New Therapies Targeting Cystogenesis in Autosomal Polycystic Kidney Disease - European Medical Journal

New Therapies Targeting Cystogenesis in Autosomal Polycystic Kidney Disease

12 Mins
Nephrology
Download PDF
Authors:
*Maurizio Salvadori,1 Aris Tsalouchos2
Disclosure:

The authors have declared no conflicts of interest.

Acknowledgements:

The authors acknowledge Dr Stefanos Tsalouchos for his assistance and for drawing Figure 1.

Received:
24.02.17
Accepted:
23.05.17
Citation:
EMJ Nephrol. ;5[1]:102-111. DOI/10.33590/emjnephrol/10310690. https://doi.org/10.33590/emjnephrol/10310690.
Keywords:
Adult autosomal polycystic kidney disease (ADPKD), cystogenesis, vasopressin 2 receptors, somatostatin analogues, mammalian target of rapamycin (mTOR) signalling

Each article is made available under the terms of the Creative Commons Attribution-Non Commercial 4.0 License.

Abstract

Autosomal dominant polycystic kidney disease is the most common inherited kidney disease and results from mutations in the polycystin 1 gene (PKD1) or the polycystin 2 gene (PKD2). The disease is characterised by the progressive development of fluid-filled cysts derived from renal tubular epithelial cells that destroy the architecture of the renal parenchyma and lead to kidney failure. Until recently, the causes and the molecular pathways that lead to cystogenesis remained obscure. In the last decade, enormous progress has been made in understanding the pathogenesis of autosomal dominant polycystic kidney disease and developing  new therapies. The purpose of this review is to provide an update on the promising therapies that are being developed and tested, based on knowledge of recent advances in molecular and cellular targets involved in cystogenesis.

INTRODUCTION

Autosomal dominant polycystic kidney disease (ADPKD) affects 1:400–1:1,000 of live births and is the most common monogenic inherited form of kidney disease across all ethnic types. ADPKD is characterised by cyst formation and enlargement in the kidney and other organs. It represents the fourth leading global cause for kidney failure, and end-stage renal disease (ESRD) usually occurs by late middle age, requiring renal replacement therapy in ~50% of patients by 70 years of age.1,2

In 85% of cases, ADPKD occurs as a result of germ-line mutation in the polycystin 1 gene (PKD1), while in 15% of cases it is due to a germline mutation in the polycystin 2 gene (PKD2).3 Recently, a polycystic kidney and/or polycystic liver disease-3 (PKD3) caused by heterozygous mutation in the gene encoding for glucosidase II subunit-a (GANAB) has been described.4 Polycystin-1 (PC1) and polycystin-2 (PC2) interact with each other through their C-terminal cytoplasmic domains and are known to form a complex that functions as a transient receptor potential channel involved in the regulation of intracellular calcium homeostasis.5,6 Glucosidase II subunit-a is required for maturation and surface and ciliary localisation of PC1 and PC2.

Analysis of GANAB-null human renal cells resulted in absence of the mature N-terminal PC1 but full-length PC1 and PC2. Heterozygous-null CANAB renal cells had a 50% depletion of mature N-terminal PC1.4 On average, patients with mutations in PKD1 developed ESRD at younger ages.7

Cystogenesis follows a two-hit model. ADPKD is recessive at the cellular level and cysts develop clonally from a tubular cell only after the cell has acquired a second, somatic mutation to inactivate the remaining normal allele.8 Although the exact mechanisms of cystogenesis remain to be elucidated, the pathological processes that facilitate cyst enlargement are probably a result of two specific abnormalities: i) increased fluid secretion into the cyst lumen, and ii) inappropriately increased cell division by the epithelium lining the cyst.9 The major signalling pathways implicated in these phenotypic changes include the intracellular deregulation of calcium homeostasis, cyclic adenosine monophosphate (cAMP) accumulation and activation of protein kinase A (PKA), activation of mitogen activated protein and mammalian target of rapamycin (mTOR) kinases, and other intracellular signalling mechanisms.10-12

Until recently, the treatment of ADPKD was aimed at the management of secondary conditions, particularly hypertension, to limit morbidity and mortality after the disease becomes symptomatic. Recent developments arising from a better mechanistic understanding of the molecular pathways involved in cyst growth have allowed targeting the disease pathogenesis, rather than the disease complications. The current review focusses on these novel therapeutic approaches that interfere with the molecular pathways of cystogenesis (Figure 1).

Figure 1: Illustration of the key mechanisms of adult autosomal polycystic kidney disease pathogenesis and targets of potential treatments.
Polycystin-1 and polycystin-2 expressed in different subcellular locations and regulate 1) proliferation, 2) fluid secretion, 3) ciliary function, 4) cell-cell adhesion, and 5) cell-matrix interaction of renal epithelial cells. Dysfunction of polycystin-1 or polycystin-2 results to aberrant signalling pathways including: A) activation of cAMP, B) decreased intracellular calcium concentrations, and C) activation of mTOR.
The targets of candidate drugs are depicted as grey circles.
CFTR: cystic fibrosis transmembrane regulator; ER: endoplasmic reticulum; ERK: extracellular-signal
regulated kinase; GlcCer: glucosylceramide; HDAC: histone deacetylase; IL-6R: interleukin-6 receptor; MEK: mitogen activated protein kinase; mTOR: mammalian target of rapamycin; PC: polycystin;
PDE: phosphodiesterase; PKA: protein kinase A; SR: somatostatin receptor; TSC: tuberous
sclerosis; V2R: vasopressin V2 receptor; EGFR: estimated glomerular filtration rate; cAMP: cyclic
adenosine monophosphate.

DRUGS TARGETING cAMP-DEPENDENT CYSTIC EXPANSION

Role of cAMP in Cystogenesis

In ADPKD, disruption of intracellular Ca2+ homeostasis due to mutations in the PKD gene leads to low intracellular calcium and consequently increased levels of intracellular cAMP. Normally, the levels of cAMP are controlled by a balanced activity of membrane-bound and soluble isoforms of adenylate cyclase (AC) (which catalyses the formation of cAMP from ATP) and phosphodiesterases (which degrades cAMP to AMP). Decreased intracellular calcium inhibits the activity of phosphodiesterases and activates ACs, thus producing a net increase in cAMP concentration.10 cAMP exerts its effects via PKA, which phosphorylates a number of metabolic enzymes and promotes transepithelial fluid secretion. Chloride secretion drives sodium into the cystic cavity through paracellular mechanisms; this causes movement of water through aquaporins and cyst expansion.13 In addition, in ADPKD, cAMP promotes cyst enlargement by stimulating epithelial cell proliferation, primarily through the activation of the B-Raf/MEK/ERK pathway.14,15

Vasopressin 2 Receptor Antagonists

Normally, vasopressin (AVP) is secreted into the circulation by the posterior pituitary gland, in response to an increase in serum osmolality or a decrease in effective circulating volume. In the kidney, AVP binds to the V2 AVP receptor in the basolateral membranes of collecting-duct cells. The V2 receptor is a typical member of the large superfamily of G protein–coupled receptors. Thus, occupancy of this receptor results in mediated activation of AC and the formation of cAMP with subsequent activation of PKA, which promotes the fusion of cytoplasmic vesicles containing aquaporin-2 water-channel proteins with the apical membrane. As a result, this normally water-tight membrane becomes water-permeable. Driven by the osmotic gradient of sodium, water is then transcellularly reabsorbed, entering the cells through aquaporin-2 in the apical membrane and leaving the cells for the interstitial space through aquaporin-3 and aquaporin-4, which reside in the basolateral membrane.16 In patients with ADPKD, there is a pathologically hyperactive AVP/V2 receptor system. Serum concentrations of AVP correlate positively with both serum osmolality, as well as with total kidney size and negatively with glomerular filtration rate (GFR).17-20 The central role of cAMP in cystogenesis and the pathologically hyperactive AVP/V2 receptor system have made the blocking of V2R particularly appealing in the treatment of ADPKD.

In preclinical trials, a non-peptide AVP antagonist mozavaptan (OPC-31260), administered in murine cystic models orthologous to human disease, including the Pkd2WS25/− mouse (ADPKD), PCK rat (ARPKD), and pcy mouse (nephronophthisis Type 3), reduced renal cAMP and inhibited disease progression, as measured by the reduction in kidney volume, the cystic area, the number of mitotic and apoptotic cells, and the blood urea nitrogen.21-23 Additional studies were conducted to examine the effects of tolvaptan (OPC-4106), a more potent and highly selective human V2R antagonist, in comparison with mozavaptan.24 Tolvaptan showed similar results on renal cAMP and PKD progression in the PCK rat model using the lowest dose. Reif et al.25 in an in vitro study examined the effect of tolvaptan on intracellular cAMP, ERK activity, cell proliferation, and transcellular chloride anion secretion using human ADPKD cyst epithelial cells. Tolvaptan caused inhibition of cAMP AVP-induced production, ERK signalling AVP-induced, cell proliferation, and chloride anion secretion. These effects significantly contributed to decreased in vitro cyst growth.

The large randomised, double-blind, placebo-controlled, multinational, Phase III TEMPO 3:4 trial26 confirmed the aforementioned experimental studies. This trial enrolled 1,445 patients aged 18–50 years with ADPKD, rapidly progressive kidney growth (total kidney volume [TKV] ≥750 mL) as measured by magnetic resonance imaging (MRI) and chronic kidney disease (CKD) Stages 1–3. Tolvaptan reduced the rate of TKV growth (primary endpoint) by 49% and the rate of estimated GFR (eGFR) loss on treatment (secondary endpoint) by 26% per year during the median observation period of 3 years. The effect on TKV appeared greater during the first year of treatment than during the second or third years. Beneficial effects on renal function have been observed in all patient subgroups, especially in patients aged ≥35 years and in patients with hypertension or a TKV of ≥1,500 mL. Another important secondary endpoint was the reduction in kidney pain occurring early and throughout treatment. The results of the TEMPO 3:4 trial suggested that tolvaptan had no effect compared with placebo on albuminuria. Conversely, in a post hoc exploratory analysis, tolvaptan decreased albuminuria compared with placebo independently of blood pressure. In addition, the treatment efficacy of tolvaptan on changes in TKV and eGFR was more readily detected in patients with higher albuminuria.27

Based on the results of the TEMPO 3:4 trial, tolvaptan has been approved to delay the progression of ADPKD in patients with a rapid increase of TKV in Japan, Canada, the European Union (EU), the UK, and South Korea. The European Renal Association-European Dialysis and Transplant Association (ERA-EDTA)28 has issued detailed guidance on this topic. ERA-EDTA suggests that tolvaptan can be prescribed to adult ADPKD patients aged <50 years with CKD Stages 1–3a (eGFR >45 mL/min/1.73 m2) who have demonstrated, or who are likely to have, rapidly progressing disease. ERA-EDTA recommends not starting tolvaptan in patients aged 30–40 years with CKD Stage 1 (eGFR >90 mL/min/1.73 m2) or patients aged 40–50 years with CKD Stages 1 or 2 (eGFR >60 mL/min/1.73 m2). The organisation recommends that rapid disease progression be defined as a confirmed annual eGFR decline of ≥5 mL/min/1.73 m2 in 1 year, and/or 2.5 mL/min/ 1.73 m2 per year over a period of 5 years. It can also be defined as a >5% increase in TKV per year by repeated measurements (preferably 3, each at least 6 months apart and by MRI). With regard to dosing, ERA-EDTA suggests that tolvaptan be started with a dose of 45 mg in the morning and 15 mg in the evening, uptitrating the dose to 50/30 and 90/30 when tolerated, and discontinuing tolvaptan when patients approach ESRD.

Tolvaptan has significant adverse effects including aquaretic effects (polyuria, nocturia, and polydipsia) and elevation of aminotransferase enzyme concentrations with the potential for acute liver failure.26,29,30 Although the incidence of hepatic enzyme elevation was low, three patients treated with tolvaptan in the TEMPO 3:4 trial met criteria for Hy’s law (elevation of aminotransferase enzymes >3-times the upper limit of normal, plus elevation of total bilirubin >2-times the upper limit of normal, without other explanatory mechanisms), which have a fatality rate of ˜10% from liver injury. Therefore, the Cardiovascular and Renal Drugs Advisory Committee of the US Food and Drug Administration (FDA) declined to approve tolvaptan for ADPKD, as they were worried that liver damage might progress with long use treatment.31 Appropriate patient selection is critical to optimise long-term benefits and minimise adverse effects and hepatotoxic risk factors.

Studies to further assess the efficacy and tolerability of tolvaptan in patients with ADPKD are ongoing or just completed. TEMPO 4:4 is a 2-year, open-label extension of TEMPO 3:4 and was completed in March 2016. This study aimed to evaluate the long-term efficacy and safety of tolvaptan in patients with ADPKD; the findings are soon to be published. In addition, the long-term safety of titrated tolvaptan in patients with ADPKD is currently being assessed in a Phase III open label trial,32 while the ongoing Phase IIIb REPRISE trial aims to extend the understanding of the efficacy and safety of tolvaptan in patients with late Stage 2–early Stage 4 CKD.33

Somatostatin Analogues

Somatostatin (SST) is an endogenous hormone primarily secreted by the pancreatic islet δ-cells. SST has anti-secretory and anti-proliferative effects mediated by the interaction with five subtypes of G protein–coupled receptors (SSTR1-5).34 SST receptors are expressed by renal tubular epithelial cells and by cholangiocytes. SST selectively inhibits cAMP synthesis in the epithelial cells of the distal tubules and collecting ducts both in vitro and in vivo35,36 and exerts similar effects to cholangiocytes.37 As plasma half-life of the native SST is very short (1–3 minutes), the synthetic analogues octreotide, lanreotide, and pasireotide were developed for use in clinical practice.

In particular, octreotide and lanreotide have a half-life of 2 hours and present a high affinity for SSTR2 and SSTR3 and moderate affinity for SSTR5. By comparison, pasireotide has high affinity for all the receptors of SST, except SSTR4, and its plasma half-life is about 12 hours.38 Currently, formulations of octreotide and lanreotide with long-acting release (LAR), which allows the administration every 28 days intramuscularly or intradermally, have been introduced into clinical practice. Ruggenenti et al.39 have evaluated for the first time the effectiveness of octreotide-LAR by performing a randomised, crossover, placebo-controlled trial in 14 ADPKD patients, which demonstrated the potential efficacy in slowing the growth of TKV and the relative safety of the treatment. Van Keimpema et al.40 compared the effects of 6 months of treatment with lanreotide or placebo in 54 patients with polycystic liver disease (PLD), including 32 with ADPKD and the remaining with isolated PLD. The average volume of the liver decreased in patients treated with lanreotide while it increased in the placebo group. Moreover, in patients with ADPKD, TKV was reduced after treatment with lanreotide, while it was increased in the placebo group. In a subsequent open label extension study41 patients who participated in the initial trial were re-enrolled to complete a treatment period of 12 months with lanreotide. Liver volume decreased after 12 months of treatment with lanreotide, with the greatest effect seen during the first 6 months. In the 25 patients with ADPKD, TKV remained stable at the end of 12 months. In another 12-month study, 42 patients with PLD, including 34 with ADPKD, were randomised to receive treatment with octreotide-LAR or placebo.42 The total volume of the liver was reduced in the treatment arm with octreotide-LAR but increased in the placebo group. In patients with ADPKD, the TKV remained unchanged in the octreotide-LAR group but increased in the placebo group. In addition, renal function had a slower reduction in patients treated with octreotide-LAR, although the difference did not reach statistical significance. More recently in the ALADIN multicentre study conducted in Italy, 79 patients with ADPKD and eGFR >40 mL/min/1.73 m2 were randomised to a 3-year treatment with octreotide-LAR or placebo.43 After the first year, the average increase in TKV was significantly lower in patients treated with octreotide-LAR compared to those receiving placebo. In the third year, the average increase in TKV in the treatment arm was lower than the placebo group without reaching statistical significance. During the entire study period, the annual reduction in GFR was lower in the octreotide-LAR group than in the placebo group, although the difference did not reach statistical significance. A more recent open label clinical study evaluated the efficacy of 6 months of treatment with lanreotide in 43 patients with symptomatic PLD and ADPKD (eGFR >30 mL/min/1.73 m2).44 Compared to baseline, the median liver volume decreased significantly, as well as that of the kidney. In addition, renal function remained stable until the end of the study. A recent meta-analysis confirmed the efficacy of SST analogues in reducing the progressive increase of TKV on average, with a reduction of 9% compared to the growth observed in patients treated with placebo or conventional therapies. However, treatment with SST analogues did not demonstrate significant effects on the eGFR.45

Based on these studies, in August 2015 the European Medicines Agency (EMA) has attributed to lanreotide the ‘orphan drug’ designation for the treatment of ADPKD. Designated orphan medicinal products are products that are still under investigation and are considered for orphan designation on the basis of potential activity. Opinions on orphan medicinal product designations are based on the following three criteria: i) the seriousness of the condition; ii) the existence of alternative methods of diagnosis, prevention, or treatment; and iii) either the rarity of the condition (affecting no more than 5 in 10,000 people in the EU) or insufficient returns on investment.

In the studies mentioned above, treatment with SST analogues was generally well tolerated with no particular problems, diarrhoea being the most common adverse event. However, recently, the authors of a randomised, controlled trail documented an increased risk for hepatic cyst infection during lanreotide treatment. A literature review also suggested an increased risk for hepatic cyst infection during the use of SST analogues.46

Additional clinical trials of SST analogues for ADPKD and/or PLD are currently ongoing.47-50

DRUGS TARGETING THE mTOR SIGNALLING PATHWAY

Role of the mTOR Signalling Pathway in Cystogenesis

Serine/threonine-protein kinase mTOR is an enzyme that plays a critical role in proliferation and cell growth.51 The first suggestion of a prominent role of the mTOR pathway in the pathogenesis of ADPKD comes from studies in patients with severe infantile-onset of ADPKD due to a large deletion of chromosome 16 involving the PKD1 gene [16p13.3] and the adjacent tuberous sclerosis 2 (TSC2) genes [16p13.3].52 TSC1 and TSC2 encode for hamartin and tuberin, respectively. These two proteins together with TBC1 domain family member 7 (TBC1D7) form the TSC protein complex that acts as a critical negative regulator of mTOR complex 1 (mTORC1).53 PC1 has also an important function in the regulation of the mTOR pathway, as the C-terminal cytoplasmic tail of PC1 interacts with tuberin. In ADPKD this interaction is impaired, and the mTOR pathway is inappropriately activated in cyst-lining epithelial cells of human ADPKD patients and mouse models.54 Based on these data, a possible therapeutic role for mTOR inhibitors in ADPKD has been suggested.

mTOR Inhibitors

Sirolimus and its derivative everolimus, used in maintenance immunosuppression in patients undergoing kidney transplantation, have been proposed as potential new drugs to slow the growth of cysts and the progression of ADPKD in ESRD. The effects of treatment with mTOR inhibitors have been assessed in different experimental models of ADPKD.55-58 The first published randomised double-blind study compared the effects of 2 years of treatment with everolimus (5 mg/day) or placebo in 433 patients with ADPKD and GFR >30 mL/min/1.73 m2.59 During the first year of study, the increase of TKV was significantly lower in the treatment arm with everolimus compared to placebo. This effect was not confirmed at the end of the second year. In addition, the initial effectiveness of everolimus in slowing TKV did not translate into improvement of renal function.

The SUISSE study compared the effects of treatment for 18 months with sirolimus (2 mg/day) or conventional therapy in 100 patients with ADPKD and GFR 70 mL/min/1.73 m2.60 The median increase in TKV was comparable between the two groups as well as the eGFR throughout the entire study period. The randomised trial SIRENA compared the effects of treatment with sirolimus or with conventional therapy alone for 6 months in 21 patients with ADPKD and GFR 40 mL/min/1.73 m2.61 The treatment with sirolimus was associated with a minor increase of the TKV compared to conventional therapy. In a subsequent open label study (RAPYD), 55 patients with ADPKD and mild-to-moderate renal impairment were randomised to 24 months of treatment with ramipril (control group), ramipril in combination with high doses of sirolimus (target blood levels: 6–8 ng/mL) or ramipril in combination with low-dose sirolimus (target blood levels: 2–4 ng/mL).62 Compared to baseline, total cyst volume decreased significantly in both treatment arms with sirolimus, while increasing in the control group. In a more recent study, 30 patients with ADPKD and measured GFR ≥25 mL/min/1.73 m2 were randomised to receive low-dose sirolimus (target blood levels 2–5 ng/mL), standard doses of sirolimus (target blood levels >5–8 ng/mL) or conventional therapy for 12 months.63 TKV did not change significantly in the two treatment groups with sirolimus as in the group assigned to conventional therapy. In addition, the renal function improved (GFR measured by plasma clearance of iothalamate) with low-dose sirolimus but not with the standard dose of the drug. Currently, there are two ongoing trials that are testing mTOR inhibitors in ADPKD.64,65

OTHER THERAPEUTIC TARGETS IN PRECLINICAL STUDIES AND IN EARLY CLINICAL TRIALS

In addition, other agents targeting different molecules or pathways involved in cystogenesis have been used in preclinical studies and some of them are ongoing in early clinical trials in humans.66 All these agents are summarised in Table 1.

Table 1: New agents in preclinical models or in early clinical trials in humans.
CFTR: cystic fibrosis transmembrane conductance regulator; PPQs: pyrimido-pyrrolo-quinoxalinediones; AMP: adenosine monophosphate.

Bosutinib (SKI-606) is a Src/Abl tyrosine kinase inhibitor effective in inhibiting epithelial cell proliferation and reducing extracellular matrix adhesion. In the BPK and PCK rodent models of ADPKD, bosutinib was found to suppress kidney cyst formation by inhibiting epidermal growth factor receptor activation and downregulating B-RAF/ERK signalling.67 Based on this evidence, a Phase II, multicentre, randomised, double-blind, placebo-controlled clinical trial with bosutinib68 has been conducted and completed, and we currently are expecting publication of the study results. Tesevatinib, a new tyrosine kinase inhibitor, is being currently evaluated in two ongoing trials.69,70

Triptolide, by acting as a PC2 agonist to restore cytosolic Ca2+ release, has been effective in arresting cellular proliferation and attenuating overall cyst formation in Pkd1−/− murine kidney epithelial cells.71 A clinical trial conducted in China has been terminated due to a high rate of drop-outs72 and another trial is underway.73

In an experimental model, sorafenib, a non-selective RAF inhibitor, reduced the basal activity of ERK, inhibited cAMP-dependent activation of B-RAF and MEK/ERK signalling, and caused a concentration dependent inhibition of cell proliferation induced by cAMP and EGF. In addition, it completely blocked in vitro cyst growth of human ADPKD cystic cells.74 A different Raf inhibitor (PLX5568) has been evaluated in the Han:SPRD rat model.75 In this study, cyst enlargement attenuated without an improvement in kidney function. Furthermore, the authors reported increased renal and liver fibrosis.

A preclinical study with the CDK inhibitor (R)-Roscovitine in juvenile cystic kidney and congenital polycystic kidney mouse models of PKD effectively attenuated cystogenesis by inhibiting cell cycle progression, proliferation, and apoptosis.76 In addition, a more potent second-generation analogue of roscovitine (S-CR8) showed effective inhibition of both renal and hepatic cystogenesis in an orthologous mouse model of ADPKD with inactivated PKD1 gene.77

Altered expression of histone deacetylases (HDCAs) causes abnormal transcription of key genes controlling principal cellular functions such as cell proliferation, cell-cycle regulation, and apoptosis.78 A pan-HDAC inhibitor called trichostatin A (TSA) has been evaluated in a PKD2 zebrafish model showing the ability to suppress pronephric cyst formation.79 The same results have been obtained after administration of valproic acid (VPA), a Class I HDAC inhibitor.79 The NIAC-PKD1 trial80 has just been completed, and we are currently awaiting publication of the results. In addition, another trial, NIAC-PKD2,81 is currently recruiting participants.

Each of the three chemical classes of cystic fibrosis transmembrane conductance regulator (CFTR) inhibitors has been tested in PKD models: i) thiazolidinones, ii) glycine and malonic acid hydrazides, and iii) pyrimido-pyrrolo-quinoxalinediones. The best thiazolidinone, tetrazolo-CFTRinh-172, and the best glycine hydrazide, Ph-GlyH-101, were found to inhibit cyst formation and enlargement in Madin–Darby canine kidney cyst models and in PKD1 mice.82

AMP-activated protein kinase regulates cell growth via suppression of the mTORC1 pathway, by direct phosphorylation of the tumour suppressor TSC2 and Raptor (regulatory associated protein of mTOR).83 Recently, Takiar et al.84 showed in a PKD1 mice model, that metformin inhibited renal cystogenesis and caused a significant decrease in the cystic index by activating AMP-activated protein kinase and suppressing mTOR and CFTR. Currently, the TAME trial85 is recruiting ADPKD patients to see if metformin is safe and well tolerated.

Agonists of peroxisome proliferator-activated receptor gamma (PPAR-γ) have been shown to have anti-cystogenic properties in PKD animal models. Pioglitazone has been shown to inhibit the growth of renal and hepatic cysts in PCK rats by inhibiting the CFTR-mediated ionic current and the secretion of fluid.86 In another study, pioglitazone also reduced cellular proliferation, highlighted by a reduction in the number of cells positive for Ki67 (a proliferation marker) in the dilated tubules and in cysts from treated rats.87 Another powerful agonist of PPAR-γ, rosiglitazone, has been used to treat Han:SPRD rats. Rosiglitazone delayed the onset of renal failure but was associated with cardiac enlargement due to excessive renal sodium reabsorption.88 Based on these preclinical data, the PIOPKD trial89 is currently recruiting participants.

CONCLUSION AND FUTURE DIRECTIONS

Multiple signalling pathways are involved in cyst formation and progression, and studies of these signalling pathways have led to potential treatments for ADPKD. In this review, we have covered the successes obtained in recent years in understanding the pathogenesis of ADPKD and presented novel therapeutic strategies targeting molecular pathways of cystogenesis. V2R antagonists and SST analogues have been shown to safely slow kidney growth and protect renal function in patients with ADPKD, and represent the most well-characterised and promising candidate therapies to date. According to the results of the TEMPO3/4 study and registration by the EMA, tolvaptan seems to be the first choice drug. Some medical interventions successful in experimental models failed in clinical practice and others still need to be evaluated in clinical trials. It is possible that monotherapy may not be sufficient and that targeting multiple molecular pathways will be required to retard cyst growth and disease progression in the feature. Combination therapy is then the right direction of further clinical trials in order to find effective treatment. To date no association of drugs inhibiting cystogenesis has proven to be effective: further research is needed. The association of tolvaptan and angiotensin-converting-enzyme (ACE) inhibitors, angiotensin receptor blockers (ARB), and statins are reported as effective associations in a recent study.90

References
Chebib FT, Torres VE. Autosomal Dominant Polycystic Kidney Disease: Core Curriculum 2016. Am J Kidney Dis. 2016;67(5):792-810. Torres VE, Harris PC. Autosomal dominant polycystic kidney disease: the last 3 years. Kidney Int. 2009;76(2):149-68. Rysz J et al. Combination drug versus monotherapy for the treatment of autosomal dominant polycystic kidney disease. Expert Opin Pharmacother. 2016; 17(15):2049-56. Porath B et al. Mutations in GANAB, Encoding the Glucosidase IIa Subunit, Cause Autosomal-Dominant Polycystic Kidney and Liver Disease. Am J Hum Gen. 2016;98(6):1193-207. Vassilev PM et al. Polycystin-2 is a novel cation channel implicated in defective intracellular Ca(+) homeostasis in polycystic kidney disease. Biochem Biophys Res Commun. 2001;282(1):341-50. Anyatonwu GI, Ehrlich BE. Organic cation permeation through the channel formed by polycystin-2. J Biol Chem. 2005;280(33):29488-93. LaRiviere WB et al. Novel therapeutic approaches to autosomal dominant polycystic kidney disease. Transl Res. 2015;165(4):488-98. Belibi FA, Edelstein CL. Novel targets for the treatment of autosomal dominant polycystic kidney disease. Expert Opin Investig Drugs. 2010;19(3):315-28. Mochizuki T et al. Autosomal dominant polycystic kidney disease: recent advances in pathogenesis and potential therapies. Clin Exp Nephrol. 2013;17(3):317-26. Torres VE, Harris PC. Strategies targeting cAMP signaling in the treatment of polycystic kidney disease. J Am Soc Nephrol. 2014;25(1):18-32. Yamaguchi T et al. Calcium restores a normal proliferation phenotype in human polycystic kidney disease epithelial cells. J Am Soc Nephrol. 2006;17(1):178-87. Paavola J et al. Polycystin-2 mutations lead to impaired calcium cycling in the heart and predispose to dilated cardiomyopathy. J Mol Cell Cardiol. 2013;58:199-208. Hanaoka K et al. A role for CFTR in human autosomal dominant polycystic kidney disease. Am J Physiol. 1996;270 (1 Pt 1):C389-99. Yamaguchi T et al. cAMP stimulates the in vitro proliferation of renal cyst epithelial cells by activating the extracellular signal regulated kinase pathway. Kidney Int. 2000;57(4):1460-71. Aguiari G et al. Polycystin-1 regulates amphiregulin expression through CREB and AP1 signalling: implications in ADPKD cell proliferation. J Mol Med (Berl). 2012;90(11):1267-82. Knepper MA et al. Molecular physiology of water balance. N Engl J Med. 2015;372(14):1349-58. Meijer E et al. Copeptin, a surrogate marker of vasopressin, is associated with disease severity in autosomal dominant polycystic kidney disease. Clin J Am Soc Nephrol. 2011;6(2):361-8. Meijer E et al. Potential deleterious effects of vasopressin in chronic kidney disease and particularly autosomal dominant polycystic kidney disease. Kidney Blood Press Res. 2011;34(4):235-44. Boertien WE et al. Copeptin, a surrogate marker for vasopressin, is associated with kidney function decline in subjects with autosomal dominant polycystic kidney disease. Nephrol Dial Transplant. 2012;27(11):4131-7. Zittema D et al. Vasopressin, copeptin, and renal concentrating capacity in patients with autosomal dominant polycystic kidney disease without renal impairment. Clin J Am Soc Nephrol. 2012;7(6):906-13. Gattone VH 2nd et al. Developmental expression of urine concentrationassociated genes and their altered expression in murine infantile-type polycystic kidney disease. Dev Genet. 1999;24(3-4):309-18. Gattone VH 2nd et al. Inhibition of renal cystic disease development and progression by a vasopressin V2 receptor antagonist. Nat Med. 2003;9(10):1323-6. Torres VE et al. Effective treatment of an orthologous model of autosomal dominant polycystic kidney disease. Nat Med. 2004;10(4):363-4. Yamamura Y et al. OPC-41061, a highly potent human vasopressin V2-receptor antagonist: pharmacological profile and aquaretic effect by single and multiple oral dosing in rats. J Pharmacol Exp Ther. 1998;287(3):860-7. Reif GA et al. Tolvaptan inhibits ERKdependent cell proliferation, Cl secretion, and in vitro cyst growth of human ADPKD cells stimulated by vasopressin. Am J Physiol Renal Physiol. 2011;301(5):F1005-13. Torres VE et al.; TEMPO 3:4 Trial Investigators: Tolvaptan in patients with autosomal dominant polycystic kidney disease. N Engl J Med. 2012;367(25):2407-18. Gansevoort RT et al.; TEMPO 3:4 Investigators. Albuminuria and tolvaptan in autosomal-dominant polycystic kidney disease: results of the TEMPO 3:4 Trial. Nephrol Dial Transplant. 2016;31(11):1887-94. Gansevoort RT et al. Recommendations for the use of tolvaptan in autosomal dominant polycystic kidney disease: a position statement on behalf of the ERA-EDTA Working Groups on Inherited Kidney Disorders and European Renal Best Practice. Nephrol Dial Transplant. 2016;31(3):337-48. Torres VE et al. Effect of Tolvaptan in Autosomal Dominant Polycystic Kidney Disease by CKD Stage: Results from the TEMPO 3:4 Trial. Clin J Am Soc Nephrol. 2016;11(5):803-11. European Medicines Agency. Tolvaptan (Jinarc): Summary of product characteristics. 2015. Available at: http://www.ema.europa.eu/docs/ en_GB/document_library/EPAR_-_ Product_Information/human/002788/ WC500187921.pdf. Last accessed: 28 Feb 2017. Food and Drug Administration; Center for Drug Evaluation and Research. FDA Briefing Document: Cardiovascular and Renal Drug Advisory Committee Meeting. 2013. Available at: http://www.fda. gov/downloads/AdvisoryCommittees/CommitteesMeetingMaterials/Drugs/CardiovascularandRenalDrugsAdvisoryCommittee/UCM363343.pdf. Last accessed: 20 June 2017. Otsuka Pharmaceutical Development and Commercialization, Inc. Open-label Trial to Evaluate the Long Term Safety of Titrated Immediate-release Tolvaptan in Subjects With Autosomal Dominant Polycystic Kidney Disease. NCT02251275. http://www.clinicaltrials.gov/ct2/show/ NCT02251275. Otsuka Pharmaceutical Development and Commercialization, Inc. Efficacy and Safety of Tolvaptan in Subjects With Chronic Kidney Disease Between Late Stage 2 to Early Stage 4 Due to Autosomal Dominant Polycystic Kidney Disease. NCT02160145. https://www. clinicaltrials.gov/ct2/show/NCT02160145. Rai U et al. Therapeutic uses of somatostatin and its analogues: Current view and potential applications. Pharmacol Ther. 2015;152:98-110. Friedlander G, Amiel C. Somatostatin and alpha 2-adrenergic agonists selectively inhibit vasopressin-induced cyclic AMP accumulation in MDCK cells. FEBS Lett. 1986;198(1):38-42. Winkler SN et al. Effect of somatostatin on vasopressin-induced antidiuresis and renal cyclic AMP of rats. Miner Electrolyte Metab. 1982;7(1):8-14. Tan CK et al. Human cholangiocarcinomas express somatostatin receptors and respond to somatostatin with growth inhibition. Gastroenterology. 1995;108(6):1908-16. Irazabal MV, Torres VE Experimental therapies and ongoing clinical trials to slow down progression of ADPKD. Curr Hypertens Rev. 2013;9(1):44-59. Ruggenenti P et al. Safety and efficacy of long-acting somatostatin treatment in autosomal-dominant polycystic kidney disease. Kidney Int. 2005;68(1):206-16. van Keimpema L et al. Lanreotide reduces the volume of polycystic liver: a randomized, double-blind, placebocontrolled trial. Gastroenterology. 2009; 137(5):1661-8.e1-2. Chrispijn M et al. The long-term outcome of patients with polycystic liver disease treated with lanreotide. Aliment Pharmacol Ther. 2012;35(2):266-74. Hogan MC et al. Randomized clinical trial of long-acting somatostatin for autosomal dominant polycystic kidney and liver disease. J Am Soc Nephrol. 2010; 21(6):1052-61. Caroli A et al. Effect of long acting somatostatin analogue on kidney and cyst growth in autosomal dominant polycystic kidney disease (ALADIN): a randomised, placebo-controlled, multicentre trial. Lancet (London, England) 2013;382(9903):1485-95. Gevers TJ et al. Effect of lanreotide on polycystic liver and kidneys in autosomal dominant polycystic kidney disease: an observational trial. Liver Int. 2015;35(5):1607-14. Myint TM et al. Treatments to slow progression of autosomal dominant polycystic kidney disease: systematic review and meta-analysis of randomized trials. Nephrology (Carlton). 2014;19(4):217-26. Lantinga MA et al.; DIPAK Consortium. Hepatic Cyst Infection During Use of the Somatostatin Analog Lanreotide in Autosomal Dominant Polycystic Kidney Disease: An Interim Analysis of the Randomized Open-Label Multicenter DIPAK-1 Study. Drug Saf. 2017;40(2):153-67. University Medical Center Groningen. Study of Lanreotide to Treat Polycystic Kidney Disease (DIPAK1). NCT01616927. https://clinicaltrials.gov/ct2/show/NCT01 616927?term=NCT01616927%3A+Study+ of+Lanreotide+to+Treat+Polycystic+Kidn ey+Disease&rank=1. Mario Negri Institute for Pharmacological Research. Somatostatin in Patients with Autosomal Dominant Polycystic Kidney Disease and Moderate to Severe Renal Insufficiency (ALADIN 2). NCT01377246. https://clinicaltrials. gov/ct2/results?term=NCT01377246& Search=Search. Assistance Publique - Hôpitaux de Paris. Lanreotide in Polycystic Kidney Disease Study (LIPS). NCT02127437. https://clinicaltrials.gov/ct2/show/NCT02 127437?term=NCT02127437&rank=1. Mayo Clinic. Pasireotide LAR in Severe Polycystic Liver Disease (SOM230)]. NCT01670110. https://clinicaltrials.gov/ ct2/show/NCT01670110?term=NCT01670 110&rank=1. Ibraghimov-Beskrovnaya O, Natoli TA. mTOR signaling in polycystic kidney disease. Trends Mol Med. 2011;17(11):625-33. Brook-Carter PT et al. Deletion of the TSC2 and PKD1 genes associated with severe infantile polycystic kidney disease--a contiguous gene syndrome. Nat Genet. 1994;8(4):328-32. Dibble CC et al. TBC1D7 is a third subunit of the TSC1-TSC2 complex upstream of mTORC1. Mol Cell. 2012;47(4):535-46. Shillingford JM et al. The mTOR pathway is regulated by polycystin-1, and its inhibition reverses renal cystogenesis in polycystic kidney disease. Proc Natl Acad Sci U S A. 2006;103(14):5466-71. Shillingford JM et al. Rapamycin ameliorates PKD resulting from conditional inactivation of Pkd1. J Am Soc Nephrol. 2010;21(3):489-97. Wahl PR et al. Inhibition of mTOR with sirolimus slows disease progression in Han:SPRD rats with autosomal dominant polycystic kidney disease (ADPKD). Nephrol Dial Transplant. 2006;21(3):598-604. Wu M et al. Everolimus retards cyst growth and preserves kidney function in a rodent model for polycystic kidney disease. Kidney Blood Press Res. 2007;30(4):253-9. Baba M et al. Kidney-targeted BirtHogg-Dube gene inactivation in a mouse model: Erk1/2 and Akt-mTOR activation, cell hyperproliferation, and polycystic kidneys. J Natl Cancer Inst. 2008;100(2):140-54. Walz G et al. Everolimus in patients with autosomal dominant polycystic kidney disease. N Engl J Med. 2010;363(9):830-40. Serra AL et al. Sirolimus and kidney growth in autosomal dominant polycystic kidney disease. N Engl J Med. 2010;363(9):820-9. Perico N et al. Sirolimus therapy to halt the progression of ADPKD. J Am Soc Nephrol. 2010;21(6):1031-40. Stallone G et al. Rapamycin for treatment of type I autosomal dominant polycystic kidney disease (RAPYD-study): a randomized, controlled study. Nephrol Dial Transplant. 2012;27(9):3560-7. Braun WE et al. Low-dose rapamycin (sirolimus) effects in autosomal dominant polycystic kidney disease: an open-label randomized controlled pilot study. Clin J Am Soc Nephrol. 2014;9(5):881-8. Medical University of Vienna. Pulsed Oral Sirolimus in Autosomal Dominant Polycystic Kidney Disease (RAP). NCT02055079. https://clinicaltrials.gov/ ct2/show/NCT02055079. Assistance Publique - Hôpitaux de Paris. The Efficacy of Everolimus in Reducing Total Native Kidney Volume in Polycystic Kidney Disease Transplanted Recipients (EVERKYSTE). NCT02134899. https://clinicaltrials.gov/ct2/show/ NCT02134899. Salvadori M, Tsalouchos A. Novel therapeutic strategies targeting molecular pathways of cystogenesis in autosomal polycystic kidney disease. J Ren Hepat Disord. 2017;1(1):35-49. Sweeney WE Jr et al. Src inhibition ameliorates polycystic kidney disease. J Am Soc Nephrol. 2008;19(7):1331-41. Pfizer. Bosutinib For Autosomal Dominant Polycystic Kidney Disease. NCT01233869. https://clinicaltrials.gov/ ct2/show/NCT01233869. Kadmon Corporation, LLC. Long-Term Treatment and Follow up of Subjects Completing 24 Months of Treatment With Tesevatinib on Study KD019-101. NCT02616055. https://clinicaltrials.gov/ ct2/show/NCT02616055. Kadmon Corporation, LLC. A Safety, Pharmacokinetic and Dose-Escalation Study of KD019 in Subjects With Autosomal Dominant Polycystic Kidney Disease (ADPKD). NCT01559363. https:// clinicaltrials.gov/ct2/show/NCT01559363. Leuenroth SJ et al. Triptolide reduces cyst formation in a neonatal to adult transition Pkd1 model of ADPKD. Nephrol Dial Transplant.2010;25(7):2187-94. Zhi-Hong Liu. Randomized Clinical Trial of Triptolide Woldifii for Autosomal Dominant Polycystic Kidney Disease. NCT00801268. https://clinicaltrials.gov/ ct2/show/NCT00801268. Shanghai Changzheng Hospital. Triptolide-Containing Formulation as Treatment for Autosomal Dominant Polycystic Kidney Disease (ADPKD). NCT02115659. https://clinicaltrials.gov/ ct2/show/NCT02115659. Yamaguchi T et al. Sorafenib inhibits cAMP-dependent ERK activation, cell proliferation, and in vitro cyst growth of human ADPKD cyst epithelial cells. Am J Physiol Renal Physiol. 2010;299(5):F944-51. Buchholz B et al. The Raf kinase inhibitor PLX5568 slows cyst proliferation in rat polycystic kidney disease but promotes renal and hepaticfibrosis. Nephrol Dial Transplant. 2011;26(11):3458-65. Bukanov NO et al. Long-lasting arrest of murine polycystic kidney disease with CDK inhibitor roscovitine. Nature. 2006;444(7121):949-52. Bukanov NO et al. CDK inhibitors R-roscovitine and S-CR8effectively block renal and hepatic cystogenesis in an orthologous model of ADPKD. Cell Cycle. 2012;11(21):4040-6. Chen HP. Histone deacetylases and mechanisms of regulation of gene expression. Crit Rev Oncog. 2015;20(1–2):35-47. Cao Y et al. Chemical modifier screen identifies HDAC inhibitors as suppressors of PKD models. Proc Natl Acad Sci USA. 2009;106(51):21819-24. Alan Yu. Uncontrolled, Open Label, Pilot and Feasibility Study of Niacinamide in Polycystic Kidney Disease (NIACPKD1). NCT02140814. https://clinicaltrials. gov/ct2/show/NCT02140814. University of Kansas Medical Center. Pilot Study of Niacinamide in Polycystic Kidney Disease (NIAC-PKD2). NCT02558595. https://clinicaltrials.gov/ ct2/show/NCT02558595. Yang B et al. Small-molecule CFTR inhibitors slow cyst growth in polycystic kidney disease. J Am Soc Nephrol. 2008;19:1300-10. Gwinn DM et al. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol Cell. 2008;30(2):214-26. Takiar V et al. Activating AMPactivated protein kinase (AMPK) slows renal cystogenesis. Proc Natl Acad Sci USA. 2011;108(6):2462-7. University of Pittsburgh. Metformin as a Novel Therapy for Autosomal Dominant Polycystic Kidney Disease (TAME). NCT02656017. https://clinicaltrials.gov/ ct2/show/NCT02656017. Blazer-Yost BL et al. Pioglitazone attenuates cystic burden in the PCK rodent model of polycystic kidney disease. PPAR Res. 2010;2010:274376. Yoshihara D et al. PPAR-gamma agonist ameliorates kidney and liver disease in an orthologous rat model of human autosomal recessive polycystic kidney disease. Am J Physiol Renal Physiol. 2011;300(2):F465-74. Dai B et al. Rosiglitazone attenuates development of polycystic kidney disease and prolongs survival in Han:SPRD rats. Clin Sci (Lond). 2010;119(8):323-33. Indiana University. Use of Low Dose Pioglitazone to Treat Autosomal Dominant Polycystic Kidney Disease (PIOPKD). NCT02697617. https://clinicaltrials.gov/ ct2/show/NCT02697617. Rysz J et al. Combination drug versus monotherapy for the treatment of autosomal dominant polycystic kidney disease. Expert Opin Pharmacother. 2016;17(15)2049-56.

Please rate the quality of this content

As you found this content interesting...

Follow us on social media!

We are sorry that this content was not interesting for you!

Let us improve this content!

Tell us how we can improve this content?

Keep your finger on the pulse

Join Now

Elevating the Quality of Healthcare Globally

>