field of research by designing R1 and R2 adenine nucleotides (Figure 2) [13,15].
Since PDE4 inhibitors act on inflammatory processes, mainly on oxygen-derived free radicals [16], and that moreover some failures (unsolved mortality) were obtained during the PDE3 inhibitor development, PDE4 research was preferentially developed. In another way, the cloning and expression of cDNA for a human low-Km, rolipram-sensitive cAMP phosphodiesterase was firstly reported in 1990 [17] as well as the structure of two rat genes coding for rolipram-sensitive cAMP-PDE [18]. Thus, pharmaceutical companies have conducted many studies to design and develop increasingly potent and specific PDE4 inhibitors (for review see: Houslay et al., 2005 [19]. Altogether, knowledge of the structures, properties, and functions of the PDE4 family has increased drastically.

PDE4 Structure

The PDE4 family, which exclusively hydrolyses the cAMP (Km = 2–4 µM), is the most extensive family, coded by four genes, namely PDE4A to PDE4D. Each gene encodes multiple protein products generated by alternative splicing and/or by the use of multiple promoters (see Figure 3). More than 25 human isoforms of PDE4 are identified. The monomer of these proteins (from 50 to 125 kDa) is constituted by: (1°) a unique amino acid signature region called upstream conserved regions 1 and 2 (UCR1 and UCR2) which allows to classify PDE4 into long and short forms. The long PDE4 isozymes exhibit both UCR1 and UCR2, whereas short PDE4 isozymes lack UCR1 [20] (Figure 3). Since the UCR module provides dimerization, only long-form PDE4 splice variants containing UCR1 and UCR2 are dimerized [21]. UCR1 contains a PKA phosphorylation site that upon phosphorylation attenuates the ability of UCR1 to interact with UCR2 and thereby increases PDE4 activity; (2°) the catalytic domain, which constitutes the core of PDE, is a highly conserved region, with a 20–45% identity. It includes consensus metal-binding domains: two Zn2+-binding motifs (Hx3HxnE/D) and an Mg2+-binding motif, related to metal-ion phosphohydrolases [22], which are critical for the catalytic activity; (3°) the carboxyl-terminal end of the catalytic region contains an ERK-phosphorylation site. Its phosphorylation induces the activation of short forms of PDE4 (without UCR2) and inhibition of long forms of PDE4 [23] (Figure 3).
Rolipram, the prototype PDE4 inhibitor, binds to PDE4 at two sites, called the low-affinity rolipram binding site (LARBS) and the high-affinity binding site (HARBS) [12,24,25,26,27,28] (Figure 4).
Interestingly, it has been shown that pregnancy induces a modulation of the PDE4-conformers ratio in human myometrium, designing LPDE4 as potential pharmacological targets for the design of new tocolytic treatments [29]. In contrast, HARBS mediates the antidepressant-like effects of PDE4 inhibitors [30]. Today, it seems that not enough people are interested in the biophysical concern of this two-site concept, but they merely suggest changes in spatial conformation inducing changes in cAMP hydrolytic activities.

2. PDE4 as Therapeutic Targets

Rolipram and Ro 20–1724 acting in the µmolar range were first identified as highly selective competitive inhibitors of PDE4 [1]. Subsequently, many specific PDE4 inhibitors acting in the nanomolar range were conceived as anti-inflammatory agents for asthma and COPD, as they reduced oxidative stress, TNF-α, and cytokine production. Roflumilast was the first PDE4 inhibitor developed on the market (called Daxas®) to treat COPD [31]. Lastly, apremilast (Otezla®), an oral PDE4 inhibitor, has received FDA approval to treat patients with active psoriatic arthritis and plaque psoriasis [32]. It is interesting to note that a different type of PDE4 inhibitor, classified as an allosteric inhibitor, has been developed opening up new perspectives. By closing UCR2 across one active site in the PDE4D dimer and by decreasing the catalytic activity of the second active site, it consequently reduces emetic potential [33]. Thus, the allosteric inhibitor of PDE4D compound BPN14770 is presently under development for enhancing memory. This compound acts newly on two different rolipram binding sites [34], re-opening questions about the concept of two affinity binding sites.
Nowadays, there are different approaches under development targeting PDE4–protein interactions [35] allowing to locally and specifically alter PDE activity in the cell compartment of interest, named the signalosome, without acting on the whole cell or tissue. There may also be signalosome disruptors that affect interactions between PDEs and their regulatory partners [6,7]. In another way, the dissociation of the β1-adrenergic-receptor/PDE4 complex was reported as a novel mechanism of antagonist action [36]. Lastly, the identification of a PDE4-specific pocket design of selective inhibitors was reported that might improve the selectivity of PDE4 inhibitors [37].
Consequently, there are multiple pharmacological means to interact at the level of PDE4 changes inducing new specific and localized therapeutic approaches allowing less undesired effects.

2.1. PDE4 in Cardiovascular Diseases

PDE4 participates in the control of cardiovascular function at the level of vessels and heart.

2.1.1. PDE4, and Vessel

PDE4 has been characterized in human, rat, and bovine aortas [12]. It is clearly established that in vascular smooth muscle, PDE4 inhibitors control endothelium-dependent relaxation, opposite to PDE3 inhibitors or nitric oxide (NO) that per se relax the vessels [38,39,40]. This is related to the different kinetic properties of PDE3 and PDE4 (PDE3 Km < PDE4 Km, and PDE3 Vmax/Km > PDE4 Vmax/Km) as illustrated in Figure 5. In the presence of endothelium, at low intracellular cAMP concentrations, firstly PDE3 hydrolyses cAMP; therefore, its specific inhibition, or the endothelial NO production, induces an increase in the cAMP level thus allowing PDE4 hydrolytic function. In that condition, the inhibition of PDE4 would be effective by increasing the cAMP level and would therefore relax the whole vessel. Altogether, at the level of vascular smooth muscle, we demonstrated the existence of cAMP/cGMP cross-talk mediated by PDE3 and PDE4 [40,41,42].
Moreover, it should be noted that PDE4 is mainly present in human endothelial cells [43,44,45,46]. Surprisingly, in the presence of L-arginine, the PDE4 inhibitor, rolipram, is able to increase cGMP levels by upregulating the L-arginine/NO/cGMP pathway, revealing a cross-talk between PDE4 and cGMP regulation [47] (Figure 6).
In particular, PDE4 plays a role in the regulation of endothelial permeability [48]. Notably the PDE4D, by tethering the exchange protein activated by cAMP (EPAC)1 in a vascular endothelial cadherin-based signaling complex (VE-Cad), controls the cAMP-mediated vascular permeability [49]. It should be noted that heart failure in rat induces increases in the PDE4B protein and mRNA expressions in the aorta which could contribute to high blood pressure [50]. Lastly, it was shown that PDE4B, and not PDE4D, was upregulated in inflammatory cells from both experimental and human abdominal aortic aneurism, limiting the progressive increase in the aortic diameter without affecting the blood pressure. Rolipram has strongly mitigated the increase in vascular oxidative stress (superoxide anion) induced by angiotensin II [51]. Clinically, it is important to note that PDE4 inhibition reduces life-sustaining extracorporeal vascular permeability and improves microcirculation in an extracorporeal resuscitation rodent model [52].

2.1.2. PDE4, and Heart

The cAMP in cardiac muscle is the second intracellular messenger mediating the positive inotrope effects of β-agonists [53]. Thus, PDE3 inhibitors, such as milrinone, SK, and F 94120, were described early in the early 1980s as “new cardiotonic drugs” [54,55]. Consequently, it was clearly established that the positive inotrope effect induced by PDE3 inhibitors was mediated by cAMP [56]. However, due to some mortality in chronic treatment related to tachyarrhythmia and tachycardia, milrinone prescription in humans is now only performed for a short period, beneficially inducing reduced inflammatory and apoptotic signaling [57]. At the same time, a rolipram-sensitive cAMP-PDE has been characterized in the heart. This was carried out by anion exchange chromatography performed on cardiac tissues from the rat heart ventricle [58], as well as from canine left ventricles [59] and from frog atrial fibers [60]. Furthermore, Komas et al., 1989 [55] isolated by chromatography the rolipram-inhibited cAMP-PDE (PDE4) from dog ventricular and sino-atrial node tissues, pointing out different inhibitory effects of specific inhibitors between the ventricle and sino-atrial node PDEs. Thus, the characterization of PDE4 in cardiac tissues raised the question of its implication in cardiac contraction.
Studies performed on guinea pig cardiac ventricle showed that PDE4 might be implicated differently than PDE3 in the regulation of cardiac contraction, since in opposite to PDE3 inhibitor, PDE4 inhibitor did not per se increase cardiac inotropy, although PDE3 and PDE4 are both cytosolic and membrane bound [61,62]. However, interestingly, PDE4 inhibitors potentiate the effect of PDE3 inhibition, showing that in the presence of cyclic AMP-dependent positive inotropic agents, PDE4 inhibitors exert a positive inotropic effect which probably does not involve enhanced catecholamine release from sympathetic nerve endings [62]. This was confirmed more recently by Eschenhagen, 2013 [63] indicating that rolipram does not affect the positive inotropic effects from β1- or β2-adrenoceptor stimulation. This might also suggest that PDE4 acts in another cAMP compartment.
Thus, we firstly showed in canine and human, purified cardiac microsomal fractions, that PDE3 is associated with the microsomal membranes enriched in vesicles derived from T-tubule and junctional SR membranes; meanwhile, PDE4 is mostly associated with the enriched fraction of sarcolemma membranes, revealing different subcellular compartmentations for PDE3 and PDE4 [64]. To go further, we have canonically demonstrated that PDE4B and PDE4D are associated with the envelope of the nucleus isolated from human cardiac cells [65]. This has only been confirmed very recently by a pharmacological approach, demonstrating that PDE4 insulates a mAKAPβ-targeted PKA pool at the nuclear envelope [66]. In that way, Marco Conti and his colleagues showed that PDE4D isoforms are anchored by myomegalin colocalizing components of the cAMP-dependent pathway to the Golgi/centrosomal region of the cardiac cell [67]. Due to the different physiological and localized contributions of PDE4 in cardiac contraction, its physiopathological role in cardiac pathology could therefore be questioned.
A study using PDE4D KO mice, clearly demonstrated that PDE4D deficiencies in the ryanodine receptor (RyR) complex promote heart failure and arrhythmias [68]. Furthermore, during cardiac hypertrophy in rats, there is a decrease in PDE4A, and PDE4B proteins and mRNA expressions, that accompany β-adrenergic desensitization, indicating some failures of PDE4s in cardiac pathologies [69]. However, it must be noticed that there is a five-fold higher amount of non-PDE4 activity in human hearts compared to rodents. As a result, only a restricted number of proteins could be phosphorylated in human cardiomyocytes, which would not affect overall cAMP signaling in the human heart [70].
PDE4D represents the main PDE4 isoform in the human heart and its expression is decreased in idiopathic dilated cardiomyopathy as well as in rats during early angiotensin II-induced cardiac hypertrophy, although total cAMP-PDE activity was increased [71]. In this way, PDE4D regulates the release of baseline sarcoplasmic reticulum Ca2+ and cardiac contractility, independently of L-type Ca2+ current, allowing for precise pharmacological control of cardiac contractility [72]. This could have a useful therapeutic implication for patients, since in failing human hearts, PDE4D co-immunoprecipitated with sarcoplasmic reticular (SR) calcium-ATPase type 2a but not with RyR2 [72]. Surprisingly, although PDE4D is part of a Cav1.2 signaling complex, the inactivation of the pde4d gene had no consequence on basal or β-AR stimulated ICa,L. In another way, studies performed in KO mice reveal that PDE4B plays also a key role during β-AR stimulation of cardiac function. As a result, PDE4B, by limiting the amount of Ca2+ entering the cell through the L-type Ca2+ channels, prevents Ca2+ overload and arrhythmias [73].
Altogether, decreases in PDE4D activity, which represents the main PDE4 isoform in human cardiac tissue, might favor idiopathic dilated cardiomyopathy and its deficiency in the RyR complex promotes heart failure and arrhythmias, whereas PDE4B prevents Ca2+ overload and arrhythmias. Interestingly, UCR1C is a new, long isoform PDE4 activator that alleviates cardiomyocyte hypertrophy [74]. Lastly, it was reported that decreases in PDE4 expression in the heart, related to fibroblast growth factor 23, increase the risk of cardiac arrhythmia [75].

2.2. PDE4 in Obesity

Obesity is considered the fifth highest cause of death worldwide by the World Health Organization and is associated with pathological conditions and diseases associated with obesity, including hyperlipidemia, heart diseases such as coronary artery disease (CAD) and myocardial infarction, stroke, type 2 diabetes (T2D), hypertension, cancers, low-grade and chronic inflammation, fatty liver disease, osteoarthritis, respiratory problems, and neurodegenerative diseases [76].
Obesity is defined by the body mass index (BMI). It is calculated as weight in kilograms divided by height in meters squared, rounded to one decimal place. Obesity in adults (>20 years old) was defined as a BMI greater than or equal to 30. The prevalence of obesity was 39.8% and affected about 93.3 million US adults in 2015~2016 [77].
Therefore, it was pointed out that caloric restriction might extend lifespan and that the metabolic effect of resveratrol, a mimetic of caloric restriction, might be mediated by inhibiting cAMP phosphodiesterases [78]. In that way, alterations in PDE activities were firstly reported in omental and subcutaneous adipose tissues in human obesity, i.e., in omental and subcutaneous adipose tissues, it was a significant negative correlation between PDE4 and BMI [79].
Thus, the involvement of PDE4 in obesity has been demonstrated since, under submaximal β-adrenoceptor stimulation of brown adipocytes, the PDE4 inhibitor alone may increase lipolysis [80]. Therefore, it was reported that the PDE4 inhibitor, roflumilast, might be superior to metformin in weight loss in obese women with polycystic ovarian fibrosis [81]. In this way, PDE4B-KO mice had reduced adiposity and adipose-induced inflammation induced by a high-fat diet, highlighting the involvement of PDE4B in adiposity [82].
A study of the Saudi population found that PDE4D polymorphism (rs295978) was not associated with a risk of obesity and did not affect the blood lipid profile [83]. Nevertheless, the allosteric PDE4D inhibitor, compound D159687, induces weight loss and fat mass loss in older mice without changing lean mass and physical and cognitive function [84]. Thus, we proposed that PDEs and in particular PDE4 participate in the metabolic syndrome by regulating the AMP-activated protein kinase (AMPK) [85]. In this way, it is noteworthy that the effect of roflumilast depended on the activation of the metabolic regulator AMPKα. Consequently, suppressed adipogenesis and promoted lipolysis in cell culture and mouse models suggest the therapeutic potential of roflumilast in obesity-related diseases [86].
Altogether, PDE4 regulation seems to be implicated in obesity at the level of PDE4B and PDE4D, and an allosteric inhibitor of PDE4D would be effective in the loss of fat mass without changing cognitive function.

2.3. PDE4 in Diabetes

The global prevalence of diabetes among 20- to 79-year-olds is estimated to be about 8.8% in 2015, or 415 million, and approximately 10.4% in 2040, or 642 million [87]. Diabetes is related to alterations in glucose homeostasis governed by insulin. The failure to maintain glucose homeostasis underlies both type 1 diabetes (T1D) and type 2 diabetes (T2D). T1D is an autoimmune disease that originates when β-cells that produce insulin are destroyed. T2D is characterized by insulin resistance and the progressive loss of β-cell function. Although these two forms of diabetes are fundamentally very different, β-cell failure and death play a key role in the pathogenesis of both diseases, leading to hyperglycemia resulting from a reduced capacity to produce insulin [88]. The diagnosis of diabetes and its severity may be determined by measuring blood glycated proteins such as hemoglobin A1c (HbA1c) and glycated albumin (GA) [89]. In the 2000s, it was proposed that inhibiting PDE4 during diabetes may be beneficial against hyperglycemia, oxidative stress, and the production of TNF-α and NFκB [90]. In this way, it has previously been reported that insulin-secreting cells and Langerhans islets contain PDE4 [91]. PDE4C is the main PDE4 subtype expressed at the mRNA level in isolated rat islets. Interestingly, silencing the PDE4C as well as the specific PDE4 inhibitors, roflumilast and compound L-826,141, significantly increases glucose-dependent insulin secretion [92]. PDE4 seemed active only after stimulation with glucose, suggesting some interaction between glucose and PDE4 [93]. Following this, a study in mice showed that roflumilast improves glucose tolerance and insulin sensitivity [93].
Recently, TAK-648, a PDE4 inhibitor, with demonstrated preclinical anti-diabetic properties, has been developed for human T2D (dose prediction using HbA1c results from a db/db mouse study), opening the way for new therapeutic approaches in diabetes [94]. Interestingly, a recent review on PDE4 inhibitors in diabetic nephropathy highlights this project [95] as well as a review on PDE4 inhibitors in prediabetic patients [96].

2.4. PDE4 and Ulcerative Colitis and Crohn’Disease

Crohn’s disease is a chronic inflammatory condition of the gastrointestinal tract. It can affect any part of the gastrointestinal (GI) tract, but ulcerative colitis affects only the colon. Additionally, while Crohn’s disease can affect all layers of the bowel wall, ulcerative colitis (UC) only affects the lining of the colon. Both have been together classified as inflammatory bowel diseases (IBD). In the United States, it is currently estimated that about 1.5 million people suffer from IBD, causing considerable suffering, with a prevalence of Crohn’s disease (CD) of 201 per 100,000 population, UC is being equally prevalent [97].
Because Crohn’s disease is mainly a chronic inflammation, the inflammatory cytokine productions in intestinal biopsies have been studied in relation to the pathological grade. This study originally shows that TNF-α and IL-1β are significantly increased in endoscopic biopsies, as well as for TNF-α, IL-1β, and IL-6 in the individual culture supernatant of intestinal biopsies, opening a new way of investigation [98]. Thus, the effect of pentoxifylline was investigated on intestinal inflammation in IBD, showing that pentoxifylline downregulates in vitro TNF-α and IL-1β production by PBMCs and by intestinal organ cultures from patients with Crohn’s disease and ulcerative colitis [99]. Because pentoxifylline, acting on TNF-α and IL-1β, is a non-selective PDE inhibitor [100], PDE activities were investigated in the human normal mucosa and inflamed mucosa of patients with Crohn’s disease. These studies revealed an increase in the % of PDE4 activity (from 42% to 72.5 %, up by p < 0.05%) suggesting that the specific PDE4 inhibitor may be effective in Crohn’s disease [101]. In that way, the use of OPC-6535 (tetomilast) was suggested for the treatment of a variety of oxidative inflammatory intestinal disorders with an abnormal mucosal barrier such as inflammatory bowel disease [102]. As a result, tetomilast was developed by OTSUKA (phase III) to investigate the treatment of UC [103]. Furthermore, it has been clearly reported that the PDE4 inhibitor, rolipram, prevents and reduces experimental colitis in mice and suppresses TNF-α levels in colonic tissues [104]. Interestingly, rolipram is shown to be superior to methylprednisolone in preventing late collagen deposition [105].
Overall, today, no PDE4 inhibitor has been developed for IBDs, although it appears to be more effective than methylprednisolone [105]. This may be possible because of limitations related to side effects of PDE4-Is, such as nausea and vomiting, see Ref. [9]. Today, with the new designs of PDE4-I, this should be re-investigated. In this way, it has recently been suggested that inhibition of PDE4 (notably by apremilast) may be a new therapeutic avenue for IBD [106]. Recently, a clinical study on the use of PDE4-I in psoriatic arthritis and IBD suggested that the use of apremilast should be studied merely in enteropathic spondylarthritis [107]. However, very few studies have focused on characterizing the altered PDE4 subtype in IBD. To go further, it would be of interest to develop either the selective PDE4 inhibitor acting specifically on the PDE4 subtype implicated in the disease or to investigate a new kind of PDE4-I such as allosteric PDE4-I [6,7]. A better prognosis of patients with Crohn’s disease has been recently proposed that combines endoscopic and radiological healing, which would favor the development of PDE4-Is [108].

2.5. PDE4 and Osteoporosis

Osteoporosis is a common disease characterized by reduced bone mineral density (BMD) and increased risk of fragility fractures, which is usually caused by osteoclasts, whereas osteopenia corresponds to a lower bone density than normal and represents the stage before osteoporosis. By 1970, no medications against osteoporosis were being investigated. At the time, fractures were not even recognized as an illness but were considered part of normal aging [109]. Nonetheless, osteoporosis is a public health issue worldwide, affecting over 200 million people. An estimated 30% to 50% of postmenopausal women have this disease. The prevalence (June 2009 to June 2010) of osteoporosis and osteopenia in healthy active workers is 18% [110]. However, it must be pointed out that osteoporosis is an age-associated disease. In the Eastern Mediterranean region, the overall pooled prevalence of osteoporosis was 24.4%, pointing out the necessity of investigating therapeutic approaches [111].
In that way, a study showed that denbufylline, an inhibitor of PDE4 [112,113], was shown to increase the number of mineralized nodules and decrease the number of osteoclast-like cells, suggesting that it should be a therapeutic drug for bone loss [114]. Therefore, this study was reinforced by studying the effect of a new PDE4 inhibitor, XT-44, on mineralized nodule formation, as well as in vivo in ovariectomized female Wistar rats. XT-44 stimulated the formation of mineralized nodules, while it inhibited the formation of osteoclast-like cells in mouse bone marrow culture. Interestingly, in ovariectomized female rats, 1 mg/kg per os (every 2 days for 8 weeks) increased bone mineral density, indicating that XT-44 may be effective in osteoporosis treatment [115].
In addition, pentoxifylline and rolipram have been shown to increase bone mass in normal mice mainly primarily by accelerating bone formation [116,117]. This was confirmed in adult ovariectomized rats in which rolipram at 0.1–0.6 mg/kg dose levels prevented, while at 1 mg/kg restored ovariectomy-induced cancellous and cortical bone loss [118]. Thus, studies performed on human culture cells showed by chromatographic separation that SaOS-2 human osteosarcoma cells contain PDE4 activity, as well as PDE1 and PDE7 activities, PDE4 activity being undetectable in cultured normal human osteoblasts. Nevertheless, transcripts of PDE4A and PDE4B were characterized in both kinds of human cells. Furthermore, only in sarcoma cells does rolipram raise the cAMP level according to the dose.
This human cell study, combined with previous studies [104,105,106,107], suggests that PDE4 could be a new target in the treatment of osteoporosis [119]. Therefore, since dexamethasone glucocorticoid is known to enable the development of osteoporosis, its effect was investigated on the expression of PDEs in SaOS-2 cultured sarcoma cells. Thus, in agreement with a decrease in PDE4 activity, and rolipram sensitivity, a 50–70% decrease in the mRNA of PDE4A and PDE4B subtypes, associated with a reduction in PDE4A4 and PDE4B1 isoforms were reported following dexamethasone treatment.
In total, these data point to the participation of PDE4 in osteoporosis [120]. Furthermore, it was shown that PDE4 inhibitor enhances PGE2-induced growth arrest of osteoclast formation. It has been found that the PDE4 inhibitor in one way promotes the formation of osteoblasts, and in another way, inhibits the formation of osteoclasts, both of which might may increase bone formation [121].
Although it was shown that PDE4-Is are effective in decreasing osteoporosis, this potential therapeutic approach always seems to not be recognized for osteoporosis [122]. Hopefully, very recently, Porwal et al. 2021 suggested repurposing the potential of PDE inhibitors in osteoporosis [123].

2.6. PDE4 and Malignancies

According to MedlinePlus Medical Encyclopedia, “the malignancy” refers to the presence of cancerous cells that have the ability to spread to other sites in the body (https://medlineplus.gov/enc/article/002253.htm, accessed on 13 November 2020). The global cancer incidence and mortality in 2020 were estimated at 19.3 million, almost a million new cases of cancer and 10 million deaths from cancer. The global cancer burden is expected to be 28.4 million cases in 2040 [124].
The use of PDE4-Is in various developing cancers was first investigated in various cancer cell lines as well as on cancer tissues [125,126,127]. Therefore, it was shown that PDE4 protein and mRNA up-regulations are associated in vivo with human endothelial cell proliferation and angiogenesis [45,128] as well as in vivo cancer development in mice [129]. Similarly, there has been reported a phosphodiesterase 4B-dependent interplay between tumor cells and the microenvironment regulating angiogenesis in B-cell lymphoma [130]. Since the PDE4-I, piclamilast, was able to potentiate the cyto-differentiating action of retinoids in myeloid leukemia cells [131], we investigated PDE activity and expression in retinoic acid-resistant cell lines in acute promyelocytic leukemia and showed an increase in PDE4 activity, accompanied by an increase in PDE4D expression [132]. In another area, studies performed in mice, rats, and human tissue revealed that PDE4B protects colon adenomas and is inactivated by epigenetic silencing in colon cancer [133].
Interestingly, the focal cAMP reduction achieved by forced expression of PDE4A1 in the cortex of neurofibromatosis-1 (NF1) genetically engineered mouse (GEM) model is sufficient to induce gliomagenesis in a mouse model of NF1, and conversely, rolipram dramatically inhibits in vivo optic glioma growth and tumor size in the model [134]. In ovarian cancer development, roflumilast mediates in vivo tumor inhibition by up-regulating mitochondrial ferritin in nude mice [135]. In another area, PDE4D7 was found to be a marker for the potential diagnosis of prostate cancer [136]. Thus, PDE4 inhibition decreases the malignant properties of DLD-1 colorectal cancerous cells by repressing the AKT/mTOR/Myc cancerous signaling pathway [137]. Conversely, PDE4 and EPAC1 synergistically promote rectal carcinoma via the cAMP pathway in human tissues, highlighting the PDE4 implication in colorectal cancer [138]. It must be pointed out that PDE4 inhibition by rolipram induced analgesia in bone cancer pain by inhibiting the activation of the spinal astrocytes and is associated with the downregulation of spinal IL-1β, IL-6, and TNF-α expressions [139].
Some PDE4-Is were recently considered cancer therapeutics, including apremilast [140]. In this area, it should be noted that patients with psoriasis may present a high risk of melanoma and hematological cancers, regardless of their psoriasis treatment [141]. Lastly, the participation of PDE4 subtypes in various cancers has been reviewed by Hsien Lai et al., 2020 [142].

2.7. PDE4 and Fatty Liver Disease

Fatty liver disease, currently named nonalcoholic fatty liver disease (NAFLD) according to the practice guidance commissioned by the American Association for the Study of Liver Diseases (AASLD), is defined by: (1) evidence of hepatic steatosis (HS), either by imaging or histology, and (2) lack of secondary causes of hepatic fat accumulation such as significant alcohol consumption, long-term use of a steatogenic medication, or monogenic hereditary disorders [143]. The overall global prevalence of NAFLD diagnosed by imaging is about 25.24% (95% CI, 22.10–28.65) [144]. NAFLD and alcoholic liver disease (ALD) are the leading causes of liver-related morbidity and mortality and are important causes of liver transplantation [145]. Using bile-duct ligation, as a cholestatic liver injury model, Gobejishvili and colleagues 2013 [146] demonstrated that induction of hepatic PDE4A, B, and D plays a causal role in the development of liver injury and fibrosis. In addition, roflumilast PDE4-I improved glucose tolerance, reduced insulin resistance, and decreased steatohepatitis in mice, increasing the cellular respiratory capacity of hepatocytes [93]. Interestingly, Vonghia et al., 2019 [145] mentioned that the compound tipelukast, also known as MN-001, designed as a leukotriene receptor antagonist, orally bioavailable, being anti-fibrotic and anti-inflammatory in pre-clinical models, interestingly, inhibits PDE3 and PDE4. However, a lack of efficacy of a PDE4-I in phase 1 and 2 trials of patients with non-alcoholic steatohepatitis has already been shown [147].
Overall, these data suggest that today PDE4-Is have failed in the treatment of NAFLD but have induced interesting positive data regarding ALD treatment. Thus, it has been shown that altered PDE4B plays a critical role in alcohol-induced steatosis [148]. A recent review designs PDE4 inhibition as a therapeutic target for ALD as much as, expression of the PDE4 subfamilies is significantly up-regulated in conjunction with markedly decreased cAMP levels in hepatic tissues of patients with severe ALD [149]. It should be noted that inhibition of PDE4 decreases ethanol intake in mice, extending the contribution of PDE4 to liver diseases [150]. Thus, contrary to ALD, one could speculate that liver PDE4 is not sufficiently induced in NAFLD to demonstrate changes in PDE4 expression and efficiency of PDE4-Is. Nonetheless, overexpression of PDE4 in the mouse liver is sufficient to trigger NAFLD and hypertension and can be avoided and even reversed by roflumilast PDE4-I [151]. A very recent study has supported the clinical use of a novel liposomal rolipram formulation to reduce emesis [152].

2.8. PDE4 and Depression

Major depression is a common illness that severely limits psychosocial functioning and diminishes quality of life. In 2008, the WHO ranked major depression as the third cause of burden of disease worldwide and projected that the disease will rank first by 2030, the 12-month prevalence of major depressive disorder (MDD) being approximately 6.6%, and the lifetime risk being 15–18% [153].
The potential antidepressant activity of rolipram, as a cAMP phosphodiesterase inhibitor, was canonically demonstrated in mice by Wachtel H, 1983 [154]. In that way, rolipram was first reported as a specific PDE4-I [12]. Rolipram was shown to have antidepressant-like effects on behavior maintained by differential reinforcement of low rat response [155] and to facilitate the establishment of long-lasting long-term potentiation and improve memory [156]. In PDE4D knockout (PDE4D−/−) mice, the loss of PDE4D in the cerebral cortex and hippocampus, interestingly, induces an antidepressant-like effect on behavior and reduces sensitivity to rolipram. Furthermore, rolipram potentiated isoproterenol-induced cyclic AMP formation only in the PDE4D+/+ mice. Interestingly, the PDE4D-regulated cyclic AMP signaling may play a role in the pathophysiology and pharmacotherapy of depression [157].
According to a review on CNS and PDEs, it is pointed out that PDE4A, 4B and 4D may participate in depression [158]. Mice deficient in the PDE4B gene (PDE4B−/−) displayed long-term depression as compared to wild-type littermates [159]. In this case, the use of novel allosteric small-molecule activators of the long-form PDE4 would be helpful [160]. However, the use of a new isoform selective technological approach, reducing dramatically the catalytic enzyme activity of PDE4B1, (mainly present in hippocampal CA2 and CA3 regions) being more specific than the PDE4B gene knockouts, allows seeing increased activity in these mice without major changes in associative learning or other behavior. These data might contribute to the development of PDE4-Is [161]. The development of new technologies has allowed demonstrating the implication of PDE4 in stress-induced depressive-like behaviors, since the knockdown of long forms of PDE4D4 and PDE4D5 in the mouse prefrontal cortex alleviates stress-induced depression and chronic unpredictable stress-induced depressive-like behaviors in mice [162]. In that way, roflumilast, a more potent and selective PDE4 inhibitor than rolipram, improves memory in rodents at non-emetic doses, pointing out PDE4 as a possible therapeutic target in cognition, demonstrating the therapeutic potential of PDE4 inhibitors in the brain [163].
Therefore, strategies to develop antidepressant PDE4 inhibitors is to develop a new generation of PDE4 inhibitors which act specifically on long PDE4 isoforms. They might also be capable of targeting the interactions with their signaling “partner” proteins, such as RACK1, β-arrestin, and DISC1 in order to affect cAMP levels in specific cellular compartments [164]. In addition, the understood mechanism of PDE4D involvement in Alzheimer’s disease opens up new therapeutic approaches that may focus on the activation status of PDE4 [165]. Interestingly, the identification of diagnostic markers MDD (761 differentially expressed genes closely related to MDD) has recently been validated in total blood samples. The gene set enrichment analysis suggested that the tumor necrosis factor signaling pathway, toll-like receptor signaling pathway, apoptosis pathway, and NFkB signaling pathway are all crucial in the development of MDD [166] in accordance with known PDE4-dependent signaling. Recently, these approaches were conducted in a pre-clinical study using roflumilast in transgenic APP/PS1mice [167]. Since all these pathways concern PDE4 participation, it would be very interesting to include PDE4 as a diagnostic marker in MDD. Furthermore, these diagnostic markers will help in the development of specific PDE4 inhibitors targeting MDD with lower adverse effects. In that way, a recent review on PDE4 inhibition in the CNS, suggests new approaches focused on targeting PDE4 conformational states and on influencing the binding affinity of PDE4 subtype inhibitors [168].

2.9. PDE4, and COVID-19

As previously [169] stated, during the 1990s, PDE4 might participate in viral infections, since the PDE4 inhibitor, rolipram, was canonically shown to inhibit human immunodeficiency virus-1 (HIV-1) replication, and to decrease HIV-1 p24 antigen production in acutely HIV-1 infected PBMCs [170]. This rolipram inhibitory effect on HIV-1 replication in Jurkat and primary T-cells induced by T-cell activation has been confirmed and extended to the production of TNF-α, NFκB, and NFAT activation [171]. Beavo and colleagues showed that infection of CD4+ memory T-cells by HIV-1 requires the expression of PDE4, and that rolipram abolishes HIV-1 DNA nuclear import in memory T cells, pointing out the important contribution of PDE4 in viral infection [172]. Roflumilast inhibits respiratory syncytial virus infection in differentiated human bronchial epithelial cells [173]. However, in the human airway smooth muscle infected by rhinovirus, a non-enveloped RNA virus, surprisingly, inhibition of PDE4 did not overcome cytokine induction [174], opening the hypothesis that the viral envelope is a requisite for cytokine induction and PDE4-I effectiveness. Interestingly, it has been reported that cAMP produced in Tregs is implicated in suppressing the activation and expression of the HIV-1 gene in vivo in humanized mice [175]. Since a relationship between HIV-1 and PDE4 was clearly established, one could wonder whether COVID-19 might modify PDE4 regulation.
Interestingly, a paper published in Nature Medicine (2005), demonstrated that angiotensin-converting enzyme 2 (ACE2) plays a strategic role in SARS-CoV-2-induced lung injury. As expected, the injection of SARS-CoV-2 spikes into mice worsens acute lung failure in vivo which can be alleviated by blocking the renin–angiotensin pathway governed by ACE2 [176]. Furthermore, it has been reported that Ang-II, an AT1R agonist, induces PDE4 up-regulation, with a 44% increase in the PDE4A protein, mediating cascade inflammation and oxidative stress. Altogether, this opens a new way of PDE interaction with the renin–angiotensin system [71]. More recently, relationships between PDE4, and cytokine storm were discussed for COVID-19 [177,178,179,180]. In this way, a recent in vitro study (phase III) performed with the inhaled PDE4 inhibitor, tanamilast (CHF6001), showed that it blunts proinflammatory dendritic cell activation by SARS-CoV-2 ssRNAs [181].
Altogether, these data reported for COVID-19 suggest the hypothesis that a PDE4 inhibitor might alleviate both viral infection and tissue inflammation induced by SARS-CoV-2 [169].

3. Conclusions and Perspectives

Altogether, PDE4s are implicated in various human pathologies, such as cardiovascular diseases, obesity, diabetes, ulcerative colitis and Crohn’s, osteoporosis, malignancies, fatty liver disease, and depression, as well as COVID-19 (Figure 7). Given that PDE4 is an intracellular target, it is possible that at the same time PDE4 may participate differently in two or more diseases. Therefore, there has recently been a reported risk of malignancy, and recurrence associated with psoriasis and its treatments. This was pointed out for PDE4-Is in patients with a history of cancer, suggesting the necessity of long-term studies focused on safety surveillance [182]. In another way, the neuroprotective effect of a PDE4 inhibitor (FCPR03) against oxygen-glucose deprivation has been recently related to the activation of nuclear factor erythroid 2-related factor 2 (Nrf-2)/heme oxygenase-1(HO-1) attenuating ROS production and consequently endoplasmic reticulum stress and apoptosis in neurons, allowing to decipher its molecular mechanism [183].
To further explore the therapeutic potential of PDE4-Is, different strategic approaches could be used to reduce adverse effects. Since PDE4 could act in interactome signaling, it should be possible to act on the protein–protein interactions governing PDE activity in various cell compartments [184,185]. This was illustrated for a multifunctional docking site on the catalytic unit of PDE4 which is used by several interaction partners [186]. Of interest, a recent review analyzed in detail the advances in knowledge and in patent developments. It originally reports the diverse ways to interact with PDE signals directly or not, in order to overcome PDE dysregulations associated with various dysfunctions [187]. Nevertheless, it should be pointed out that other PDE families present in the vicinity of PDE4 might interfere with PDE4 responses such as PDE3 [39,41], PDE2 [45,188], as well as PDE1 and PDE5 [71].
The present paper reports the implication of dysregulation of PDE4 activity in various pathologic fields pointing out the effectiveness of PDE4 inhibitors as well as the increases in the knowledge of PDE4 functions related to their multiplicity at cellular, subcellular, and molecular levels. This introduces new approaches to powerful, selective, and effective treatment, with lesser adverse effects, opening up a new generation of more suitable PDE4 inhibitors. This was just newly discussed and reviewed for clinical implications of PDE4-Is [189].

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lugnier, C. Cyclic nucleotide phosphodiesterase (PDE) superfamily: A new target for the development of specific therapeutic agents. Pharmacol. Ther. 2006, 109, 366–398. [Google Scholar] [CrossRef] [PubMed]
  2. Mulller, B.; Komas, N.; Keravis, T.; Lugnier, C. Les phosphodiestérases des nucléotides cycliques. Med. Sci. 1993, 12, 1335–1341. [Google Scholar]
  3. Dessauer, C.W.; Watts, V.J.; Ostrom, R.S.; Conti, M.; Dove, S.; Seifert, R. International Union of Basic and Clinical Pharmacology. CI. Structures and small molecule modulators of mammalian adenylyl cyclases. Pharmacol. Rev. 2017, 69, 93–139. [Google Scholar] [CrossRef]
  4. Pozdniakova, S.; Ladilov, Y. Functional significance of the adcy10-Dependent intracellular cAMP Compartments. J. Cardiovasc. Dev. Dis. 2018, 5, E29. [Google Scholar] [CrossRef]
  5. Francis, S.H.; Houslay, M.D.; Conti, M. Phosphodiesterase inhibitors: Factors that influence potency, selectivity, and action. Handb. Exp. Pharmacol. 2011, 204, 47–84. [Google Scholar]
  6. Keravis, T.; Lugnier, C. Cyclic nucleotide phosphodiesterase (PDE) isozymes as targets of the intracellular signalling network: Benefits of PDE inhibitors in various diseases and perspectives for future therapeutic developments. Br. J. Pharmacol. 2012, 165, 1288–1305. [Google Scholar] [CrossRef]
  7. Maurice, D.H.; Ke, H.; Ahmad, F.; Wang, Y.; Chung, J.; Manganiello, V.C. Advances in targeting cyclic nucleotide phosphodiesterases. Nat. Rev. Drug. Discov. 2014, 13, 290–314. [Google Scholar] [CrossRef] [PubMed]
  8. Ahmad, F.; Murata, T.; Shimizu, K.; Degerman, E.; Maurice, D.; Manganiello, V. Cyclic nucleotide phosphodiesterases: Important signaling modulators and therapeutic targets. Oral Dis. 2015, 21, e25–e50. [Google Scholar] [CrossRef]
  9. Lugnier, C.; Meyer, A.; Charloux, A.; Andrès, E.; Geny, B.; Talha, S. The Endocrine function of the heart: Physiology and involvements of natriuretic peptides and cyclic nucleotide phosphodiesterases in heart failure. J. Clin. Med. 2019, 8, 1746. [Google Scholar] [CrossRef]
  10. Lugnier, C.; Meyer, A.; Talha, S.; Geny, B. Cyclic nucleotide phosphodiesterases: New targets in the metabolic syndrome? Pharmacol. Ther. 2020, 208, 107475. [Google Scholar] [CrossRef]
  11. Fertig, B.A.; Baillie, G.S. PDE4-Mediated cAMP Signalling. J. Cardiovasc. Dev. Dis. 2018, 5, E8. [Google Scholar] [CrossRef] [PubMed]
  12. Lugnier, C.; Stierlé, A.; Beretz, A.; Schoeffter, P.; Lebec, A.; Wermuth, C.G.; Cazenave, J.P.; Stoclet, J.C. Tissue and substrate specificity of inhibition by alkoxy-aryl-lactams of platelet and arterial smooth muscle cyclic nucleotide phosphodiesterases relationship to pharmacological activity. Biochem. Biophys. Res. Commun. 1983, 113, 954–959. [Google Scholar] [CrossRef]
  13. Lugnier, C.; Schoeffter, P.; Le Bec, A.; Strouthou, E.; Stoclet, J.C. Selective inhibition of cyclic nucleotide phosphodiesterases of human, bovine and rat aorta. Biochem. Pharmacol. 1986, 35, 1743–1751. [Google Scholar] [CrossRef]
  14. Schneider, H.H.; Schmiechen, R.; Brezinski, M.; Seidler, J. Stereospecific binding of the antidepressant rolipram to brain protein structures. Eur. J. Pharmacol. 1986, 127, 105–115. [Google Scholar] [CrossRef]
  15. Marivet, M.C.; Bourguignon, J.J.; Lugnier, C.; Mann, A.; Stoclet, J.C.; Wermuth, C.G. Inhibition of cyclic adenosine-3′,5′-monophosphate phosphodiesterase from vascular smooth muscle by rolipram analogues. J. Med. Chem. 1989, 32, 1450–1457. [Google Scholar] [CrossRef]
  16. Dent, G.; Giembycz, M.A.; Rabe, K.F.; Barnes, P.J. Inhibition of eosinophil cyclic nucleotide PDE activity and opsonised zymosan-stimulated respiratory burst by ‘type IV’-selective PDE inhibitors. Br. J. Pharmacol. 1991, 103, 1339–1346. [Google Scholar] [CrossRef]
  17. Livi, G.P.; Kmetz, P.; McHale, M.M.; Cieslinski, L.B.; Sathe, G.M.; Taylor, D.P.; Davis, R.L.; Torphy, T.J.; Balcarek, J.M. Cloning and expression of cDNA for a human low-Km, rolipram-sensitive cyclic AMP phosphodiesterase. Mol. Cell. Biol. 1990, 10, 2678–2686. [Google Scholar]
  18. Monaco, L.; Vicini, E.; Conti, M. Structure of two rat genes coding for closely related rolipram-sensitive cAMP phosphodiesterases. Multiple mRNA variants originate from alternative splicing and multiple start sites. J. Biol. Chem. 1994, 269, 347–357. [Google Scholar] [CrossRef]
  19. Houslay, M.D.; Schafer, P.; Zhang, K.Y. Keynote review: Phosphodiesterase-4 as a therapeutic target. Drug Discov. Today 2005, 10, 1503–1519. [Google Scholar] [CrossRef]
  20. Houslay, M.D.; Sullivan, M.; Bolger, G.B. The multienzyme PDE4 cyclic adenosine monophosphate-specific phosphodiesterase family: Intracellular targeting, regulation, and selective inhibition by compounds exerting anti-inflammatory and antidepressant actions. Adv. Pharmacol. 1998, 44, 225–342. [Google Scholar]
  21. Richter, W.; Conti, M. The oligomerization state determines regulatory properties and inhibitor sensitivity of type 4 cAMP-specific phosphodiesterases. J. Biol. Chem. 2004, 279, 30338–30348. [Google Scholar] [CrossRef] [PubMed]
  22. Francis, S.H.; Turko, I.V.; Corbin, J.D. Cyclic nucleotide phosphodiesterases: Relating structure and function. Prog. Nucleic Acid Res. Mol. Biol. 2001, 65, 1–52. [Google Scholar] [PubMed]
  23. Baillie, G.S.; MacKenzie, S.J.; McPhee, I.; Houslay, M.D. Sub-family selective actions in the ability of Erk2 MAP kinase to phosphorylate and regulate the activity of PDE4 cyclic AMP-specific phosphodiesterases. Br. J. Pharmacol. 2000, 131, 811–819. [Google Scholar] [CrossRef]
  24. Bolger, G.B.; Conti, M.; Houslay, M.D. Cellular functions of PDE4 enzymes. In Cyclic Nucleotide Phosphodiesterases in Health and Disease; Beavo, J.A., Francis, S.H., Houslay, M., Eds.; CRC Press: Boca Raton, FL, USA, 2006; pp. 99–129. [Google Scholar]
  25. Barnette, M.S.; Manning, C.D.; Cieslinski, L.B.; Burman, M.; Christensen, S.B.; Torphy, T.J. The ability of phosphodiesterase IV inhibitors to suppress superoxide production in guinea pig eosinophils is correlated with inhibition of phosphodiesterase IV catalytic activity. J. Pharmacol. Exp. Ther. 1995, 273, 674–679. [Google Scholar] [PubMed]
  26. Barnette, M.S.; Grous, M.; Cieslinski, L.B.; Burman, M.; Christensen, S.B.; Torphy, T.J. Inhibitors of phosphodiesterase IV (PDE IV) increase acid secretion in rabbit isolated gastric glands: Correlation between function and interaction with a high-affinity rolipram binding site. J. Pharmacol. Exp. Ther. 1995, 273, 1396–1402. [Google Scholar] [PubMed]
  27. Jacobitz, S.; McLaughlin, M.M.; Livi, G.P.; Burman, M.; Torphy, T.J. Mapping the functional domains of human recombinant phosphodiesterase 4A: Structural requirements for catalytic activity and rolipram binding. Mol. Pharmacol. 1996, 50, 891–899. [Google Scholar]
  28. Souness, J.E.; Rao, S. Proposal for pharmacologically distinct conformers of PDE4 cyclic AMP phosphodiesterases. Cell Signal. 1997, 9, 227–236. [Google Scholar] [CrossRef]
  29. Méhats, C.; Tanguy, G.; Paris, B.; Robert, B.; Pernin, N.; Ferré, F.; Leroy, M.J. Pregnancy induces a modulation of the cAMP phosphodiesterase 4-conformers ratio in human myometrium: Consequences for the utero-relaxant effect of PDE4-selective inhibitors. J. Pharmacol. Exp. Ther. 2000, 292, 817–823. [Google Scholar]
  30. Zhang, H.T.; Zhao, Y.; Huang, Y.; Deng, C.; Hopper, A.T.; De Vivo, M.; Rose, G.M.; O’Donnell, J.M. Antidepressant-like effects of PDE4 inhibitors mediated by the high-affinity rolipram binding state (HARBS) of the phosphodiesterase-4 enzyme (PDE4) in rats. Psychopharmacology 2006, 186, 209–217. [Google Scholar] [CrossRef]
  31. Hatzelmann, A.; Morcillo, E.J.; Lungarella, G.; Adnot, S.; Sanjar, S.; Beume, R.; Schudt, C.; Tenor, H. The preclinical pharmacology of roflumilast–A selective, oral phosphodiesterase 4 inhibitor in development for chronic obstructive pulmonary disease. Pulm. Pharmacol. Ther. 2010, 23, 235–256. [Google Scholar] [CrossRef]
  32. Fala, L. Otezla (Apremilast), an oral PDE-4 inhibitor, receives FDA approval for the treatment of patients with active psoriatic arthritis and plaque psoriasis. Am. Health Drug Benefits 2015, 8, 105–110. [Google Scholar] [PubMed]
  33. Burgin, A.B.; Magnusson, O.T.; Singh, J.; Witte, P.; Staker, B.L.; Bjornsson, J.M.; Thorsteinsdottir, M.; Hrafnsdottir, S.; Hagen, T.; Kiselyov, A.S.; et al. Design of phosphodiesterase 4D (PDE4D) allosteric modulators for enhancing cognition with improved safety. Nat. Biotechnol. 2010, 28, 63–70. [Google Scholar] [CrossRef]
  34. Zhang, C.; Xu, Y.; Chowdhary, A.; Fox, D., 3rd; Gurney, M.E.; Zhang, H.T.; Auerbach, B.D.; Salvi, R.J.; Yang, M.; Li, G.; et al. Memory enhancing effects of BPN14770, an allosteric inhibitor of phosphodiesterase-4D, in wild-type and humanized mice. Neuropsychopharmacology 2018, 43, 2299–2309. [Google Scholar] [CrossRef] [PubMed]
  35. Keravis, T.; Lugnier, C. Cyclic nucleotide phosphodiesterases (PDE) and peptide motifs. Curr. Pharm. Des. 2010, 16, 1114–1125. [Google Scholar] [CrossRef] [PubMed]
  36. Richter, W.; Mika, D.; Blanchard, E.; Day, P.; Conti, M. ß1-adrenergic receptor antagonists signal via PDE4 translocation. EMBO Rep. 2013, 14, 276–283. [Google Scholar] [CrossRef] [PubMed]
  37. Feng, X.; Wang, H.; Ye, M.; Xu, X.T.; Xu, Y.; Yang, W.; Zhang, H.T.; Song, G.; Ke, H. Identification of a PDE4-Specific Pocket for the Design of Selective Inhibitors. Biochemistry 2018, 57, 4518–4525. [Google Scholar] [CrossRef] [PubMed]
  38. Stoclet, J.C.; Keravis, T.; Komas, N.; Lugnier, C. Cyclic nucleotide phosphodiesterases as therapeutic targets in cardiovascular diseases. Expert Opin. Investig. Drugs 1995, 4, 1081–1100. [Google Scholar]
  39. Komas, N.; Lugnier, C.; Stoclet, J.C. Endothelium-dependent and independent relaxation of the rat aorta by cyclic nucleotide phosphodiesterase inhibitors. Br. J. Pharmacol. 1991, 104, 495–503. [Google Scholar] [CrossRef]
  40. Lugnier, C.; Komas, N. Modulation of vascular cyclic nucleotide phosphodiesterases by cyclic GMP: Role in vasodilatation. Eur. Heart J. 1993, 14 (Suppl. 1), 141–148. [Google Scholar]
  41. Eckly, A.E.; Lugnier, C. Role of phosphodiesterases III and IV in the modulation of vascular cyclic AMP content by the NO/cyclic GMP pathway. Br. J. Pharmacol. 1994, 113, 445–450. [Google Scholar] [CrossRef]
  42. Keravis, T.; Lugnier, C. Cyclic nucleotide phosphodiesterase (PDE) superfamily and smooth muscle signaling. In New Frontiers in Smooth Muscle Biology and Physiology; Savineau, J.P., Ed.; Transworld Research, Network: Kerala, India, 2007; pp. 269–289. [Google Scholar]
  43. Lugnier, C.; Schini, V.B. Characterization of cyclic nucleotide phosphodiesterases from cultured bovine aortic endothelial cells. Biochem. Pharmacol. 1990, 39, 75–84. [Google Scholar] [CrossRef]
  44. Keravis, T.; Komas, N.; Lugnier, C. Cyclic nucleotide hydrolysis in bovine aortic endothelial cells in culture: Differential regulation in cobblestone and spindle phenotypes. J. Vasc. Res. 2000, 37, 235–249. [Google Scholar] [CrossRef] [PubMed]
  45. Favot, L.; Keravis, T.; Lugnier, C. Modulation of VEGF-induced endothelial cell cycle protein expression through cyclic AMP hydrolysis by PDE2 and PDE4. Thromb. Haemost. 2004, 92, 634–645. [Google Scholar] [CrossRef]
  46. Keravis, T.; Silva, A.P.; Favot, L.; Lugnier, C. Role of PDEs in vascular health and disease: Endothelial PDEs and angiogenesis. In Cyclic Nucleotide Phosphodiesterases in Health and Disease; Beavo, J.A., Francis, S.H., Houslay, M., Eds.; CRC Press: Boca Raton, FL, USA, 2006; pp. 417–439. [Google Scholar]
  47. Kessler, T.; Lugnier, C. Rolipram increases cyclic GMP content in L-arginine-treated cultured bovine aortic endothelial cells. Eur. J. Pharmacol. 1995, 290, 163–167. [Google Scholar] [CrossRef]
  48. Suttorp, N.; Ehreiser, P.; Hippenstiel, S.; Fuhrmann, M.; Krüll, M.; Tenor, H.; Schudt, C. Hyperpermeability of pulmonary endothelial monolayer: Protective role of phosphodiesterase isoenzymes 3 and 4. Lung 1996, 174, 181–194. [Google Scholar] [PubMed]
  49. Rampersad, S.N.; Ovens, J.D.; Huston, E.; Umana, M.B.; Wilson, L.S.; Netherton, S.J.; Lynch, M.J.; Baillie, G.S.; Houslay, M.D.; Maurice, D.H. Cyclic AMP phosphodiesterase 4D (PDE4D) tethers EPAC1 in a vascular endothelial cadherin (VE-Cad)-based signaling complex and controls cAMP-mediated vascular permeability. J. Biol. Chem. 2010, 285, 33614–33622. [Google Scholar] [PubMed]
  50. Hubert, F.; Belacel-Ouari, M.; Manoury, B.; Zhai, K.; Domergue-Dupont, V.; Mateo, P.; Joubert, F.; Fischmeister, R.; Leblais, V. Alteration of vascular reactivity in heart failure: Role of phosphodiesterases 3 and 4. Br. J. Pharmacol. 2014, 171, 5361–5375. [Google Scholar] [CrossRef]
  51. Varona, S.; Puertas, L.; Galán, M.; Orriols, M.; Cañes, L.; Aguiló, S.; Camacho, M.; Sirvent, M.; Andrés, V.; Martínez-González, J.; et al. Rolipram prevents the formation of abdominal aortic aneurysm (AAA) in Mice: PDE4B as a target in AAA. Antioxidants 2021, 10, 460. [Google Scholar] [CrossRef]
  52. Wollborn, J.; Siemering, S.; Steiger, C.; Buerkle, H.; Goebel, U.; Schick, M.A. Phosphodiesterase-4 inhibition reduces ECLS-induced vascular permeability and improves microcirculation in a rodent model of extracorporeal resuscitation. Am. J. Physiol. Heart. Circ. Physiol. 2019, 316, H751–H761. [Google Scholar]
  53. Tsien, R.W. Cyclic AMP and contractile activity in heart. Adv. Cycl. Nucleotide Res. 1977, 8, 363–420. [Google Scholar]
  54. Harrison, S.A.; Reifsnyder, D.H.; Gallis, B.; Cadd, G.G.; Beavo, J.A. Isolation and characterization of bovine cardiac muscle cGMP- inhibited phosphodiesterase: A receptor for new cardiotonic drugs. Mol. Pharmacol. 1986, 29, 506–514. [Google Scholar] [PubMed]
  55. Komas, N.; Lugnier, C.; Le Bec, A.; Serradeil-Le Gal, C.; Barthelemy, G.; Stoclet, J.C. Differential sensitivity to cardiotonic drugs of cyclic AMP phosphodiesterases isolated from canine ventricular and sinoatrial-enriched tissues. J. Cardiovasc. Pharmacol. 1989, 14, 213–220. [Google Scholar] [CrossRef]
  56. Muller, B.; Lugnier, C.; Stoclet, J.C. Implication of cyclic AMP in the positive inotropic effects of cyclic GMP-inhibited cyclic AMP phosphodiesterase inhibitors on guinea pig isolated left atria. J. Cardiovasc. Pharmacol. 1990, 15, 444–451. [Google Scholar] [CrossRef]
  57. Lanfear, D.E.; Hasan, R.; Gupta, R.C.; Williams, C.; Czerska, B.; Tita, C.; Bazari, R.; Sabbah, H.N. Short term effects of milrinone on biomarkers of necrosis, apoptosis, and inflammation in patients with severe heart failure. J. Transl. Med. 2009, 7, 67. [Google Scholar] [CrossRef]
  58. Prigent, A.F.; Fougier, S.; Nemoz, G.; Anker, G.; Pacheco, H.; Lugnier, C.; Lebec, A.; Stoclet, J.C. Comparison of cyclic nucleotide phosphodiesterase isoforms from rat heart and bovine aorta. Separation and inhibition by selective reference phosphodiesterase inhibitors. Biochem. Pharmacol. 1988, 37, 3671–3681. [Google Scholar] [CrossRef]
  59. Stoclet, J.C.; Boulanger-Saunier, C.; Lassègue, B.; Lugnier, C. Cyclic nucleotides and calcium regulation in heart and smooth muscle cells. Ann. N. Y. Acad. Sci. 1988, 522, 106–115. [Google Scholar] [CrossRef]
  60. Lugnier, C.; Gauthier, C.; Le Bec, A.; Soustre, H. Cyclic nucleotide phosphodiesterases from frog atrial fibers: Isolation and drug sensitivities. Am. J. Physiol. 1992, 262, H654–H660. [Google Scholar] [CrossRef] [PubMed]
  61. Muller, B.; Stoclet, J.C.; Lugnier, C. Cytosolic and membrane-bound cyclic nucleotide phosphodiesterases from guinea pig cardiac ventricles. Eur. J. Pharmacol. 1992, 225, 263–272. [Google Scholar] [CrossRef]
  62. Muller, B.; Lugnier, C.; Stoclet, J.C. Involvement of rolipram-sensitive cyclic AMP phosphodiesterase in the regulation of cardiac contraction. Cardiovasc. Pharmacol. 1990, 16, 796–803. [Google Scholar]
  63. Eschenhagen, T. PDE4 in the human heart–major player or little helper? Br. J. Pharmacol. 2013, 168, 524–527. [Google Scholar] [CrossRef]
  64. Lugnier, C.; Muller, B.; Le Bec, A.; Beaudry, C.; Rousseau, E. Characterization of indolidan- and rolipram-sensitive cyclic nucleotide phosphodiesterases in canine and human cardiac microsomal fractions. J. Pharmacol. Exp. Ther. 1993, 265, 1142–1151. [Google Scholar] [PubMed]
  65. Lugnier, C.; Keravis, T.; Le Bec, A.; Pauvert, O.; Proteau, S.; Rousseau, E. Characterization of cyclic nucleotide phosphodiesterase isoforms associated to isolated cardiac nuclei. Biochim. Biophys. Acta 1999, 1472, 431–446. [Google Scholar] [CrossRef]
  66. Bedioune, I.; Lefebvre, F.; Lechêne, P.; Varin, A.; Domergue, V.; Kapiloff, M.S.; Fischmeister, R.; Vandecasteele, G. PDE4 and mAKAPβ are nodal organizers of β2-ARs nuclear PKA signalling in cardiac myocytes. Cardiovasc. Res. 2018, 114, 1499–1511. [Google Scholar] [CrossRef] [PubMed]
  67. Verde, I.; Pahlke, G.; Salanova, M.; Zhang, G.; Wang, S.; Coletti, D.; Onuffer, J.; Jin, S.L.; Conti, M. Myomegalin is a novel protein of the Golgi/centrosome that interacts with a cyclic nucleotide phosphodiesterase. J. Biol. Chem. 2001, 276, 11189–11198. [Google Scholar] [CrossRef]
  68. Lehnart, S.E.; Wehrens, X.H.; Reiken, S.; Warrier, S.; Belevych, A.E.; Harvey, R.D.; Richter, W.; Jin, S.L.; Conti, M.; Marks, A.R. Phosphodiesterase 4D deficiency in the ryanodine-receptor complex promotes heart failure and arrhythmias. Cell 2005, 123, 25–35, Erratum in Cell 2005, 123, 535–536. [Google Scholar]
  69. Abi-Gerges, A.; Richter, W.; Lefebvre, F.; Mateo, P.; Varin, A.; Heymes, C.; Samuel, J.L.; Lugnier, C.; Conti, M.; Fischmeister, R.; et al. Decreased expression and activity of cAMP phosphodiesterases in cardiac hypertrophy and its impact on beta-adrenergic cAMP signals. Circ. Res. 2009, 105, 784–792. [Google Scholar]
  70. Richter, W.; Xie, M.; Scheitrum, C.; Krall, J.; Movsesian, M.A.; Conti, M. Conserved expression and functions of PDE4 in rodent and human heart. Basic Res. Cardiol. 2011, 106, 249–262. [Google Scholar]
  71. Mokni, W.; Keravis, T.; Etienne-Selloum, N.; Walter, A.; Kane, M.O.; Schini-Kerth, V.B.; Lugnier, C. Concerted regulation of cGMP and cAMP phosphodiesterases in early cardiac hypertrophy induced by angiotensin II. PLoS ONE 2010, 5, e14227. [Google Scholar] [CrossRef]
  72. Beca, S.; Helli, P.B.; Simpson, J.A.; Zhao, D.; Farman, G.P.; Jones, P.; Tian, X.; Wilson, L.S.; Ahmad, F.; Chen, S.R.W.; et al. Phosphodiesterase 4D (PDE4D) regulates baseline sarcoplasmic reticulum Ca2+ release and cardiac contractility, independently of L-type Ca2+ current. Circ. Res. 2011, 109, 1024–1030. [Google Scholar]
  73. Leroy, J.; Richter, W.; Mika, D.; Castro, L.R.; Abi-Gerges, A.; Xie, M.; Scheitrum, C.; Lefebvre, F.; Schittl, J.; Mateo, P.; et al. Phosphodiesterase 4B in the cardiac L-type Ca2+ channel complex regulates Ca2+ current and protects against ventricular arrhythmias in mice. J. Clin. Investig. 2011, 121, 2651–2661. [Google Scholar] [CrossRef]
  74. Wang, L.; Burmeister, B.T.; Johnson, K.R.; Baillie, G.S.; Karginov, A.V.; Skidgel, R.A.; O’Bryan, J.P.; Carnegie, G.K. UCR1C is a novel activator of phosphodiesterase 4 (PDE4) long isoforms and attenuates cardiomyocyte hypertrophy. Cell Signal. 2015, 27, 908–922. [Google Scholar] [PubMed]
  75. Lindner, M.; Mehel, H.; David, A.; Leroy, C.; Burtin, M.; Friedlander, G.; Terzi, F.; Mika, D.; Fischmeister, R.; Prié, D. Fibroblast growth factor 23 decreases PDE4 expression in heart increasing the risk of cardiac arrhythmia; Klotho opposes these effects. Basic Res. Cardiol. 2020, 115, 51. [Google Scholar] [PubMed]
  76. Levian, C.; Ruiz, E.; Yang, X. The pathogenesis of obesity from a genomic and systems biology perspective. Yale J. Biol. Med. 2014, 87, 113–126. [Google Scholar]
  77. Hales, C.M.; Carroll, M.D.; Fryar, C.D.; Ogden, C.L. Prevalence of obesity among adults and youth: United States, 2015–2016. NCHS Data Briefs 2020, 360, 1–8. [Google Scholar]
  78. Chung, J.H.; Manganiello, V.; Dyck, J.R. Resveratrol as a calorie restriction mimetic: Therapeutic implications. Trends Cell Biol. 2012, 22, 546–554. [Google Scholar] [CrossRef]
  79. Omar, B.; Banke, E.; Ekelund, M.; Frederiksen, S.; Degerman, E. Alterations in cyclic nucleotide phosphodiesterase activities in omental and subcutaneous adipose tissues in human obesity. Nutr. Diabetes 2011, 1, e13. [Google Scholar] [CrossRef]
  80. Kraynik, S.M.; Miyaoka, R.S.; Beavo, J.A. PDE3 and PDE4 isozyme-selective inhibitors are both required for synergistic activation of brown adipose tissue. Mol. Pharmacol. 2013, 83, 1155–1165. [Google Scholar] [CrossRef] [Green Version]
  81. Jensterle, M.; Salamun, V.; Kocjan, T.; Vrtacnik Bokal, E.; Janez, A. Short term monotherapy with GLP-1 receptor agonist liraglutide or PDE 4 inhibitor roflumilast is superior to metformin in weight loss in obese PCOS women: A pilot randomized study. J. Ovarian Res. 2015, 8, 32. [Google Scholar] [CrossRef]
  82. Zhang, R.; Maratos-Flier, E.; Flier, J.S. Reduced adiposity and high-fat diet-induced adipose inflammation in mice deficient for phosphodiesterase 4B. Endocrinology 2008, 150, 3076–3082. [Google Scholar]
  83. Al Amrani, A.; Abdel Karim, M.; Al Zoghaibi, M. PRDM16 Gene Polymorphism is associated with obesity and blood lipids profiles in Saudi population. J. Clin. Med. 2018, 7, E141. [Google Scholar] [CrossRef]
  84. Muo, I.M.; Park, S.J.; Smith, A.; Springer, A.A.; Allen, M.D.; Hagen, T.J.; Chung, J.H. Compound D159687, a phosphodiesterase 4D inhibitor, induces weight and fat mass loss in aged mice without changing lean mass, physical and cognitive function. Biochem. Biophys. Res. Commun. 2018, 506, 1059–1064. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, W.; Zhang, J.; Xiao, J. Roflumilast suppresses adipogenic differentiation via AMPK mediated pathway. Front. Endocrinol. 2021, 12, 662451. [Google Scholar] [CrossRef] [PubMed]
  86. Ogurtsova, K.; da Rocha Fernandes, J.D.; Huang, Y.; Linnenkamp, U.; Guariguata, L.; Cho, N.H.; Cavan, D.; Shaw, J.E.; Makaroff, L.E. IDF Diabetes Atlas: Global estimates for the prevalence of diabetes for 2015 and 2040. Diabetes Res. Clin. Pract. 2017, 128, 40–50. [Google Scholar] [CrossRef] [PubMed]
  87. Marasco, M.R.; Linnemann, A.K. β-Cell Autophagy in Diabetes Pathogenesis. Endocrinology 2018, 159, 2127–2141. [Google Scholar] [CrossRef]
  88. Yazdanpanah, S.; Rabiee, M.; Tahriri, M.; Abdolrahim, M.; Rajab, A.; Jazayeri, H.E.; Tayebi, L. Evaluation of glycated albumin (GA) and GA/HbA1c ratio for diagnosis of diabetes and glycemic control: A comprehensive review. Crit. Rev. Clin. Lab. Sci. 2017, 54, 219–232. [Google Scholar] [CrossRef]
  89. Lugnier, C. Cyclic nucleotide phosphodiesterase families in intracellular signaling and diabetes. Adv. Exp. Med. Biol. 2001, 498, 253–261. [Google Scholar]
  90. Parker, J.C.; VanVolkenburg, M.A.; Ketchum, R.J.; Brayman, K.L.; Andrews, K.M. Cyclic AMP phosphodiesterases of human and rat islets of Langerhans: Contributions of types III and IV to the modulation of insulin secretion. Biochem. Biophys. Res. Commun. 1995, 217, 916–923. [Google Scholar] [CrossRef]
  91. Waddleton, D.; Wu, W.; Feng, Y.; Thompson, C.; Wu, M.; Zhou, Y.P.; Howard, A.; Thornberry, N.; Li, J.; Mancini, J.A. Phosphodiesterase 3 and 4 comprise the major cAMP metabolizing enzymes responsible for insulin secretion in INS-1 (832/13) cells and rat islets. Biochem. Pharmacol. 2008, 76, 884–893. [Google Scholar] [CrossRef]
  92. Tian, G.; Sågetorp, J.; Xu, Y.; Shuai, H.; Degerman, E.; Tengholm, A. Role of phosphodiesterases in the shaping of sub-plasma-membrane cAMP oscillations and pulsatile insulin secretion. J. Cell Sci. 2012, 125, 5084–5095. [Google Scholar] [CrossRef] [Green Version]
  93. Möllmann, J.; Kahles, F.; Lebherz, C.; Kappel, B.; Baeck, C.; Tacke, F.; Werner, C.; Federici, M.; Marx, N.; Lehrke, M. The PDE4 inhibitor roflumilast reduces weight gain by increasing energy expenditure and leads to improved glucose metabolism. Diabetes Obes. Metab. 2017, 19, 496–508. [Google Scholar] [CrossRef]
  94. Plock, N.; Vollert, S.; Mayer, M.; Hanauer, G.; Lahu, G. Pharmacokinetic/Pharmacodynamic Modeling of the PDE4 Inhibitor TAK-648 in Type 2 Diabetes: Early Translational Approaches for Human Dose Prediction. Clin. Transl. Sci. 2017, 10, 185–193. [Google Scholar] [PubMed]
  95. Ookawara, M.; Nio, Y. Phosphodiesterase 4 inhibitors in diabetic nephropathy. Cell Signal. 2022, 90, 110185. [Google Scholar] [CrossRef] [PubMed]
  96. Muo, I.M.; MacDonald, S.D.; Madan, R.; Park, S.J.; Gharib, A.M.; Martinez, P.E.; Walter, M.F.; Yang, S.B.; Rodante, J.A.; Courville, A.B.; et al. Early effects of roflumilast on insulin sensitivity in adults with prediabetes and overweight/obesity involve age-associated fat mass loss–results of an exploratory study. Diabetes Metab. Syndr. Obes. 2019, 12, 743–759. [Google Scholar] [CrossRef] [PubMed]
  97. Gajendran, M.; Loganathan, P.; Catinella, A.P.; Jana, G.; Hashash, J.G. A comprehensive review and update on Crohn’s disease. Dis. Mon. 2018, 64, 20–57. [Google Scholar] [CrossRef]
  98. Reimund, J.-M.; Wittersheim, C.; Dumont, S.; Muller, C.D.; Kenney, J.S.; Baumann, R.; Poindron, P.; Duclos, B. Increased production of tumour necrosis factor-α, interleukin-1β, and interleukin-6 by morphologically normal intestinal biopsies from patients with Crohn’s disease. Gut 1996, 39, 684–689. [Google Scholar] [CrossRef]
  99. Reimund, J.M.; Dumont, S.; Muller, C.D.; Kenney, J.S.; Kedinger, M.; Baumann, R.; Poindron, P.; Duclos, B. In vitro effects of oxpentifylline on inflammatory cytokine release in patients with inflammatory bowel disease. Gut 1997, 40, 475–480. [Google Scholar] [CrossRef]
  100. Tetsi, L.; Charles, A.L.; Paradis, S.; Lejay, A.; Talha, S.; Geny, B.; Lugnier, C. Effects of cyclic nucleotide phosphodiesterases (PDEs) on mitochondrial skeletal muscle functions. Cell. Mol. Life Sci. 2017, 74, 1883–1893. [Google Scholar] [CrossRef]
  101. Arondel, Y.; Keravis, T.; Le Bec, A.; Baumann, R.; Duclos, B.; Reimund, J.M.; Lugnier, C. First characterisation of cyclic nucleotide phosphodiesterase isoforms in normal human mucosa and inflamed mucosa from Crohn’s disease patients. Gastroenterology 1999, 116, G3716. [Google Scholar]
  102. Banan, A.; Fitzpatrick, L.; Zhang, Y.; Keshavarzian, A. OPC-compounds prevent oxidant-induced carbonylation and depolymerization of the F-actin cytoskeleton and intestinal barrier hyperpermeability. Free Radic. Biol. Med. 2001, 30, 287–298. [Google Scholar]
  103. O’Mahony, S. Tetomilast. IDrugs 2005, 8, 502–507. [Google Scholar]
  104. Hartmann, G.; Bidlingmaier, C.; Siegmund, B.; Albrich, S.; Schulze, J.; Tschoep, K.; Eigler, A.; Lehr, H.A.; Endres, S. Specific type IV phosphodiesterase inhibitor rolipram mitigates experimental colitis in mice. J. Pharmacol. Exp. Ther. 2000, 292, 22–30. [Google Scholar] [PubMed]
  105. Videla, S.; Vilaseca, J.; Medina, C.; Mourelle, M.; Guarner, F.; Salas, A.; Malagelada, J.R. Selective inhibition of phosphodiesterase-4 ameliorates chronic colitis and prevents intestinal fibrosis. J. Pharmacol. Exp. Ther. 2006, 316, 940–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Spadaccini, M.; D’Alessio, S.; Peyrin-Biroulet, L.; Danese, S. PDE4 Inhibition and Inflammatory Bowel Disease: A novel therapeutic avenue. Int. J. Mol. Sci. 2017, 18, E1276. [Google Scholar] [CrossRef]
  107. Picchianti-Diamanti, A.; Spinelli, F.R.; Rosado, M.M.; Conti, F.; Laganà, B. Inhibition of Phosphodiesterase-4 in Psoriatic Arthritis and Inflammatory Bowel Diseases. Int. J. Mol. Sci. 2021, 22, 2638. [Google Scholar] [CrossRef]
  108. Oh, K.; Oh, E.-H.; Noh, S.M.; Park, S.H.; Kim, N.; Hwang, S.W.; Park, S.H.; Yang, D.-H.; Byeon, J.-S.; Myung, S.-J.; et al. Combined endoscopic and radiologic healing is associated with a better prognosis than endoscopic healing only in patients with Crohn’s disease receiving anti-TNF therapy. Clin. Transl. Gastroenterol. 2022, 13, e00442. [Google Scholar] [CrossRef] [PubMed]
  109. Gallagher, J.C. Advances in osteoporosis from 1970 to 2018. Menopause 2018, 25, 1403–1417. [Google Scholar] [CrossRef]
  110. Barrios-Moyano, A.; De la Peña-García, C. Prevalencia de osteoporosis y osteopenia en pacientes laboralmente activos. Acta Ortop. Mex. 2018, 32, 131–133. [Google Scholar]
  111. Zamani, M.; Zamani, V.; Heidari, B.; Parsian, H.; Esmaeilnejad-Ganji, S.M. Prevalence of osteoporosis with the World Health Organization diagnostic criteria in the Eastern Mediterranean Region: A systematic review and meta-analysis. Arch. Osteoporos. 2018, 13, 129. [Google Scholar] [CrossRef]
  112. Lugnier, C.; Stoclet, J.C.; Wilke, R. Phosphodiesterase isoenzymes in different tissues and the selective inhibition by denbufylline. Naunyn-Schmiedeberg’s Arch. Pharmacol. 1989, 339, R106. [Google Scholar]
  113. Nicholson, C.D.; Jackman, S.A.; Wilke, R. The ability of denbufylline to inhibit cyclic nucleotide phosphodiesterase and its affinity for adenosine receptors and the adenosine re-uptake site. Br. J. Pharmacol. 1989, 97, 889–897. [Google Scholar] [CrossRef]
  114. Miyamoto, K.; Waki, Y.; Horita, T.; Kasugai, S.; Ohya, K. Reduction of bone loss by denbufylline, an inhibitor of phosphodiesterase 4. Biochem. Pharmacol. 1997, 54, 613–617. [Google Scholar] [CrossRef]
  115. Waki, Y.; Horita, T.; Miyamoto, K.; Ohya, K.; Kasugai, S. Effects of XT-44, a phosphodiesterase 4 inhibitor, in osteoblastgenesis and osteoclastgenesis in culture and its therapeutic effects in rat osteopenia models. Jpn. J. Pharmacol. 1999, 79, 477–483. [Google Scholar] [PubMed]
  116. Kinoshita, T.; Kobayashi, S.; Ebara, S.; Yoshimura, Y.; Horiuchi, H.; Tsutsumimoto, T.; Wakabayashi, S.; Takaoka, K. Phosphodiesterase inhibitors, pentoxifylline and rolipram, increase bone mass mainly by promoting bone formation in normal mice. Bone 2000, 27, 811–817. [Google Scholar] [CrossRef]
  117. Boudjema, N. Inhibiteurs de phosphodiestrase type 4: Intérêt potentiel dans le traitement de l’ostéoporose. Master’s Thesis, Faculty of Pharmacy, Strasbourg, France, 15 March 2002. [Google Scholar]
  118. Yao, W.; Tian, X.Y.; Chen, J.; Setterberg, R.B.; Lundy, M.W.; Chmielzwski, P.; Froman, C.A.; Jee, W.S. Rolipram, a phosphodiesterase 4 inhibitor, prevented cancellous and cortical bone loss by inhibiting endosteal bone resorption and maintaining the elevated periosteal bone formation in adult ovariectomized rats. J. Musculoskelet. Neuronal Interact. 2007, 7, 119–130. [Google Scholar]
  119. Ahlström, M.; Pekkinen, M.; Huttunen, M.; Lamberg-Allardt, C. Cyclic nucleotide phosphodiesterases (PDEs) in human osteoblastic cells; the effect of PDE inhibition on cAMP accumulation. Cell. Mol. Biol. Lett. 2005, 10, 305–319. [Google Scholar]
  120. Ahlström, M.; Pekkinen, M.; Huttunen, M.; Lamberg-Allardt, C. Dexamethasone down-regulates cAMP-phosphodiesterase in human osteosarcoma cells. Biochem. Pharmacol. 2005, 69, 267–275. [Google Scholar] [CrossRef]
  121. Chen, L.; Zheng, T.; Park, H.; Noh, A.L.; Lee, J.M.; Lee, D.S.; Yim, M. PDE4 inhibitor suppresses PGE2-induced osteoclast formation via COX-2-mediated p27(KIP1) expression in RAW264.7 cells. Pharmazie 2011, 66, 201–206. [Google Scholar]
  122. Stern, A.R.; Yao, X.; Wang, Y.; Berhe, A.; Dallas, M.; Johnson, M.L.; Yao, W.; Kimmel, D.B.; Lane, N.E. Effect of osteoporosis treatment agents on the cortical bone osteocyte microenvironment in adult estrogen-deficient, osteopenic rats. Bone Rep. 2018, 8, 115–124. [Google Scholar] [CrossRef]
  123. Porwal, K.; Pal, S.; Bhagwati, S.; Siddiqi, M.I.; Chattopadhyay, N. Therapeutic potential of phosphodiesterase inhibitors in the treatment of osteoporosis: Scopes for therapeutic repurposing and discovery of new oral osteoanabolic drugs. Eur. J. Pharmacol. 2021, 899, 174015. [Google Scholar] [CrossRef]
  124. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  125. Kim, D.H.; Lerner, A. Type 4 cyclic adenosine monophosphate phosphodiesterase as a therapeutic target in chronic lymphocytic leukemia. Blood 1998, 92, 2484–2494. [Google Scholar] [CrossRef] [PubMed]
  126. Marko, D.; Romanakis, K.; Zankl, H.; Fürstenberger, G.; Steinbauer, B.; Eisenbrand, G. Induction of apoptosis by an inhibitor of cAMP-specific PDE in malignant murine carcinoma cells overexpressing PDE activity in comparison to their nonmalignant counterparts. Cell. Biochem. Biophys. 1998, 28, 75–101. [Google Scholar] [CrossRef] [PubMed]
  127. Narita, M.; Murata, T.; Shimizu, K.; Sugiyama, T.; Nakagawa, T.; Manganiello, V.C.; Tagawa, T. Phosphodiesterase 4 in osteoblastic osteosarcoma cells as a potential target for growth inhibition. Anticancer. Drugs 2003, 14, 377–381. [Google Scholar] [CrossRef]
  128. Favot, L.; Keravis, T.; Holl, V.; Le Bec, A.; Lugnier, C. VEGF-induced HUVEC migration and proliferation are decreased by PDE2 and PDE4 inhibitors. Thromb. Haemost. 2003, 90, 334–343. [Google Scholar] [CrossRef] [PubMed]
  129. Abusnina, A.; Keravis, T.; Zhou, Q.; Justiniano, H.; Lobstein, A.; Lugnier, C. Tumour growth inhibition and anti-angiogenic effects using curcumin correspond to combined PDE2 and PDE4 inhibition. Thromb. Haemost. 2015, 113, 319–328. [Google Scholar] [CrossRef] [PubMed]
  130. Suhasini, A.N.; Wang, L.; Holder, K.N.; Lin, A.P.; Bhatnagar, H.; Kim, S.W.; Moritz, A.W.; Aguiar, R.C.T. A phosphodiesterase 4B-dependent interplay between tumor cells and the microenvironment regulates angiogenesis in B-cell lymphoma. Leukemia 2016, 30, 617–626. [Google Scholar] [CrossRef] [PubMed]
  131. Parrella, E.; Gianni, M.; Cecconi, V.; Nigro, E.; Barzago, M.M.; Rambaldi, A.; Rochette-Egly, C.; Terao, M.; Garattini, E. Phosphodiesterase IV inhibition by piclamilast potentiates the cytodifferentiating action of retinoids in myeloid leukemia cells. Cross-talk between the cAMP and the retinoic acid signaling pathways. J. Biol. Chem. 2004, 279, 42026–42040. [Google Scholar] [CrossRef] [Green Version]
  132. Keravis, T.; Justiniano, H.; Guillemin, M.C.; de Thé, H.; Rochette-Egly, C.; Lugnier, C. cAMP-PDE activity, PDE4 activity and PDE4D protein expression are increased in RA-resistant NB4-R2 cells. In Proceedings of the EMBO Retinoids 2011: Mechanisms, Biology and pathology of Signaling by Retinoic acid and Retinoic acid Receptors, Strasbourg, France, 22–25 September 2011.Poster. [Google Scholar]
  133. Pleiman, J.K.; Irving, A.A.; Wang, Z.; Toraason, E.; Clipson, L.; Dove, W.F.; Dustin, A.; Deming, D.A.; Newton, M.A. The conserved protective cyclic AMP-phosphodiesterase function PDE4B is expressed in the adenoma and adjacent normal colonic epithelium of mammals and silenced in colorectal cancer. PLoS Genet. 2018, 14, e1007611. [Google Scholar] [CrossRef]
  134. Warrington, N.M.; Gianino, S.M.; Jackson, E.; Goldhoff, P.; Garbow, J.R.; Piwnica-Worms, D.; Gutmann, D.H.; Rubin, J.B. Cyclic AMP suppression is sufficient to induce gliomagenesis in a mouse model of Neurofibromatosis-1. Cancer Res. 2010, 70, 5717–5727. [Google Scholar] [CrossRef]
  135. Gong, S.; Chen, Y.; Meng. F.; Zhang. Y.; Wu, H.; Wu, F. Roflumilast restores cAMP/PKA/CREB signaling axis for FtMt-mediated tumor inhibition of ovarian cancer. Oncotarget 2017, 8, 112341–112353. [Google Scholar] [CrossRef]
  136. Henderson, D.J.P.; Houslay, M.D.; Bangma, C.H.; Hoffmann, R. Creating a potential diagnostic for prostate cancer risk stratification (InformMDx™) by translating novel scientific discoveries concerning cAMP degrading phosphodiesterase-4D7 (PDE4D7). Clin. Sci. 2019, 133, 269–286. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, D.U.; Nam, J.; Cha, M.D.; Kim, S.W. Inhibition of phosphodiesterase 4D decreases the malignant properties of DLD-1 colorectal cancer cells by repressing the AKT/mTOR/Myc signaling pathway. Oncol. Lett. 2019, 17, 3589–3598. [Google Scholar] [CrossRef] [PubMed]
  138. Kong, X.; Ai, G.; Wang, D.; Chen, R.; Guo, D.; Yao, Y.; Wang, K.; Liang, G.; Qi, F.; Liu, W.; et al. PDE4 and Epac1 Synergistically Promote Rectal Carcinoma via the cAMP Pathway. Anal. Cell. Pathol. 2019, 2019, 7145198. [Google Scholar] [CrossRef]
  139. Guo, C.H.; Bai, L.; Wu, H.H.; Yang, J.; Cai, G.H.; Wang, X.; Wu, S.X.; Ma, W. The analgesic effect of rolipram is associated with the inhibition of the activation of the spinal astrocytic JNK/CCL2 pathway in bone cancer pain. Int. J. Mol. Med. 2016, 38, 1433–1442. [Google Scholar] [CrossRef] [PubMed]
  140. Peng, T.; Gong, J.; Jin, Y.; Zhou, Y.; Tong, R.; Wei, X.; Bai, L.; Shi, J. Inhibitors of phosphodiesterase as cancer therapeutics. Eur. J. Med. Chem. 2018, 150, 742–756. [Google Scholar] [CrossRef] [PubMed]
  141. Reddy, S.P.; Martires, K.; Wu, J.J. The risk of melanoma and hematologic cancers in patients with psoriasis. J. Am. Acad. Dermatol. 2017, 76, 639–647. [Google Scholar] [PubMed]
  142. Hsien Lai, S.; Zervoudakis, G.; Chou, J.; Gurney, M.E.; Quesnelle, K.M. PDE4 subtypes in cancer. Oncogene 2020, 39, 3791–3802. [Google Scholar] [CrossRef]
  143. Chalasani, N.; Younossi, Z.; Lavine, J.E.; Charlton, M.; Cusi, K.; Rinella, M.; Harrison, S.A.; Brunt, E.M.; Sanyal, A.J. The diagnosis and management of nonalcoholic fatty liver disease: Practice guidance from the American Association for the Study of Liver Diseases. Hepatology 2018, 67, 328–357. [Google Scholar] [CrossRef]
  144. Younossi, Z.M.; Koenig, A.B.; Abdelatif, D.; Fazel, Y.; Henry, L.; Wymer, M. Global epidemiology of nonalcoholic fatty liver disease-Meta-analytic assessment of prevalence, incidence, and outcomes. Hepatology 2016, 64, 73–84. [Google Scholar] [CrossRef]
  145. Vonghia, L.; Van Herck, M.A.; Weyler, J.; Francque, S. Targeting myeloid-derived cells: New frontiers in the treatment of non-alcoholic and alcoholic liver disease. Front. Immunol. 2019, 10, 563. [Google Scholar] [CrossRef]
  146. Gobejishvili, L.; Barve, S.; Breitkopf-Heinlein, K.; Li, Y.; Zhang, J.; Avila, D.V.; Dooley, S.; McClain, C.J. Rolipram attenuates bile duct ligation-induced liver injury in rats: A potential pathogenic role of PDE4. J. Pharmacol. Exp. Pathol. 2013, 347, 80–90. [Google Scholar] [CrossRef] [PubMed]
  147. Ratziu, V.; Bedossa, P.; Francque, S.M.; Larrey, D.; Aithal, G.P.; Serfaty, L.; Voiculescu, M.; Preotescu, L.; Nevens, F.; De Lédinghen, V.; et al. Lack of efficacy of an inhibitor of PDE4 in phase 1 and 2 trials of patients with nonalcoholic steatohepatitis. Clin. Gastroenterol. Hepatol. 2014, 12, 1724–1730.e5. [Google Scholar] [CrossRef] [PubMed]
  148. Avila, D.V.; Barker, D.F.; Zhang, J.; McClain, C.J.; Barve, S.; Gobejishvili, L. Dysregulation of hepatic cAMP levels via altered Pde4b expression plays a critical role in alcohol-induced steatosis. J. Pathol. 2016, 240, 96–107. [Google Scholar] [CrossRef] [PubMed]
  149. Rodriguez, W.E.; Wahlang, B.; Wang, Y.; Zhang, J.; Vadhanam, M.V.; Joshi-Barve, S.; Bauer, P.; Cannon, R.; Ahmadi, A.R.; Sun, Z.; et al. Phosphodiesterase 4 inhibition as a therapeutic target for alcoholic liver disease: From bedside to bench. Hepatology 2019, 70, 1958–1971. [Google Scholar] [CrossRef]
  150. Hu, W.; Lu, T.; Chen, A.; Huang, Y.; Hansen, R.; Chandler, L.J.; Zhang, H.-T. Inhibition of phosphodiesterase-4 decreases ethanol intake in mice. Psychopharmacology 2011, 218, 331–339. [Google Scholar] [CrossRef]
  151. Tao, X.; He, H.; Peng, J.; Xu, R.; Fu, J.; Hu, Y.; Li, L.; Yang, X.; Feng, X.; Zhang, C.; et al. Overexpression of PDE4D in mouse liver is sufficient to trigger NAFLD and hypertension in a CD36-TGF-beta1 pathway: Therapeutic role of roflumilast. Pharmacol. Res. 2022, 175, 106004. [Google Scholar] [CrossRef]
  152. Gobejishvili, L.; Rodriguez, W.E.; Bauer, P.; Wang, Y.; Soni, C.; Lydic, T.; Barve, S.; McClain, C.; Maldonado, C. Novel liposomal rolipram formulation for clinical application to reduce emesis. Drug Des. Devel. Ther. 2022, 16, 1301–1309. [Google Scholar] [CrossRef]
  153. Malhi, G.S.; Mann, J.J. Depression. Lancet 2018, 392, 2299–2312. [Google Scholar] [CrossRef]
  154. Wachtel, H. Potential antidepressant activity of rolipram and other selective cyclic adenosine 3′,5′-monophosphate phosphodiesterase inhibitors. Neuropharmacology 1983, 22, 267–272. [Google Scholar] [CrossRef]
  155. O’Donnell, J.M. Antidepressant-like effects of rolipram and other inhibitors of cyclic adenosine monophosphate phosphodiesterase on behavior maintained by differential reinforcement of low response rate. J. Pharmacol. Exp. Ther. 1993, 264, 1168–1178. [Google Scholar]
  156. Barad, M.; Bourtchouladze, R.; Winder, D.G.; Golan, H.; Kandel, E. Rolipram, a type IV-specific phosphodiesterase inhibitor, facilitates the establishment of long-lasting long-term potentiation and improves memory. Proc. Natl. Acad. Sci. USA 1998, 95, e15020–e15025. [Google Scholar] [CrossRef] [PubMed]
  157. Zhang, H.T.; Huang, Y.; Jin, S.L.; Frith, S.A.; Suvarna, N.; Conti, M.; O’Donnell, J.M. Antidepressant-like profile and reduced sensitivity to rolipram in mice deficient in the PDE4D phosphodiesterase enzyme. Neuropsychopharmacology 2002, 27, 587–595. [Google Scholar] [PubMed]
  158. Menniti, F.S.; Faraci, W.S.; Schmidt, C.J. Phosphodiesterases in the CNS: Targets for drug development. Nat. Rev. Drug Discov. 2006, 5, 660–670. [Google Scholar] [CrossRef] [PubMed]
  159. Rutten, K.; Wallace, T.L.; Works, M.; Prickaerts, J.; Blokland, A.; Novak, T.J.; Santarelli, L.; Misner, D. L Enhanced long-term depression and impaired reversal learning in phosphodiesterase 4B- knockout (PDE4B−/−) mice. Neuropharmacology 2011, 61, 138–147. [Google Scholar]
  160. Omar, F.; Jane, E.; Findlay, J.E.; Carfray, G.; Allcock, R.W.; Jiang, Z.; Moore, C.; Muir, A.L.; Lannoy, M.; Fertig, B.A.; et al. Small-molecule allosteric activators of PDE4 long form cyclic AMP phosphodiesterases. Proc. Natl. Acad. Sci. USA 2019, 116, 13320–13329. [Google Scholar] [CrossRef] [PubMed]
  161. Campbell, S.L.; van Groen, T.; Kadish, I.; Smoot, L.H.M.; Bolger, G.B. Altered phosphorylation, electrophysiology, and behavior on attenuation of PDE4B action in hippocampus. BMC Neurosci. 2017, 18, 77. [Google Scholar] [CrossRef]
  162. Wang, Z.Z.; Yang, W.X.; Zhang, Y.; Zhao, N.; Zhang, Y.Z.; Liu, Y.Q.; Xu, Y.; Wilson, S.P.; O’Donnell, J.M.; Zhang, H.T.; et al. Phosphodiesterase-4D knock-down in the prefrontal cortex alleviates chronic unpredictable stress-induced depressive-like behaviors and memory deficits in mice. Sci. Rep. 2015, 5, 11332. [Google Scholar] [CrossRef]
  163. Vanmierlo, T.; Creemers, P.; Akkerman, S.; van Duinen, M.; Sambeth, A.; De Vry, J.; Uz, T.; Blokland, A.; Prickaerts, J. The PDE4 inhibitor roflumilast improves memory in rodents at non-emetic doses. Behav. Brain. Res. 2016, 303, 26–33. [Google Scholar] [CrossRef]
  164. Bolger, G.B. The PDE4 cAMP-specific phosphodiesterases: Targets for drugs with antidepressant and memory-enhancing action. Adv. Neurobiol. 2017, 17, 63–102. [Google Scholar]
  165. Tibbo, A.J.; Tejeda, G.S.; Baillie, G.S. Understanding PDE4′s function in Alzheimer’s disease; a target for novel therapeutic approaches. Biochem. Soc. Trans. 2019, 47, 1557–1565. [Google Scholar]
  166. Wang, H.; Zhang, M.; Xie, Q.; Yu, J.; Qi, Y.; Yue, Q. Identification of diagnostic markers for major depressive disorder by cross-validation of data from whole blood samples. PeerJ 2019, 7, e7171. [Google Scholar] [CrossRef] [PubMed]
  167. Wang, H.; Zhang, F.F.; Xu, Y.; Fu, H.R.; Wang, X.D.; Wang, L.; Chen, W.; Xu, X.Y.; Gao, Y.F.; Zhang, J.G.; et al. The Phosphodiesterase-4 Inhibitor Roflumilast, a Potential Treatment for the Comorbidity of Memory Loss and Depression in Alzheimer’s Disease: A Preclinical Study in APP/PS1 Transgenic Mice. Int. J. Neuropsychopharmacol. 2020, 23, 700–711. [Google Scholar]
  168. Blokland, A.; Heckman, P.; Vanmierlo, T.; Schreiber, R.; Paes, D.; Prickaerts, J. Phosphodiesterase type 4 inhibition in CNS diseases. TIPS 2019, 40, 971–985. [Google Scholar] [CrossRef]
  169. Lugnier, C.; Al-Kuraishy, H.M.; Rousseau, E. PDE4 inhibition as a therapeutic strategy for improvement of pulmonary dysfunctions in Covid-19 and cigarette smoking. Biochem. Pharmacol. 2021, 185, 114431. [Google Scholar] [PubMed]
  170. Angel, J.B.; Saget, B.M.; Walsh, S.P.; Greten, T.F.; Dinarello, C.A.; Skolnik, P.R.; Endres, S. Rolipram, a specific type IV phosphodiesterase inhibitor, is a potent inhibitor of HIV-1 replication. AIDS 1995, 9, 1137–1144. [Google Scholar] [PubMed]
  171. Navarro, J.; Punzon, C.; Jimenez, J.L.; Fernandez-Cruz, E.; Pizarro, A.; Fresno, M.; Muñoz-Fernández, M.A. Inhibition of phosphodiesterase type IV suppresses human immunodeficiency virus type 1 replication and cytokine production in primary T cells: Involvement of NF-κB and NFAT. J. Virol. 1998, 72, 4712–4720. [Google Scholar] [PubMed]
  172. Sun, Y.; Li, L.; Lau, F.; Beavo, J.A.; Clark, E.A. Infection of CD4+ memory T cells by HIV-1 requires expression of phosphodiesterase 4. J. Immunol. 2000, 165, 1755–1761. [Google Scholar] [CrossRef]
  173. Mata, M.; Martinez, I.; Melero, J.A.; Tenor, H.; Cortijo, J. Roflumilast inhibits respiratory syncytial virus infection in human differentiated bronchial epithelial cells. PLoS ONE 2013, 8, e69670. [Google Scholar] [CrossRef]
  174. Van Ly, D.; De Pedro, M.; James, P.; Morgan, L.; Black, J.L.; Burgess, J.K.; Oliver, B.G. Inhibition of phosphodiesterase 4 modulates cytokine induction from toll like receptor activated, but not rhinovirus infected, primary human airway smooth muscle. Respir. Res. 2013, 14, 127. [Google Scholar]
  175. Li, G.; Nunoya, J.I.; Cheng, L.; Reszka-Blanco, N.; Tsao, L.C.; Jeffrey, J.; Su, L. Regulatory T Cells Contribute to HIV-1 Reservoir persistence in CD4+ T cells through cyclic adenosine monophosphate–dependent mechanisms in humanized mice in vivo. J. Inf. Dis. 2017, 216, 1579–1591. [Google Scholar]
  176. Kuba, K.; Imai, Y.; Rao, S.; Gao, H.; Feng Guo, F.; Guan, B.; Huan, Y.; Yang, P.; Zhang, Y.; Deng, W.; et al. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS Coronavirus-Induced lung injury. Nat. Med. 2005, 11, 875–879. [Google Scholar] [CrossRef] [PubMed]
  177. Dalamaga, M.; Karampela, I.; Mantzoros, C.S. Commentary: Phosphodiesterase 4 inhibitors as potential adjunct treatment targeting the cytokine storm in COVID-19. Metabolism 2020, 109, 154282. [Google Scholar] [CrossRef] [PubMed]
  178. Chemboli, R.; Kapavarapu, R.; Deepti, K.; Prasad, K.R.S.; Reddy, A.G.; Kumar, A.V.D.N.; Rao, M.V.B.; Pal, M. Pyrrolo[2,3-b]quinoxalines in attenuating cytokine storm in COVID-19: Their sonochemical synthesis and in silico/in vitro assessment. J. Mol. Struct. 2021, 1230, 129868. [Google Scholar] [CrossRef] [PubMed]
  179. Bridgewood, C.; Damiani, G.; Sharif, K.; Watad, A.; Bragazzi, N.L.; Quartuccio, L.; Savic, S.; McGonagle, D. Rationale for evaluating PDE4 inhibition for mitigating against severe inflammation in COVID-19 pneumonia and beyond. Isr. Med. Assoc. J. 2020, 22, 335–339. [Google Scholar] [PubMed]
  180. Sugin Lal Jabaris, S.; Ranju, V. Scope of adjuvant therapy using roflumilast, a PDE-4 inhibitor against COVID-19. Pulm. Pharmacol. Ther. 2021, 66, 101978. [Google Scholar]
  181. Nguyen, H.O.; Schioppa, T.; Tiberio, L.; Facchinetti, F.; Villetti, G.; Civelli, M.; Del Prete, A.; Sozio, F.; Gaudenzi, C.; Passari, M.; et al. The PDE4 Inhibitor Tanimilast Blunts Proinflammatory Dendritic Cell Activation by SARS-CoV-2 ssRNAs. Front. Immunol. 2022, 12, 797390. [Google Scholar] [CrossRef]
  182. Geller, S.; Xu, H.; Lebwohl, M.; Nardone, B.; Lacouture, M.E.; Kheterpal, M. Malignancy risk and recurrence with psoriasis and its treatments: A concise update. Am. J. Clin. Dermatol. 2018, 19, 363–375. [Google Scholar] [CrossRef]
  183. Xu, B.; Qin, Y.; Li, D.; Cai, N.; Wu, J.; Jiang, L.; Jie, L.; Zhou, Z.; Xu, J.; Wang, H. Inhibition of PDE4 protects neurons against oxygen-glucose deprivation-induced endoplasmic reticulum stress through activation of the Nrf-2/HO-1 pathway. Redox Biol. 2020, 28, 101342. [Google Scholar] [CrossRef]
  184. Lugnier, C. PDE inhibitors: A new approach to treat metabolic syndrome? Curr. Opin. Pharmacol. 2011, 11, 698–706. [Google Scholar] [CrossRef]
  185. Klussmann, E. Protein–protein interactions of PDE4 family members–Functions, interactions and therapeutic value. Cell Signal. 2016, 28, 713–718. [Google Scholar] [CrossRef]
  186. Houslay, K.F.; Christian, F.; MacLeod, R.; Adams, D.R.; Houslay, M.D.; Baillie, G.S. Identification of a multifunctional docking site on the catalytic unit of phosphodiesterase-4 (PDE4) that is utilized by multiple interaction partners. Biochem. J. 2017, 474, 597–609. [Google Scholar] [CrossRef] [PubMed]
  187. Baillie, G.S.; Tejeda, G.S.; Kelly, M.P. Therapeutic targeting of 3′,5′-cyclic nucleotide phosphodiesterases: Inhibition and beyond. Nat. Drug Discov. 2019, 18, 770–796. [Google Scholar] [CrossRef] [PubMed]
  188. Viña, D.; Seoane, N.; Vasquez, E.C.; Campos-Toimil, M. cAMP compartmentalization in cerebrovascular endothelial cells: New therapeutic opportunities in Alzheimer’s disease. Cells 2021, 10, 1951. [Google Scholar] [CrossRef] [PubMed]
  189. Schick, M.A.; Schlegel, N. Clinical implication of phosphodiesterase-4-inhibition. Int. J. Mol. Sci. 2022, 23, 1209. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Intracellular signaling and cyclic nucleotide phosphodiesterases [2].
Figure 1. Intracellular signaling and cyclic nucleotide phosphodiesterases [2].
Ijms 23 10616 g001
Figure 2. Design of adenine R1 and R2 as PDE4 inhibitors [15] based on rolipram and trequinsin structures identified as PDE4 inhibitors, with trequinsin inhibiting both PDE3 and PDE4 [13].
Figure 2. Design of adenine R1 and R2 as PDE4 inhibitors [15] based on rolipram and trequinsin structures identified as PDE4 inhibitors, with trequinsin inhibiting both PDE3 and PDE4 [13].
Ijms 23 10616 g002
Figure 3. Structure of the PDE4 family. (A) Long forms of PDE4 contain both UCR1 and UCR2 and could be phosphorylated by PKA increasing PDE4 activity, whereas short forms lacking UCR1 could not be regulated by PKA and only exist as monomer [1]. (B) Diversity of PDE4D variants, Reprinted/adapted with permission from Ref. [24].
Figure 3. Structure of the PDE4 family. (A) Long forms of PDE4 contain both UCR1 and UCR2 and could be phosphorylated by PKA increasing PDE4 activity, whereas short forms lacking UCR1 could not be regulated by PKA and only exist as monomer [1]. (B) Diversity of PDE4D variants, Reprinted/adapted with permission from Ref. [24].
Ijms 23 10616 g003
Figure 4. PDE4 inhibitors molecular mechanisms [24,25,26].
Figure 4. PDE4 inhibitors molecular mechanisms [24,25,26].
Ijms 23 10616 g004
Figure 5. PDE3 and PDE4 cross-talk in vasodilation. NO or NO donor, by increasing soluble guanylyl cyclase, increases intracellular cGMP level. Therefore, this increase in cGMP induces preferentially PDE3 inhibition and consequently, an increase in cAMP in smooth muscle cells. In that way PDE4 might be able to hydrolyze cAMP in that way, PDE4 inhibitors inhibit PDE4 activity. In the presence of endothelial cells this effect is potentiated. Altogether, increases in cAMP altogether induces vasodilatation which might be potentiated in the presence of endothelium [38,39,40,41].
Figure 5. PDE3 and PDE4 cross-talk in vasodilation. NO or NO donor, by increasing soluble guanylyl cyclase, increases intracellular cGMP level. Therefore, this increase in cGMP induces preferentially PDE3 inhibition and consequently, an increase in cAMP in smooth muscle cells. In that way PDE4 might be able to hydrolyze cAMP in that way, PDE4 inhibitors inhibit PDE4 activity. In the presence of endothelial cells this effect is potentiated. Altogether, increases in cAMP altogether induces vasodilatation which might be potentiated in the presence of endothelium [38,39,40,41].
Ijms 23 10616 g005
Figure 6. Increase in cGMP in human endothelial cells related to PDE4 inhibition [47].
Figure 6. Increase in cGMP in human endothelial cells related to PDE4 inhibition [47].
Ijms 23 10616 g006
Figure 7. PDE4, a therapeutic target in various human diseases.
Figure 7. PDE4, a therapeutic target in various human diseases.
Ijms 23 10616 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lugnier, C. The Complexity and Multiplicity of the Specific cAMP Phosphodiesterase Family: PDE4, Open New Adapted Therapeutic Approaches. Int. J. Mol. Sci. 2022, 23, 10616. https://doi.org/10.3390/ijms231810616

AMA Style

Lugnier C. The Complexity and Multiplicity of the Specific cAMP Phosphodiesterase Family: PDE4, Open New Adapted Therapeutic Approaches. International Journal of Molecular Sciences. 2022; 23(18):10616. https://doi.org/10.3390/ijms231810616

Chicago/Turabian Style

Lugnier, Claire. 2022. "The Complexity and Multiplicity of the Specific cAMP Phosphodiesterase Family: PDE4, Open New Adapted Therapeutic Approaches" International Journal of Molecular Sciences 23, no. 18: 10616. https://doi.org/10.3390/ijms231810616

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

Article Metrics

Back to TopTop