Skip to main content

Pharmacokinetic and pharmacodynamic analyses of terlipressin in patients with hepatorenal syndrome

Abstract

The objective of this population pharmacokinetics (PK) analysis was to characterize the PK of terlipressin and its active metabolite, lysine-vasopressin (L-VP), in patients with hepatorenal syndrome (HRS), following intravenous administration of terlipressin 1 mg to 2 mg every 6 h. Sparse PK samples from 69 patients with HRS who participated in terlipressin phase 3 clinical studies were used for model development. In addition, mean arterial pressure (MAP) and heart rate (HR) from 40 patients with HRS were available to explore the relationship between terlipressin and L-VP plasma concentrations and pharmacodynamic (PD) response. A two-compartment model with first-order elimination adequately described the PK of terlipressin. L-VP was well characterized as the active metabolite of terlipressin by a one-compartment model with first-order elimination. The population PK modeling results showed that the estimated clearances for terlipressin and L-VP are 27.4 L/h and 318 L/h, respectively, for a typical patient with a body weight of 86 kg. Body weight was identified as the only covariate for the clearance of terlipressin. However, simulation suggested that body weight had no clinically meaningful effects on the exposure of L-VP through terlipressin. Therefore, no weight-based dose is needed for terlipressin to treat HRS patients. PD response, change in MAP, and HR were well correlated to L-VP concentrations; compared with baseline values, the estimated maximum decrease in HR would be 10.6 bpm and the estimated maximum increase in MAP would be 16.2 mm Hg.

Introduction

Hepatorenal syndrome (HRS), a potentially reversible renal failure, is a serious, rapidly progressing disease complicating decompensated chronic liver disease associated with cirrhosis (Arroyo et al. 2006; Salerno et al. 2007; Angeli et al. 2015). If left untreated, patients with HRS have a poor prognosis (Gines et al. 1993; Alessandria et al. 2005). At present, there are no approved therapies available in the United States or Canada for the treatment of HRS. An increasing body of knowledge of the pathophysiology of HRS has demonstrated that vasoconstrictive drug therapy may improve renal function in patients with HRS (Salerno et al. 2007; European Association for the Study of the Liver 2018).

Terlipressin, a synthetic vasopressin analog, is a 12-amino-acid peptide with the chemical name N-[N-(N-glycylglycyl)glycyl]-8-l-lysinevasopressin (Jamil et al. 2018; Forsling et al. 1980) with an average molecular mass of 1227.4 Da (as a free base). It differs from endogenous human vasopressin by the substitution of lysine for arginine at the eighth position of the endogenous molecule (lys8) and the addition of 3 glycyl residues at the amino terminus (Jamil et al. 2018). Terlipressin acts as a systemic vasoconstrictor via the vascular vasopressin V1 receptors, primarily due to its metabolite lysine-vasopressin (L-VP), albeit of its own low potency. Results from the binding of terlipressin and L-VP investigation suggest that the binding affinity of L-VP to the V1 or V2 receptor is 600–700-fold greater than that of terlipressin. Cellular activity indicates that terlipressin is a partial agonist at V1a, and a full agonist at V2; while L-VP is a full agonist at both V1 and V2 receptors. It is believed that the in vivo response to terlipressin administration is likely due to partial V1a agonist activity of terlipressin and full V1 agonist activity of L-VP (Jamil et al. 2018; Forsling et al. 1980; Pliska et al. 1976; Nilsson et al. 1990; Wisniewski et al. 2006; Colson et al. 2016; Bernadich et al. 1998). In patients with HRS, the potent agonist activity of terlipressin and L-VP at V1 receptors leads to splanchnic vasoconstriction that results in increased mean arterial pressure (MAP) and effective intravascular volume, as well as decreased heart rate (HR), which improves renal function (European Association for the Study of the Liver 2018; Jamil et al. 2018; Serpa Neto et al. 2012; Narahara et al. 2009; Boyer et al. 2016; Sanyal et al. 2008).

After intravenous (IV) administration of terlipressin, the glycyl residues of terlipressin are cleaved by endogenous tissue proteases resulting in a rapid decrease of terlipressin levels in blood circulation. Terlipressin is almost completely metabolized in the tissues by ubiquitous peptidases and L-VP is released gradually from tissues into the circulation, which likely contributes to the longer duration of action of terlipressin than that of vasopressin. The prolonged pharmacologic effect following injection of terlipressin at 7.5 μg/kg was from the conversion of terlipressin to L-VP in humans. Further, the clinical use of terlipressin lies in its ability to generate L-VP over a prolonged period. Adding amino acid residues to the amino terminus can prolong the duration of biological activity of vasopressin (Forsling et al. 1980; Cort et al. 1975; Kyncl and Rudinger 1970). The pharmacokinetics (PK) of terlipressin and L-VP has been characterized in healthy subjects (Forsling et al. 1980; Nilsson et al. 1990). Plasma concentrations of L-VP peak at 60 to 120 min after IV administration of terlipressin (Forsling et al. 1980). There is a biphasic decline in plasma terlipressin concentration with a rapid initial distribution phase (half-life of 8 min) followed by a slower second phase with an elimination half-life of 50 min (Nilsson et al. 1990). The volume of distribution is 0.7 L/kg and the plasma clearance is 9 mL/kg/min (Nilsson et al. 1990). Less than 1% of administered terlipressin is excreted in urine, and less than 0.1% is excreted as L-VP in urine (Forsling et al. 1980).

Based on the elimination half-life of 50 min (Nilsson et al. 1990) and its prolonged pharmacological effect due to the conversion to L-VP (Forsling et al. 1980; Cort et al. 1975), terlipressin is given as a bolus injection every 4 to 6 h (2 to 6 mg/day, up to 8 to 12 mg/day). More recently, a continuous intravenous infusion (CIV) of terlipressin has been used in HRS in bovine sepsis models, as well as in several small clinical studies. A recent Australian study in five patients with refractory ascites, using 3.4 mg/24 h CIV therapy of terlipressin for 4 weeks, resulted in promising clinical response and safety profiles (Gow et al. 2016).

Owing to the restriction of blood sampling in target patients, the PK data of terlipressin and L-VP are very limited. Understanding the PK and pharmacodynamic (PD) characteristics of terlipressin and L-VP in patients with HRS is necessary to guide the usage of terlipressin, such as dosing regimen and dose adjustment, in a clinical setting. Using the PK and PD data from two clinical studies in patients with HRS, the objectives of this analysis were to characterize the PK and PD, as measured by MAP and HR, of terlipressin and L-VP.

Materials and methods

Data

The data used for population PK/PD model development are from two phase 3 randomized, multicenter, double-blind, placebo-controlled studies of IV bolus administration of terlipressin in patients with HRS (study 1: OT-0401, NCT00089570; study 2: REVERSE, NCT01143246) (Boyer et al. 2016; Sanyal et al. 2008). Each study protocol was reviewed and approved by an institutional review board at each study site and written informed consent was obtained from each patient before enrollment in the studies. A summary of the study designs is provided in Supplementary Appendix 1.

In study 1, patients received up to 14 days of study treatment with terlipressin or matched placebo. Terlipressin was administered as slow IV bolus doses of 1 mg every 6 h (4 mg/day), with a dose increase to 2 mg every 6 h if serum creatinine (SCr) decreased by less than 30% after a minimum of 10 doses were administered. One baseline sample and additional 2 samples per day were collected on days 3, 6, 9, and day 14 (or last treatment day, whichever came first). The sampling time points for each day included one time point from TIMEPOINT 1 (0.083, 0.5, 1.0, and 2.0 h) and one time point from TIMEPOINT 2 (3.0, 4.0, 5.0, and 6.0 h) (Table 1).

Table 1 Demographic and selected characteristics of patients with HRS in the population PK analysis

In study 2, patients received up to 14 days of study treatment with terlipressin or matched placebo, IV, every 6 h (maximum of 15 or 16 days if HRS reversal was first achieved on days 13 or 14, respectively). Blood samples were taken, one predose and three postdose, based on randomization assignments from one of each group during the first dosing interval on day 1. The first time point was randomly assigned from one of the six time points of TIMEPOINT A, (5, 6, 7, 8, 9, and 10 min [± 1 min]), the second time point from one of the four time points of TIMEPOINT B, (0.5, 1, 1.5, and 2 h [± 5 min]), and the third time point from one of the four time points of TIMEPOINT C, (3, 3.5, 4, and 5 h [± 10 min]) for each patient (Table 1).

Bioanalytical methods

Blood samples (7 mL each) were processed within 30 min after they were obtained. Preliminary stability data indicated that terlipressin and L-VP were stable in plasma at −20 °C and plasma samples were to be stored under controlled conditions at −20 °C; however, it was later determined that terlipressin and L-VP in human plasma were stable at −20 °C only up to 148 days. Plasma samples that were stored at −20 °C and assayed within 148 days after collection were considered to be within the stability period, and the data are considered valid for PK use. Plasma samples that were stored at −20 °C and assayed after 148 days had elapsed were notated as being outside the stability period. For study 1, the lower limit of quantitation (LLOQ) was 0.5 ng/mL for terlipressin and 0.1 ng/mL for L-VP. For study 2, the LLOQ was 0.25 ng/mL for terlipressin and 0.05 ng/mL for L-VP.

Population PK and PK/PD analysis methods

Population PK and PK/PD analyses were performed using nonlinear mixed-effect modeling with NONMEM (version 7.3.0, ICON Development Solutions, Ellicott City, MD) in accordance with the Guidance for Industry: Population Pharmacokinetics of the US Food and Drug Administration. The first-order conditional estimation method, as well as the Laplace (INTER) with Method 3 (M3) to handle below LLOQ (below quantitation limit [BQL]) data, was used for the analysis (Ahn et al. 2008; Bergstrand and Karlsson 2009). The plasma concentration data, dosing history, sampling date/time, demographic data, baseline laboratory test results, and concomitant medications were merged using SAS software version 9.3 or higher (SAS Institute, Cary, NC). Data summary, plotting, and model diagnostics were completed with R, version 3.5.2.

Population PK model development

Per PK information from healthy subjects, a linear two-compartment model was adopted to describe the PK data for terlipressin. A third compartment was added to describe the PK of L-VP, as depicted in Fig. 1. Model parameters are presented in the footnote of Fig. 1; these included total body clearance of terlipressin (CLA), central compartment volume (V1A), intercompartment flow rate (QA), and peripheral compartment volume (V2A) for terlipressin; and the apparent clearance (CLM/Fm) and apparent central compartment volume (VM/Fm) for L-VP. The fraction of terlipressin dose that is metabolized to L-VP, Fm, is unknown and cannot be estimated. Thus, CLM/Fm and VM/Fm were estimated from the population PK analysis. Log normal distribution was assumed for between-subject variability (BSV) for PK parameters. An additive error model was used for log-transformed plasma concentrations. Different residual errors were assumed for terlipressin and L-VP.

Fig. 1
figure 1

Structural pharmacokinetic model for terlipressin and its active metabolite lysine-vasopressin. CLA, total body clearance for terlipressin; CLM/Fm and VM/Fm, apparent volume of central compartment and clearance for lysine-vasopressin; IV, intravenous; QA, intercompartmental flow rate; V1A, central volume of distribution for terlipressin; V2A, peripheral volume of distribution for terlipressin

Continuous covariates included in the model were body weight, age, renal function (described using creatinine clearance [CRCL]), hepatic function (aspartate aminotransferase [AST], alanine aminotransferase [ALT], alkaline phosphatase [ALKP], bilirubin). Discrete covariates included age group (< 65 years versus ≥ 65 years of age), sex, race, Child-Pugh score for hepatic function, and concomitant medications. Graphical exploratory evaluation of covariate-parameter relationships was performed, followed by the stepwise method. Covariate analysis was also assisted by clinical relevance, reduction in BSV, and the improvement of the Loess fitting.

Plasma concentration values BQL were treated as the lower limit of quantitation divided by 2 (LLOQ/2). In addition, M3 was used for parameter estimation (Ahn et al. 2008; Bergstrand and Karlsson 2009). Comparison between the parameters from two methods was conducted.

Population PK/PD models

Sequential PK/PD analysis was performed to develop the population PK/PD model. Predicted individual terlipressin and L-VP concentrations at the same time points as the PD measurements were used in the PD model development. Both linear and nonlinear PK/PD models for MAP or HR were tested. A log-normal distribution for BSV of the baseline MAP or HR was assumed. Normal distribution was assumed to describe BSV in the slope in the linear model or maximum change in HR (Imax) in the Imax model. An additive error model was assumed.

Model evaluation

Model evaluation was conducted using standard diagnostic tools such as goodness-of-fit criteria, diagnostic plots, and a prediction-corrected visual predictive check (VPC) (Holford 2005). Model stability and performance were also evaluated using the bootstrap method.

Simulation to evaluate dosing regimen

Model-based simulations were conducted to evaluate the impact of identified covariates on the exposure of terlipressin and L-VP, and their potential impact on the dosing regimen such as weight-based or flat dose of terlipressin.

Results

The population PK analysis included 69 adult patients with HRS who had 227 measurable terlipressin and 246 measurable L-VP plasma samples collected from study 1 and study 2 altogether. In studies 1 and 2, 16.5% of total terlipressin PK samples were BQL and 10.2% of total L-VP PK samples were BQL.

Demographics and selected characteristics at baseline for the study population are shown in Table 1. Two patients (2.9%) had missing CRCL, weight, and Child-Pugh score data, and one had missing race information. The majority of the patient population was White (84%), male (58%), and younger than 65 years of age (87%). The median (minimum, maximum) body weight was 86 (43, 170) kg. The missing CRCL, weight, and Child-Pugh score values were imputed as the median of the study population, and the patient with missing race data was classified as White. As per the nature of HRS, all patients from studies 1 and 2 exhibited renal impairment (CRCL < 60 mL/min), and a majority (77%) had severe hepatic impairment, with a Child-Pugh score ≥ 10.

Because the BQL data points for both terlipressin and L-VP were > 10%, the M3 method to handle BQL data was also applied to the parameter estimation in addition to handling BQL as LLOQ/2. Results of the two methods of handling BQL were in good agreement. For example, the difference in parameter estimates including error models of the two methods ranged from −0.4 to 7.6%. The difference in estimated BSV from the two methods was less than 5.5% except for the central compartment volume of terlipressin, which was 35%. Optimization using M3 took longer computational time. Therefore, the method of setting BQL as LLOQ/2 was used in subsequent analyses. Diagnostic plots of the PK base model derived by setting BQL=LLOQ/2 are shown in Supplementary Fig. S1.

Covariate analysis

Graphical exploratory evaluation of covariate-parameter relationships from the base population PK model is presented in Supplementary Fig. S2. It appeared that only body weight and CRCL had a consistent trend with the PK parameters of terlipressin and L-VP.

During the stepwise forward selection step, body weight (WT), sex, and renal function were added to the clearance of terlipressin (CLA) and L-VP; sex and Child-Pugh score were added to the central compartment of terlipressin and L-VP. Stepwise backward elimination revealed that body weight on CLA was the only significant covariate (Table 2). The body weight effect on CLA can be described as below:

$$\mathrm{CLA}=27.4\times {\left(\mathrm{WT}/86\right)}^{0.549}\ \mathrm{L}/\mathrm{h}$$
Table 2 Population pharmacokinetic parameters of terlipressin and L-VP in patients with HRS (final PK model)

For patients with a body weight of 86 kg, the median weight in our population, the typical value of CLA was 27.4 L/h. Given the body weight range of the two studies, from 43 kg to 170 kg, CLA varied from 18.7 L/h to 39.8 L/h, respectively. The typical value of the central compartment volume for terlipressin was 6.31 L. The mean CLM/Fm and VM/Fm for L-VP were 318 L/h and 1370 L, respectively. A large degree of BSV was derived from the analysis, with the BSV in V1A, CLM/Fm, and VM/Fm above 60%.

Evaluation of the population PK model

Diagnostic plots for the final population PK model shown in Fig. 2 suggested that the final population PK model was adequate in describing plasma concentration-time profiles in patients with HRS.

Fig. 2
figure 2

Goodness-of-fit diagnostic plots of the final model of terlipressin. CWRES, conditional weighted residual value; DV, observed concentration; IPRE, individual predicted value; L-VP, lysine-vasopressin; PRED, typical value. Log transformed observed data used for model development

Based on the bootstrap results shown in Supplementary Table S1, the parameter estimates of the final population PK model were comparable to the median values of 1110 resampled bootstrapping runs. All of the parameter estimates for the final population PK model fell within the 95% confidence intervals (CIs) obtained from the bootstrap, the results of the bootstrap runs demonstrated sufficient stability of the final PK model.

Figure 3 presents prediction-corrected VPC plots of terlipressin and L-VP for the final population PK model. In general, the dashed and solid red lines (observations) for both terlipressin and L-VP run within their designated blocks (model-predicted 90% CI), indicating that the final population PK model had good predictability.

Fig. 3
figure 3

Visual predictive check: model-predicted and observed plasma concentrations of terlipressin and L-VP in patients with HRS. Blue circle markers represent observed PK data; solid red line represents the median observed values; solid red areas indicate the model-predicted 90% confidence interval for the median. The 2.5th and 97.5th percentiles of observed data are presented with dashed red lines, and the blue solid areas represent the model-predicted 90% confidence intervals for the corresponding percentiles. HRS, hepatorenal syndrome; L-VP, lysine-vasopressin; PK, pharmacokinetic

PK/PD model

HR and MAP measurements were available from 40 patients in study 2. Increasing MAP and decreasing HR were associated with the increase of L-VP concentration (data not shown). The Imax model was selected as the final PK/PD model for MAP based on its lower Akaike Information Criterion (AIC) value compared with the linear model. A summary of PK/PD parameter estimates of the MAP Imax model is presented in Table 3. The estimated L-VP potency (concentration of L-VP that produces 50% of the maximal response [IC50]) was 0.26 ng/mL (95% CI = 0.008 to 1.28). The maximum increase in MAP Imax due to L-VP exposure was 16.2 mm Hg (95% CI = 9.0 to 37.8).

Table 3 Parameter estimates of the MAP Imax model

Similarly, the Imax model was selected as the final PK/PD model for HR based on its lower AIC value. Model parameters are presented in Table 4. The estimated potency (IC50) was 0.25 ng/mL (95% CI = 0.056 to 1.06). The Imax due to L-VP exposure was −10.6 bpm (95% CI = −24.0 to −5.8).

Table 4 Parameter estimates of the HR Imax model

The model-predicted relationship between L-VP plasma concentrations and MAP and HR measurements is illustrated in Supplementary Fig. S3.

Simulation to evaluate dosing algorithm

Although body weight has no direct impact on the exposure of L-VP, the impact of body weight on terlipressin might alter the exposure of L-VP. To assess this hypothesis, model-based simulations were conducted. Figure 4 presents the relationship between body weight and the exposure (average concentration within one dosing interval) of terlipressin and L-VP. Results presented in Fig. 4 demonstrated that the changes in terlipressin exposure with body weight had no noticeable effect on L-VP exposure. Because L-VP is the primary pharmacologically active moiety, no body weight-based dosing regimen of terlipressin would be necessary in patients with HRS.

Fig. 4
figure 4

Simulated terlipressin and L-VP plasma concentrations at steady state following terlipressin 1 mg administered every 6 h in patients with HRS with different body weights. The solid blue line indicates the Loess fitting and the shaded area indicates the 90% confidence interval. Conc, concentration; HRS, hepatorenal syndrome; L-VP, lysine-vasopressin; Q6h, every 6 h; SS, steady-state.

Discussion

Patients with HRS are expected to have advanced, progressive liver and renal disease. Therefore, to identify the effect of severe hepatic and renal dysfunction on terlipressin PK is important. Covariate analysis in the population PK modeling suggested that hepatic dysfunction measured by baseline values for ALT, AST, total bilirubin, ALKP, and Child-Pugh scores did not appear to have any significant effect on the PK of terlipressin and L-VP. This is likely due to the fact that the liver is not involved in the conversion of terlipressin to L-VP. That renal dysfunction had no statistically significant impact is in good agreement with reported clinical results that less than 1% of administered terlipressin is excreted in urine, and less than 0.1% is excreted as L-VP in urine (Forsling et al. 1980). As such, no dose adjustment of terlipressin was recommended for patients with HRS and varying degrees of hepatic or renal dysfunction.

Age effect on the PK of terlipressin and L-VP was tested either as a continuous covariate or a categorical covariate. Neither approach showed that age or race were statistically significant covariates. Of the 69 patients evaluated, 9 (13%) were elderly (aged ≥ 65 years) and most patients were White (83.8%). Therefore, the conclusion derived from the covariate analysis that neither age nor race had any significant effect on the PK of terlipressin had its limitations.

Two different methods to handle BQL were implemented. Parameter estimates using the two methods are in good agreement. Keizer and colleagues (Keizer et al. 2015) have reported similar outcomes between different methods in handling BQL. In particular, for a population PK model with IV administration, when the percentage data below LLOQ is less than 20%, setting BQL = LLOQ/2 has a similar performance to that of the M3 method.

PK/PD modeling

Increased MAP after administration of terlipressin in patients with HRS has been reported in the literature (Boyer et al. 2016; Sanyal et al. 2008), and this effect results in improved renal function and has provided the therapeutic rationale for terlipressin treatment in this setting. Our studies confirmed the reported effect of MAP increasing with terlipressin treatment.

Both linear and nonlinear models were explored to characterize the correlations between terlipressin and L-VP concentrations and MAP as well as HR. These exploratory investigations suggested that a mixed model with linear correlation for MAP increase attributed to terlipressin and an Imax model attributed to L-VP are appropriate. However, because of the fast decline in terlipressin concentration (half-life is 0.9 h), the early time concentration and MAP data points were not sufficient to differentiate the response contributed by terlipressin from L-VP. A simplified model assuming that the increase in MAP primarily related to L-VP described the observed MAP data equally well following the administration of terlipressin.

Conclusion

Plasma terlipressin concentrations in patients with HRS were well characterized by a two-compartment PK model while the L-VP plasma concentrations were well characterized by a one-compartment PK model. None of the covariates examined were found to have a significant effect on the PK of terlipressin and L-VP, except for body weight on clearance of terlipressin. Results from the model-based simulations suggested that although terlipressin exposure decreases with increasing body weight, the change in L-VP levels with body weight is negligible. Because L-VP is the primary pharmacologically active moiety, no body weight-based dosage adjustment of terlipressin would be necessary when treating patients with HRS. The PK/PD relationship of L-VP to MAP was well characterized by an Imax model, with a maximum estimated increase in MAP of 16.2 mm Hg and a decrease in HR of 10.6 bpm.

Availability of data and materials

Discussion of statistical endpoints and analysis are included in the manuscript. Summary aggregate (basic) results (including information on adverse events) and the study protocol will be available on clinicaltrials.gov (OT-0401, NCT00089570; REVERSE, NCT01143246) when required by regulation. Requests for additional information should be directed to the trial sponsor company at medinfo@mnk.com.

Abbreviations

AIC:

Akaike Information Criterion

ALKP:

Alkaline phosphatase

ALT:

Alanine aminotransferase

AST:

Aspartate aminotransferase

bpm:

Beats per minute

BQL:

Below quantitation limit

BSV:

Between-subject variability

CI:

Confidence interval

CIV:

Continuous intravenous infusion

CLA:

Total body clearance of terlipressin

CLM/Fm :

Apparent clearance for L-VP

Conc:

Concentration

CRCL:

Creatinine clearance

CV%:

Coefficient of variation

CWRES:

Conditional weighted residual value

DV:

Observed concentration

HR:

Heart rate

HRS:

Hepatorenal syndrome

IC50 :

50% of the maximal response

Imax:

Maximum change in HR

IPRE:

Individual predicted value

IV:

Intravenous

LLOQ:

Lower limit of quantitation

LLOQ/2:

Lower limit of quantitation divided by 2

L-VP:

Lysine-vasopressin

M3:

Method 3

MAP:

Mean arterial pressure

NA:

Not applicable

PD:

Pharmacodynamic

PK:

Pharmacokinetic

PRED:

Typical value

Q6h:

Every 6 h

QA:

Intercompartmental flow

RSE:

Relative standard error

RUE1:

Residual unexplained variability in study 1

RUE2:

Residual unexplained variability in study 2

SCr:

Serum creatinine

SS:

Steady-state

VM/Fm :

Apparent volume of L-VP

VPC:

Visual predictive check

WT:

Body weight

References

  • Ahn JE, Karlsson MO, Dunne A, Ludden TM (2008) Likelihood based approaches to handling data below the quantification limit using NONMEM VI. J Pharmacokinet Pharmacodyn 35:401–421

    Article  Google Scholar 

  • Alessandria C, Ozdogan O, Guevara M, Restuccia T, Jimenez W, Arroyo V et al (2005) MELD score and clinical type predict prognosis in hepatorenal syndrome: relevance to liver transplantation. Hepatology 41:1282–1289

    Article  Google Scholar 

  • Angeli P, Gines P, Wong F, Bernardi M, Boyer TD, Gerbes A et al (2015) Diagnosis and management of acute kidney injury in patients with cirrhosis: revised consensus recommendations of the International Club of Ascites. J Hepatol 62:968–974

    Article  Google Scholar 

  • Arroyo V, Terra C, Gines P (2006) New treatments of hepatorenal syndrome. Semin Liver Dis 26:254–264

    Article  Google Scholar 

  • Bergstrand M, Karlsson MO (2009) Handling data below the limit of quantification in mixed effect models. AAPS J 11:371–380

    Article  CAS  Google Scholar 

  • Bernadich C, Bandi JC, Melin P, Bosch J (1998) Effects of F-180, a new selective vasoconstrictor peptide, compared with terlipressin and vasopressin on systemic and splanchnic hemodynamics in a rat model of portal hypertension. Hepatology 27:351–356

    Article  CAS  Google Scholar 

  • Boyer TD, Sanyal AJ, Wong F, Frederick RT, Lake JR, O’Leary JG et al (2016) Terlipressin plus albumin is more effective than albumin alone in improving renal function in patients with cirrhosis and hepatorenal syndrome type 1. Gastroenterology 150:1579–1589

    Article  CAS  Google Scholar 

  • Colson P, Virsolvy A, Gaudard P, Charrabi A, Corbani M, Maniere M et al (2016) Terlipressin, a vasoactive prodrug recommended in hepatorenal syndrome, is an agonist of human V1, V2, and V1B receptors: implications for its safety profile. Pharmacol Res 113:257–264

    Article  CAS  Google Scholar 

  • Cort JH, Albrecht I, Novakova J, Mulder JL, Jost K (1975) Regional and systemic haemodynamic effects of some vasopressins: structural features of the hormone which prolong activity. Eur J Clin Invest 5:165–175

    Article  CAS  Google Scholar 

  • European Association for the Study of the Liver (2018) EASL Clinical Practice Guidelines for the management of patients with decompensated cirrhosis. J Hepatol 69:406–460

    Article  Google Scholar 

  • Forsling ML, Aziz LA, Miller M, Davies R, Donovan B (1980) Conversion of triglycylvasopressin to lysine-vasopressin in man. J Endocrinol 85:237–244

    Article  CAS  Google Scholar 

  • Gines A, Escorsell A, Gines P, Salo J, Jimenez W, Inglada L et al (1993) Incidence, predictive factors, and prognosis of the hepatorenal syndrome in cirrhosis with ascites. Gastroenterology 105:229–236

    Article  CAS  Google Scholar 

  • Gow PJ, Ardalan ZS, Vasudevan A, Testro AG, Ye B, Angus PW (2016) Outpatient terlipressin infusion for the treatment of refractory ascites. Am J Gastroenterol 111:1041–1042

    Article  CAS  Google Scholar 

  • Holford N. The visual predictive check superiority to standard diagnostic (Rorschach) plots [poster 738]. Presented at: Annual Meeting of the Population Approach Group in Europe; June 16-17, 2005; Pamplona, Spain.

  • Jamil K, Pappas SC, Devarakonda KR (2018) In vitro binding and receptor-mediated activity of terlipressin at vasopressin receptors V1 and V2. J Exp Pharmacol 10:1–7

    Article  CAS  Google Scholar 

  • Keizer RJ, Jansen RS, Rosing H, Thijssen B, Beijnen JH, Schellens JH et al (2015) Incorporation of concentration data below the limit of quantification in population pharmacokinetic analyses. Pharmacol Research Perspectives 3:e00131

    Article  Google Scholar 

  • Kyncl J, Rudinger J (1970) Excretion of antidiuretic activity in the urine of cats and rats after administration of the synthetic hormonogen, N alpha-glycyl-glycyl-glycyl-[8-lysine]-vasopressin (triglycylvasopressin). J Endocrinol 48:157–165

    Article  CAS  Google Scholar 

  • Narahara Y, Kanazawa H, Taki Y, Kimura Y, Atsukawa M, Katakura T et al (2009) Effects of terlipressin on systemic, hepatic and renal hemodynamics in patients with cirrhosis. J Gastroenterol Hepatol 24:1791–1797

    Article  CAS  Google Scholar 

  • Nilsson G, Lindblom P, Ohlin M, Berling R, Vernersson E (1990) Pharmacokinetics of terlipressin after single i.v. doses to healthy volunteers. Drugs Exp Clin Res 16:307–314

    CAS  PubMed  Google Scholar 

  • Pliska V, Chard T, Rudinger J, Forsling ML (1976) In vivo activation of synthetic hormonogens of lysine vasopressin: Na-glycyl-glycyl-glycyl-[8-lysine]vassopressin in the cat. Acta Endocrinol (Copenh) 81:474–481

    Article  CAS  Google Scholar 

  • Salerno F, Gerbes A, Gines P, Wong F, Arroyo V (2007) Diagnosis, prevention and treatment of hepatorenal syndrome in cirrhosis. Gut 56:1310–1318

    CAS  PubMed  PubMed Central  Google Scholar 

  • Sanyal AJ, Boyer T, Garcia-Tsao G, Regenstein F, Rossaro L, Appenrodt B et al (2008) A randomized, prospective, double-blind, placebo-controlled trial of terlipressin for type 1 hepatorenal syndrome. Gastroenterology 134:1360–1368

    Article  CAS  Google Scholar 

  • Serpa Neto A, Nassar AP, Cardoso SO, Manetta JA, Pereira VG, Espósito DC et al (2012) Vasopressin and terlipressin in adult vasodilatory shock: a systematic review and meta-analysis of nine randomized controlled trials. Crit Care 16:R154

    Article  Google Scholar 

  • Wisniewski K, Alagarsamy S, Taki H, Miampamba M, Laporte R, Galyean R, et al (2006) Synthesis and biological activity of terlipressin and its putative metabolites. In: Blondelle SE (ed) Understanding biology using peptides: Proceedings of the Nineteenth American Peptide Symposium: 489-490. Springer.

Download references

Acknowledgements

This study was funded by Mallinckrodt Pharmaceuticals. Editorial support, which was conducted in accordance with Good Publication Practice (GPP3) and the International Committee of Medical Journal Editors (ICMJE) guidelines, was provided by Michael D. Morren, RPh, MBA, of Peloton Advantage, LLC, an OPEN Health company, and Oxford PharmaGenesis Inc, Newtown, PA, funded by Mallinckrodt Pharmaceuticals. The authors would like to thank Loan Pham, PhD, former employee of Mallinckrodt Pharmaceuticals, for her input on this study and Greg Tardie, PhD, of Mallinckrodt Pharmaceuticals for his scientific and editorial contributions to the draft.

Author disclosures

No portion of this manuscript has been submitted for publication elsewhere, and it is not under consideration for publication in another journal, website, or textbook.

Funding

Editorial support for this report was funded by Mallinckrodt Pharmaceuticals. The authors were responsible for all content and editorial decisions, and have not received honoraria related to the development of this publication. Editorial support, which was conducted in accordance with Good Publication Practice (GPP3) and the International Committee of Medical Journal Editors (ICMJE) guidelines, was provided by Peloton Advantage, LLC, an OPEN Health company, and Oxford PharmaGenesis Inc, Newtown, PA.

Author information

Authors and Affiliations

Authors

Contributions

Khurram Jamil reviewed each draft and provided comments on the analyses. Xiaofeng Wang provided interpretation of data, reviewed each draft, provided input on the discussion, and contributed written content. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Khurram Jamil.

Ethics declarations

Competing interests

Xiaofeng Wang reports that she is a former employee of Mallinckrodt Pharmaceuticals and that she does not hold shares in Mallinckrodt Pharmaceuticals.

Khurram Jamil reports that he is an employee of Mallinckrodt Pharmaceuticals and that he does not hold shares in Mallinckrodt Pharmaceuticals.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Xiaofeng Wang was an employee of Mallinckrodt Pharmaceuticals at the time the research was conducted; currently an employee of Alexion Pharmaceuticals.

Additional file

Additional file 1.

(DOCX 370 kb)

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, X., Jamil, K. Pharmacokinetic and pharmacodynamic analyses of terlipressin in patients with hepatorenal syndrome. AAPS Open 8, 7 (2022). https://doi.org/10.1186/s41120-022-00052-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s41120-022-00052-7

Keywords