Next Article in Journal
Application of Metal–Organic Framework in Diagnosis and Treatment of Diabetes
Next Article in Special Issue
Sex Differences in the Neuropsychiatric Effects and Pharmacokinetics of Cannabidiol: A Scoping Review
Previous Article in Journal
Research Progress on Nanoparticles-Based CRISPR/Cas9 System for Targeted Therapy of Tumors
Previous Article in Special Issue
Design, Synthesis and Pharmacological Evaluation of Novel Conformationally Restricted N-arylpiperazine Derivatives Characterized as D2/D3 Receptor Ligands, Candidates for the Treatment of Neurodegenerative Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Peptidomimetics and Their Applications for Opioid Peptide Drug Discovery

Department of Pharmacology, College of Medicine, University of Arizona, Tucson, AZ 85724, USA
Current Address: Calviri Inc., Phoenix, AZ 85004, USA.
Biomolecules 2022, 12(9), 1241; https://doi.org/10.3390/biom12091241
Submission received: 17 August 2022 / Revised: 29 August 2022 / Accepted: 2 September 2022 / Published: 5 September 2022
(This article belongs to the Special Issue Recent Advances in Central Nervous System Drug Discovery)

Abstract

:
Despite various advantages, opioid peptides have been limited in their therapeutic uses due to the main drawbacks in metabolic stability, blood-brain barrier permeability, and bioavailability. Therefore, extensive studies have focused on overcoming the problems and optimizing the therapeutic potential. Currently, numerous peptide-based drugs are being marketed thanks to new synthetic strategies for optimizing metabolism and alternative routes of administration. This tutorial review briefly introduces the history and role of natural opioid peptides and highlights the key findings on their structure-activity relationships for the opioid receptors. It discusses details on opioid peptidomimetics applied to develop therapeutic candidates for the treatment of pain from the pharmacological and structural points of view. The main focus is the current status of various mimetic tools and the successful applications summarized in tables and figures.

Graphical Abstract

1. Introduction

Peptidomimetics are synthetically altered peptides with adjusted molecular properties for specific biological or therapeutic applications and have been an important class of drug molecules due to their potential features, high potency, and low toxicity since the term was created first in the late 1970s [1]. Although endogenous opioid peptides such as endorphin (END), enkephalins (ENKs), and dynorphins (DYNs) play various complex roles in the body, they have been shown to have a key role in regulating pathological states of pain as well as in multiple behavioral processes: addiction, reward, sedation, and depression. Their activities in the central nervous system (CNS) have been of particular interest for the treatment of pain, considering the centrally mediated actions of the pain process. Despite their critical roles, the endogenous peptides have limited clinical use due to their lack of drug-like properties: low metabolic stability, poor bioavailability, and low blood-brain barrier (BBB) permeability [2].
To overcome the intrinsic disadvantages and to optimize the therapeutic potential, various strategies have been applied to the design of novel opioid peptidomimetics through in-depth structural analyses and structure–activity relationships (SAR) studies while maintaining key structural features responsible for the biological activity. These, in general, include deletion, addition and/or replacement of certain natural amino acids with non-natural amino acids, and introduction of steric constraints into flexible peptide bonds, which resulted in the discovery of various new opioid peptidomimetic structures with enhanced efficacy and therapeutic potentials.
This article reviews the current development of opioid peptidomimetics from the pharmacological and chemical points of view, which aims toward the development of opioid peptide-derived therapeutics for the treatment of pain. It includes typical mimetic strategies of which the representative molecules with drug-like agonist or antagonist activity are selected and discussed. Emphasis is given to the relatively short length (up to around ten residues) of structures with high receptor selectivity and potency or unique biological profile.
In the following second chapter of this review, opioid receptors and natural peptides are described together with a brief introduction about the discovery and progress. The third chapter covers pharmacological and chemical structural aspects of opioid peptidomimetics. In the former part, pharmacological efforts aimed at developing molecules with enhanced biological activity, receptor selectivity, metabolic stability, BBB permeability, and oral bioavailability are discussed together with successful examples. In the latter part, chemical aspects of opioid peptidomimetics are discussed based on various structural approaches and representative molecules. Starting with the introduction of the well-known “message–address” concept for the opioid receptor-peptide interactions, the part discusses opioid peptidomimetics at MOR, DOR, and KOR for which the relationships to natural peptide scaffolds and activities have been established. The part also outlines useful tools for the opioid peptidomimetic design: local modifications, global restrictions, and secondary structure mimetics. These include the cyclization, N/C-terminal modifications, peptide linking, local backbone modifications, simple non-natural amino acid replacements, etc. Successful cases are discussed in detail, sometimes with failed ones that are worth introducing. Finally, it details prospective aspects of opioid peptidomimetics as therapeutics aiming to inspire future research.

2. Opioid Receptors and Natural Opioid Peptides

Opioid receptors which belong to G-protein coupled receptors are known to be involved in pain modulation, numerous physiological functions, and behavioral effects and are characterized in three subtypes, mu- (MOR), delta- (DOR), and kappa-opioid receptor (KOR) with overall 60–65% high structural homology [3]. The extracellular region has much lower homology, and the differences in the region are responsible for the subtype-selectivity of endogenous opioid peptides [4]. There are three main families of the endogenous opioid peptides, END, ENK, and DYN, which are derived from three different precursor proteins, pro-ENK, pro-DYN, and pro-opiomelanocortin, and prefer to bind at the MOR, DOR, and KOR, respectively, with low selectivity but strong analgesic effects in vivo with milder side effects, unlike morphine [3]. Regardless of the receptor selectivity, all of the endogenous opioid peptides share the same N-terminal tetrapeptide sequence (YGGF) that acts as the message part for the receptor, while their C-terminal acts as the address part for selectivity (Table 1).
Since their discovery, endogenous peptides, mostly smaller peptides, have been studied extensively to discover a safe analgesic with minimal alkaloid opioid-related adverse effects. Although END was found to be the most potent analgesic due to high MOR selectivity, fewer studies have been done due to the relatively larger size [5]. A SAR study of END showed that DAla substitution at position 2 did not change metabolic stability and biological activity, contrary to the ENK analogs that enhanced the stability and activity [6].
ENKs that are slightly selective for DOR over MOR have been selected as a good scaffold to modify for druggable molecules due to the small size, lower toxicity, and lack of MOR-caused adverse effects. Activation of DOR with an agonist is not strongly analgesic, like that of MOR, but seems to cause less addictive and relatively fewer severe side effects [7,8,9]. C2,5[DPen2, DPen5]-ENK (DPDPE), [DAla2, DLeu5]-ENK (DADLE), and [DSer2, DLeu5]-ENK (DSLET) are representative molecules modified from the ENK structure showing enhanced selectivity and affinity with full agonist activity for the DOR [10,11,12]. Interestingly, [DAla2, NMePhe4, Gly-ol5]-ENK (DAMGO) is a highly selective MOR agonist derived from ENK structure as well [13].
DYNs that are formed by cleavage of the precursor prodynorphin contain a high proportion of basic amino acids and exert their analgesic effects primarily through the KOR with slight selectivity over the other subtypes. Activation of the DYN A/KOR system produces similar actions to other opioids but also opposite ones to those of MOR [14]. The KOR appears to be involved in reward, mood state, and cognitive function in the CNS, and its inhibition by an antagonist has recently drawn more attention as a therapeutic target for the treatment of stress-related mood disorders, drug-seeking, and relapse [14,15,16,17,18]. Nonetheless, KOR agonists are still useful if the activity is localized in the peripheral region due to the potential to avoid serious central side effect, dysphoria. The selectivity of DYN for the KOR is known to come from C-terminal region including Arg6 to Lys11, and DYN A-(1-11) is identified as a short fragment that retains similar KOR agonist activity to that of DYN A [19]. Its further modifications led to the development of potent and selective KOR agonists (ex. N-benzyl[DPro10]-Dyn A-(1-11)) as well as antagonists (ex. N-benzyl-c5,8[DAsp5, Dap8]-DYN A-(1-11)-NH2 (Zyklophin)) [18,20].
Endomorphin (EM)-1 and -2, which are structurally distinct, are atypical endogenous opioid peptides consisting of more constrained amino acid residues (Pro-Trp and Pro-Phe, respectively) in the middle positions and showing higher MOR selectivity over DOR and KOR [21,22,23]. These endogenous peptides are localized in the CNS associated with pain mechanism and in the ending of sensory neurons, and their interactions with the opioid receptors produce clinically effective analgesia, which seems to be dissociated by rewarding, and respiratory depression [23,24,25]. However, their low ability to penetrate the BBB and low metabolic stability prohibit therapeutic uses as analgesics, and therefore significant efforts have been made to improve metabolic stability and to obtain longer lasting antinociceptive effects through various modifications [26,27]. Cyclic EM-1 analogs such as ZH853 (Tyr-c[DLys-Trp-Phe-Glu]-Gly-NH2) are recent discoveries showing longer duration of action and effective antinociception in multiple pain models with reduced adverse side effects [28,29,30].
Other naturally occurring peptides are dermorphines (DERs) and deltorphins (DLTs), which are isolated from frog skin and milk peptides and are the most selective for MOR and DOR, respectively [31,32]. The amino acid sequences of these two peptides are quite different from the other endogenous opioids and have been extensively used as a starting point for the development of highly selective opioid peptidomimetics. The N-terminal tetrapeptide of DERs, YaFG, is a minimum structure for the MOR activity and its modified analog, [DArg2, Lys4]-DER (DALDA), is a highly selective MOR agonist but has a limit in crossing the BBB because of multiple positive charges [33]. DLTs are the most selective naturally occurring opioid peptide with a high affinity and potency for the DOR as well as a high BBB penetration rate [34].

3. Peptidomimetics for Opioid Receptors

3.1. Pharmacological Aspect

3.1.1. Improving Bioactivity, Specificity, and Selectivity for the Opioid Receptors

Structural homology and the natural flexibility of endogenous linear opioid peptides limit their uses for a specific subtype receptor because of the possible interactions with more than one subtype receptors and drastic SAR results caused by subtle structural change. For this reason, selectivity, mostly between MOR and DOR, has been a major problem for the linear opioid peptides, and tremendous efforts have been made to reduce their structural flexibility and thereby enhance the subtype selectivity and potency. Incorporation of a constrained amino acid, configurational change from L to D, and cyclization of a backbone have been the most useful tool to achieve high selectivity and specificity. Figure 1 displays highly selective opioid peptidomimetics derived from ENK and DYN A for respective subtypes with high potency. DAMGO is a potent and selective MOR agonist to be rendered by introducing an N-methyl group into ENK to reduce rotational freedom around the peptide bond (φ/ψ constraint) [13]. Its crystal structure bound to the MOR showed a very similar pose in the binding pocket to those of other small molecule morphinans [35]. DPDPE is the most successful example of cyclization through a disulfide bond that constrains the ENK to an active conformation state for the DOR and has been used the most for scientific purposes [36,37]. It is noteworthy that both selective molecules have originated from less selective ENK. [DAla3]-DYN A-(1-11)-NH2 is a highly selective KOR agonist with a similar affinity to DYN A-(1-11)-NH2 showing how simple substitution can affect receptor selectivity drastically [38].

3.1.2. Transforming Bioactivity from Agonism to Antagonism

While opioid agonists are used for pain relief, antagonists are also substantial in preventing or reversing the opioid agonist effects and intoxication as well as in pharmacological studies to investigate any endogenous opioid regulation. However, there is no opioid antagonist that naturally occurs, and currently available ones have been developed from opioid agonists by structural modifications. In general, agonist versus antagonist behavior of opioid peptides depends on very subtle structural differences, as shown in many cases [3,39,40].
The most effective and generalized structural modification for the conversion into an antagonist is the deletion of an Nα-amino group or its replacement with the other groups such as allyl and hydroxybenzyl [39,41,42]. These modifications similarly affect the three subtypes receptors. Substitution of Tyr1 with 3-(2,6-dimethyl-4-hydroxyphenyl)propanoic acid (Dhp) or (2S)-2-methyl-3-(2,6-dimethyl-4-hydroxyphenyl)propanoic acid (Mdp) converted opioid agonist peptides to corresponding antagonist peptides with high potency and selectivity [43]. The substitution of position 2 with 1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid (Tic) in ENK and DER analogs converted their respective agonist activities at DOR and MOR to DOR selective antagonists, and NMR analysis indicated a 90-degree arrangement of the two aromatic rings in the cis-Tyr-L-Tic moiety responsible for the conversion [44]. It was also shown that modifications of EM-2 and morphiceptin (YPFP-NH2) at position 3 with DPhe or D-2-Nal converted the MOR agonist activity to the antagonist activity [40]. Details are discussed in the later part, 3.2. Structural aspects.

3.1.3. Optimizing Drug-like Characteristics including Metabolic Stability, BBB Permeability, and Oral Bioavailability

Clinical application of opioid peptides has been limited due to poor metabolic stability in the body and low ability to cross the BBB [2]. The ability to cross the BBB and reach the CNS to interact with opioid receptors for analgesic effects relies on physicochemical properties including size, lipophilicity, enzymatic stability, hydrogen bonding potential, and other structural features [45]. Physicochemical properties of opioid peptides have been far from optimal and have restricted their oral bioavailability. The nonpeptidic nature of peptidomimetics has the potential to overcome the detrimental therapeutic characteristics and prolong and enhance biological activities at the target site [2,46]. Therefore, various peptidomimetic approaches have been devised to develop molecules with optimized CNS activity, although there are still no definitive alternatives to using the classical analgesics for pain treatment [45,47].
The difficulty of eliciting analgesia with three endogenous opioid peptides is mainly due to the rapid cleavage of the Tyr1-Gly2 bond, and simple modification of Gly2 with DAla2 in ENK resulted in long-lasting analgesic effects by avoiding the critical degradation [47,48]. Replacing a peptide bond susceptible to peptidases with a dipeptide isostere such as olefin, ester, and triazole has also been a useful tool to increase peptide stability as well as conjugating with antibodies [47,49,50]. A recent study showed that β-Ala at the N-terminus of an ENK-like tetrapeptide amide resulted in longer analgesic effects owing to less accessibility of endopeptidases [51]. Another recent study showed that Nα-guanidinyl group at the N-terminus, which mimics the ionic state of the parent peptide, significantly improved the stability and lipophilicity as well as affinity and potency, when combined with C-terminal tetrazolate and DAla in guanidyl-Tyr-DAla-Gly-Phe-Leu-tetrazole (Figure 2) [52].
Although metabolic stability has been improved by various structural modifications, the efforts to improve the drug-like property of opioid peptides have been made with limited success for the CNS analgesic effect due to the poor permeability across the BBB. Pharmacological approaches to enhance a peptide delivery to the CNS are to increase biochemical attributes by modifying its structure or conjugating a molecule like lipophilic enhancer, polymer, and antibody. Increasing lipophilicity has been the most useful tool for BBB penetration despite the risk of plasma protein binding [45,53,54]. DPDPE with poor BBB permeability has been modified by methylations, halogenations, Nα-acylation, and C-terminal esterification, and the resulting lipophilic analogs have produced significantly increased central analgesic effects as well as BBB permeability [53,55,56,57]. Unlike bis[Tyr-NH]-alkyl peptides, short bivalent analogs linked via an alkyl group, 2′,6′-dimethyl-L-tyrosine (Dmt)-substituted analogs produced strong analgesic effects after systemic (s.c. and oral) administrations through the MOR [58]. Among those, pyrazinone-containing analog (Figure 2) was the most potent analgesia in tail flick and hot plate tests. The molecular modeling studies suggested that the bis-Dmt analog contains the identical message and address regions and fits the MOR binding pocket, unlike large dimeric ENK or DER analogs [59]. A study showed that simple C-terminal esterification of EM can significantly increase lipophilicity, metabolic stability, and systemic antinociceptive activity after i.c.v. or oral administration, although it provided no explanation on how the change in physicochemical properties are responsible for the side effects [60].
Interestingly, [Dmt1]DALDA, a highly polar ligand with a 3+ net charge, produced 36 times more potent antinociceptive effects than morphine in mice after subcutaneous (s.c.) administration indicating high BBB transport to the brain [61]. Many studies also demonstrated that hydrophilic peptides could penetrate the cell membrane via absorptive-mediated endocytosis. The common feature was that those peptides contained multiple Arg and/or Lys. E-2078 is also a polycationic DYN A-(1-8) analog internalizing to brain capillaries by the endocytosis [62]. MOR selective DER analogs, H-Tyr-DArg-Phe-Sar/β-Ala-OH, showed potent analgesic effect with low physical and psychological dependence. Their modifications with Nα-amidinoTyr1 resulted in the slower onset of the analgesic effect after systemic administration because of the loss of a positive charge and the resulting slow BBB transport [63,64]. Based on the observation, modifying the positive charges might be a useful tool to control the BBB transport and the pharmacological effects in the CNS.
The addition of carbohydrates to a peptide, glycosylation, has also been used to improve BBB penetration and enzymatic resistance while maintaining the biological activity of a parent peptide. The applications for ENKs, DER, DLT, and EM have been successfully performed to enhance central analgesic effects after systemic administration and oral bioavailability [65,66,67,68,69,70,71]. A glycosylated hexapeptide, H-Tyr-DThr-Gly-Phe-Ile-Ser(β-DGlc)-NH2 (Figure 2), increased metabolic stability and BBB penetration as well as systemic (i.v.) antinociceptive activity, and its further modification with β-lactose and β-melibioside produced highly potent antinociception [67]. The same efforts, glucosylation or galactosylation at the C-terminal region, were made for DER and DLT and resulted in a remarkable increase of antinociceptive activity following systemic administration while retaining high in vitro activity as well [72,73]. Thr4 glycosylated analog, however, reduced activity dramatically, indicating the important role of the position for opioid activity. Orally active EM-1 analog in which a lactose moiety was linked to the N-terminus through a succinamic acid spacer showed a strong antineuropathic effect without producing constipation [71]. A cyclic glycosylated pentapeptide H-Dmt-c2,4(-SCH2CH2S-)[DCys-Aic-DPen]Ser(Glc)-NH2 possessing potent mixed MOR agonism/DOR antagonism and poor bioavailability was also successfully modified to afford high antinociceptive efficacy after intraperitoneal (i.p.) administration without acute tolerance [74]. Modeling studies and NMR analysis of glycosylated ENK analogs indicated the modification did not disturb the peptide backbone [75].
Conjugation of polyethylene glycol (PEG) has been used to enhance the therapeutic potential of opioid peptides by reducing enzymatic degradation while maintaining the biological activity of a parent peptide [45]. I.v administration of pegylated DPDPE, PEG-CH2-CH2-CO-DPDPE (Figure 2), showed an increased analgesic effect despite 176-fold lower binding affinity than DPDPE, which indicated better ability to cross the BBB and undergo hydrolysis in the brain [76]. Nonetheless, there are potential risks of losing activity and brain uptake due to increased molecular size and hydrophilicity or improper applications. To improve the brain uptake, pegylated liposomes as carriers have also been used as a carrier, and glutathione pegylated liposomal formulation of DAMGO was shown to increase and prolong the brain uptake significantly [77].
Prodrugs have been designed to improve the exposure of opioid peptides at target sites, mostly the brain, after systemic administration. Esterification of amine, hydroxyl, or carboxylic acid improves brain uptake by increasing lipophilicity. Likewise, the addition of a Phe residue to DPDPE was shown to increase the BBB permeability as a prodrug [78]. Cyclic prodrugs have also been built using linkers susceptible to esterase hydrolysis for improved delivery characteristics [79,80,81]. For a prodrug, its bioconversion to an active parent peptide within the target sites is the most critical as well as stability in the blood. A study on cyclic prodrugs of DADLE was not successful due to the formation of stable intermediate and the substrate activity for efflux transporters on the BBB nonetheless improved metabolic stability (Figure 2). The enzymatic bioconversion rates were dependent on the chirality of specific amino acids, and therefore alterations in their chirality might be critical for the success of prodrug [82].
Although oral delivery is the most preferred mode of drug administration, the bioavailability of peptides remains low due to degradation and limited absorption in the gastrointestinal tract. The development of oral delivery for opioid peptides requires the same strategies as those applied to optimize the BBB permeability, such as increasing lipophilicity and structural constraint. An amine-modified prodrug of ENK by a reversible aqueous lipidization (REAL) (Figure 2) approach produced prolonged strong oral antinociceptive effects in an inflammatory pain model along with increased metabolic stability, gastrointestinal absorption, and limited CNS-penetration [83]. Thiazole-containing cyclic DAMGO, N-terminal amidated-DERs, and phenolic group acylated-DERs are successful examples showing good oral analgesic effect [84,85]. Likewise, Nα-glycosylated EM showed a strong antinociceptive effect after oral administration in the chronic constriction injury (CCI) model but no significant constipation in contrast to morphine [71].

3.1.4. Reducing Opioid Side Effects

There appears to be no substitute for opioids in achieving adequate pain relief in some cases. MOR is primarily responsible for the antinociception but also causes the most undesirable adverse effects, limiting its clinical use. Acute administration causes respiratory depression, constipation, sedation, dizziness, and nausea, and chronic use causes tolerance, dependence, and abuse liability. Despite extensive research in the field, reducing the serious adverse effects of the MOR agonists remains unsolved. The DOR and KOR are also involved in the antinociceptive effect and related side effects, such as convulsion and dysphoria, respectively, to a lesser extent, in vivo. There have been several strategies to design opioid peptidomimetics with reduced adverse effect occurrence [86]. Those are to develop novel ligands that target multiple subtype opioid receptors, interact with receptors peripherally, produce a biased signaling, etc.

Multifunctional Ligands

Of those, multifunctional opioids that can act on more than one subtype receptor as a single molecule emerged as a promising approach to lessen the opioid-related adverse effects and provide a safer alternative to traditional opioid analgesics [87,88,89]. Mixed agonist activity at the MOR and DOR may bring synergistic antinociception that can reduce adverse effect occurrence by reducing the amount given for the same effect. It was shown that the occupation of DORs by an agonist or antagonist prevented tolerance and physical dependence on morphine [90]. Based on the involvement of DOR in MOR activity proven in numerous studies, such bifunctional ligands with a MOR agonist/DOR antagonist property as AAH8, UFP-505, KSK-103, DIPP-NH2[Ψ], [Dmt1, 2′,4′,6′-trimethyl phenylalanine (Tmp)3]-EM-2, or with a MOR/DOR agonist property as biphalin, [Dmt1, 2′-ethyl-6′-methylphenylalanine (Emp)3]-EM-2, and LYS739 seemed to be a good approach [74,87,91,92,93] (Figure 3). Biphalin is a homo-bivalent ligand consisting of two tetrapeptides and showing MOR/DOR agonist activities [94]. It produces a substantial antinociceptive effect but causes fewer side effects, suggesting synergistic effects obtained from the activation of both receptors [94,95].
A glycosylated DTLES analog, NMP-2200 (YrGFLS(OD-lactose)-NH2) is a centrally active MOR/DOR agonist demonstrated to reduce addiction liability associated with the MOR agonist analgesics through the simultaneous activation of the MOR and DOR [68,96]. Alteration of EM-2 with a Dmt1 and an alkylated Phe3 (Dmp, Tmp, Emp) led to the bifunctional activities depending on the substitution of Phe3: di- and tri-methyl phenylalanine for the DOR antagonism and ethyl methyl phenylalanine for the DOR agonism and increased DOR affinity, yet retained MOR activity, leading to potent MOR/DOR agonism or MOR agonism/DOR antagonim [97]. Metkephamid (H-Tyr-DAla-Gly-NMeMet-NH2) is a balanced MOR/DOR agonist producing central analgesic effects after systemic administration in animal models [98,99]. It showed a lesser degree of side effects in extensive clinical tests, indicating DOR activation effect on the CNS. Clinical development, however, was stopped after phase 1 due to unusual side effects, which might be caused by the DOR activation in the CNS. For this reason, MOR agonist/DOR antagonist has been pursued more vigorously than MOR/DOR agonist.
TIPP-NH2 was the first to show a mixed MOR agonist/DOR antagonist property, and its further modifications resulted in the discovery of H-Dmt-Tic-[CH2NH]Phe-Phe-NH2 (DIPP-NH2[Ψ]) possessing a balanced MOR agonist/DOR antagonist property and exhibiting a reduced tolerance and dependence upon chronic administration [100]. Several other mixed MOR agonist/DOR antagonists containing a Dmt-Tic pharmacophore were also developed. Intrathecal (i.th.) injection of UFP-505 (H-Dmt-Tic-Gly-NH-Bzl) showed less tolerance in rats than morphine [101]. A 2-aminoindane (Aic)-substituted cyclic analog (KSK-103, Dmt-c(-SCH2CH2S-)[DCys-Aic-DPen]-OH) with a poor bioavailability also developed fewer tolerance and reward symptoms through the MOR agonist/DOR antagonist activities after C-terminal glycosylation [74,102]. Recently, numerous stable structures of MOR agonist/DOR antagonist have been developed showing central analgesic effects after systemic administration, but no in vivo study on the side effects has been reported [92,103,104,105].
Other subtype combinations are MOR/KOR, DOR/KOR, and MOR/DOR/KOR proposed for reducing side effects, but little peptidomimetic structure has been identified [87,106]. Cyclic EM-2 analog, H-Dmt-c[DLys-Phe-2′-MePhe-Asp]-NH2 was an agonist for MOR, DOR, and KOR possessing strong antinociceptive effects potentially through the concomitant activation of three receptors [107]. The substitution of β-amino acids has been shown to alter a peptide’s selectivity to a mixed receptor property. Likewise, modifications of a MOR selective morphiceptin with β2- or β3-amino acid residues resulted in various multifunctional opioid profiles. Dmt-DAla-(R)-β2-1-Nal-Pro-NH2 was a potent MOR/DOR/KOR agonist exhibiting a strong peripheral antinociceptive effect after i.p. and oral administration [108].

Localization of the Site: Peripherally Mediated Analgesics

Emerging evidence indicates that opioids also act in the periphery to contribute to analgesic actions, although much less is known about this, compared to the central functions [86,109]. During inflammation, peripheral opioids interact with upregulated opioid receptors in damaged tissue, and the localized interactions can induce therapeutically safe analgesic effects by avoiding undesirable centrally mediated adverse effects [86,109]. The main strategy to localize the interactions in the PNS was to increase the hydrophilicity of the opioid peptides to inhibit the brain uptake. It afforded numerous peripherally acting opioid peptidomimetics with a variety of biological profiles [110].
[β-Pro2]-EM-1 is a potent MOR agonist showing peripheral antinociceptive activity after systemic administration along with increased metabolic stability [111]. DAMGO, CR845 (Difelikefalin, H-DPhe-DPhe-DLeu-DLys-[γ-(4-N-piperidinyl)amino carboxylic acid]), DALDA, PL017 ([NMePhe3, DPro4]morphiceptin), CJC-1008, and BW443C (H-Tyr-DArg-Gly-Phe(p-NO2)-Pro-NH2 were shown to produce peripherally localized analgesic effects [112,113,114]. Preliminary clinical studies of BW443C demonstrated peripheral opioid analgesic effects avoiding relevant central effects [115]. CR845, the first approved opioid peptide drug, is in medical use of moderate to severe itching and is under development for the treatment of postoperative and osteoarthritis pain [116]. CJC-1008, a maleimido propionyl group-attached DYN A-(1-13) analog, was designed to promote covalent bond formation with serum albumin and showed a greater peripheral analgesic effect compared to placebo with prolonged activity in Phase II clinical study [113]. Likewise, the conjugation of a biocompatible nano-carrier with a big molecular size also localizes the opioid peptides by blocking the BBB transport [50,86,109].

Biased Ligands, etc.

In β-arrestin-2 knockout mice, morphine enhanced analgesia with reduced constipation and respiratory depression [117]. Based on the result, biased ligands that preferentially activate one downstream pathway, G-protein pathway over β-arrestin, have been suggested as a safer, better tolerated, and more efficacious opioid analgesic despite a recent study to the contrary [118,119,120,121]. A study discovered a cyclic peptide, H-Dmt-c[DLys-Phe(p-CF3)-Phe-Asp]-NH2, with high MOR affinity and in vitro functional activity, which turned out to be a G-protein biased MOR agonist [122]. Studies suggested positions 4 and 5 of ENK might be responsible for the biased signaling because the substitutions at the positions resulted in distinctly biased signaling [123,124,125]. Furthermore, positive allosteric modulators were shown to enhance the activities of endogenous opioid peptides, maintain their temporal and spatial action, and potentially limit the adverse effects [126]. Although a few allosteric modulators were identified for opioid receptors and characterized in vitro, their utility in vivo is yet to be determined [127].

3.2. Structural Aspect

3.2.1. Conformational Studies and Design from Pharmacophore

The hierarchical approach to peptidomimetics is to (i) identify the conformational structure along with the side chain functional requirements for a target receptor, (ii) reflect the outcomes in building a constrained structure, and (iii) determine the three-dimensional arrangement of the critical side chain and backbone functionalities. Based on this, constrained peptidomimetics in which the peptide scaffold is replaced globally or locally (at a particular amino acid residue or peptide bond) by other organic moieties are designed and synthesized [49,128]. The use of constrained peptidomimetics, new insight into their interactions with opioid receptors, and additional fine-tuning of the constrained structure are critical in developing a novel therapeutic analgesic agent. Recent advances in structural biology using crystallography, molecular docking, molecular dynamics simulations, and NMR spectroscopy provided key insights into the binding modes of opioid peptidomimetics to the receptors. Interestingly, a study using cryo-electron microscopy indicated that the N-terminus of DAMGO interact with conserved receptor residue of the morphinan ligand pocket while the C-terminus occupies the regions for MOR selectivity [35].
Regardless of the sequences, natural opioid peptides contain multiple aromatic amino acids in common as key pharmacophoric residues, and their SARs have long been analyzed based on the message–address concept (Figure 4). The message region of opioid peptides is the N-terminal three or tetrapeptide responsible for biological activity, and the address region is the variable structure following the message region responsible for receptor selectivity [129,130,131,132]. In the message region, Tyr and Phe are key pharmacophoric residues, and their Nα-amino, phenolic, and aromatic groups are known to be critical for receptor recognition. Relative orientation and length of the two aromatic rings are also critical for the subtype selectivity. Overall longer backbones are preferred for MOR and shorter ones for DOR. DOR or KOR selectivity over the other subtypes is attributed to the C-terminal region, as shown in DLT (the hydrophobic Val5-Val6 residues) and DYN A (the basic Arg6-Lys11 residues), respectively. The message–address concept has contributed to the design of highly potent and selective opioid peptides and their modifications for the development of novel opioid peptidomimetics [33,132,133,134].

3.2.2. Peptide Scaffolds as a Starting Point for Mimetics: EMs, DER, Morphiceptin, ENKs, DLTs, and DYN

MOR Peptidomimetics

EMs have been suggested to inhibit pain with reduced side effects, particularly the reward effect and respiratory depression [24,25]. Their simple structures, which differ from other endogenous opioids, and high MOR specificity over DOR and KOR have stimulated the structural modifications for overcoming their limits, mainly metabolic instability and low BBB permeability, as therapeutic agents [23,71,107,135,136]. L-configuration of Pro2, which is a spacer between two aromatic amino acid residues, was shown to be critical, and numerous chemical modifications were focused on the position [23]. NMR analysis and X-ray crystallography revealed that it folded into cis-form around the Tyr-Pro amide bond and formed β-turn structure [137,138,139]. Substitution of Pro2 with a pseudoproline, β-Pro, Cα,α-disubstituted glycine, alicyclic β-amino acids, or other substituents was well tolerated, and piperidine-3-carboxylic acid (Nip) derivative, [(R)-Nip2]EM-2 showed picomolar range of high affinity for MOR (Figure 5 and Table 2) [133,136,140,141,142,143,144]. Further substitution of Tyr1 with Dmt1 led to a potent MOR agonist with high stability and analgesic effect, while the selectivity was reduced by increasing DOR affinity [145]. The role of Phe4 residue, which is the only distinct residue from morphiceptin, is still not identified, although the majority of results have shown little effect on MOR activity. C-terminal modification of Dmt-Pro-Phe-NH-X with various aromatic or aliphatic groups showed the picomolar range of high affinities at MOR as agonists together with weak DOR antagonist activities [146]. Due to its essential role, Tyr1 has been the most conserved residue, but several studies showed that its substitution could be tolerated in some cases. c[Tyr-DPro-DTrp-Phe-Gly] contains a Gly bridge for lipophilic property and showed the nanomolar range of binding affinity at the MOR nonetheless lacking the protonable amine [147]. The Phe residue at position 3 is also critical for MOR, and its substitution with sterically hindered Phe residue improved MOR affinities while abolished DOR affinities [148]. The suitable spatial arrangement and conformational restriction of the Phe3 seemed to be significant for the MOR activity.
DER is a potent and selective MOR agonist showing a strong analgesic effect after peripheral administration due to slow metabolic degradation [149,150]. The N-terminal tetrapeptide is the minimal pharmacophore for the activity, and the D-configuration of the second position is critical for the MOR agonist activity [149]. DER tetrapeptide analogs containing DArg2 showed very potent and long-lasting antinociceptive activity with various peripheral injections due to high resistance to enzymatic degradation [32,149]. It has been derivatized over the years to yield other highly selective agonists such as [DArg2, βAla4]-DER-OH (TAPA), [DArg2, Sar4]-DER-OH (TAPS), N-amidino-[DArg2, MeβAla4]-DER-OH (ADAMB), and DALDA [33,85,151,152]. I.th., i.c.v., or s.c. administration of TAPA produced very potent and long-lasting antinociceptive effects with reduced physical dependence but very low bioavailability [63,151,153]. ADAMB is the most potent DER analog with increased oral bioavailability and slow onset of antinociceptive effect after systemic administration [64,85]. DALDA that contains two positively charged amino acid showed extremely high selectivity for MOR but relatively lower antinociceptive effect after systemic administration due to low BBB permeability [33]. [Dmt1]-DALDA displayed very potent antinociceptive effects after systemic administration due to increased resistance to enzymatic degradation, and further modifications resulted in an extremely potent MOR/DOR agonist, Dmt-NMe-DAla-[(S)-4-amino-1,2,4,5-tetrahydro-2-benzazepine-3-one (Aba)-Gly]-NH2, with picomolar potencies [154,155]. Interestingly, these DER analogs containing DArg2 showed distinct antinociceptive profiles from the traditional MOR agonists: release of DYN and suppression of the antinociceptive effects by KOR antagonists, as observed in EM-2 [149].
Morphiceptin is a MOR selective agonist identified from β-casein with a unique structure, two Pro residues which allow cis/trans isomerization around the amide bonds between positions 1 and 2, and 3 and 4 [129,156]. The L-configuration of Phe3 was not critical and its modifications with DPhe or D-1-Nal improved analgesic effects significantly (>150 folds), and cis-2-aminocyclopentane carboxylic acid (βAc5) substituted analogs improved binding affinities through the formation of MOR favorable cis-conformation [40,129,137,157,158,159]. Recent multiple modifications with Dmt1, DAla2, (R)-β3-1-Nal3 yielded a highly selective and potent MOR agonist, while substitution with (R)-β2-1-Nal at position 3 produced mixed MOR/DOR/KOR agonist activity showing strong peripheral antinociceptive effects after i.p. or oral administration [108].
Deletion of the Nα-amino group of an opioid agonist or its replacement with the other groups such as allyl and hydroxybenzyl has been utilized for the development of an opioid antagonist with limited success [39,41]. The modifications in Dhp1-c[Nε,Nβ-carbonyl-DLys2,Dap5]-ENK-NH2 and [N,N-bis(p-hydroxybenzyl)-Gly]-DArg-Phe-βAla-NH2 were less successful due to the diminished MOR selectivity [43,160]. Successful modifications occurred by N-allylation in [Dmt1]-EMs. The N-allylation was the first successful case to convert an alkaloid MOR agonist into an antagonist, which can be a useful tool to investigate the pharmacological role of MOR and the related addiction and abuse [39]. H-DPhe-c[Cys-Tyr-DTrp-Arg/Orn-Thr-Pen]-Thr-NH2 (CTAP/CTOP) is a commonly used, highly selective MOR antagonist derived from somatostatin, and its structural dissimilarity to known opioid peptides made it an attractive starting point. The substitution of N-terminal DPhe with a constrained analog, DTic, led to the most potent and selective MOR antagonists TCAP and TCOP [161].

DOR Peptidomimetics

A potential advantage of targeting DOR in the modulation of pain is reducing opioid-related side effects such as respiratory depression, addiction, and immune functions, and therefore its endogenous ligands, ENKs, have been extensively studied to develop a potent DOR agonist [7,8,9,20,162]. The Phe residue at position 4 was shown to be critical for the DOR activity and selectivity together with Tyr1 and C-terminal carboxylate [162]. The replacement of Gly at position 2 with DAla reduced DOR selectivity in numerous ENK analogs by increasing MOR activity [163]. However, further replacement of a Leu residue with a DLeu in DADLE recovered the DOR selectivity and produced highly potent agonist activity. Interestingly, C-terminal truncated [DAla2]-ENK tetrapeptide retained the DOR activity, unlike its parent tetrapeptide. Substitutions at the positions other than position 2 with more lipophilic and constrained amino acids occasionally increased binding affinity, selectivity, and potency at the DOR. Despite the critical role of Tyr1 and Phe4, the first and third amide bonds seemed to be replaceable with other functional groups such as a fluoroalkene without a large loss in the activity [164,165,166]. However, the relative proximity between two aromatic rings at positions 1 and 4 was considered to be critical for the DOR versus MOR selectivity. Based on this, more constrained cyclic analogs were developed using a disulfide bond.
Considering its high selectivity and potency, DPDPE is a major success in DOR peptidomimetic and has been the most commonly used for the studies of biochemical, pharmacological, and physiological properties [20,36,167]. [Des-Gly3]-DPDPE (JOM-13, Tyr-c[DCys-Phe-DPen]-OH) designed for additional conformational restriction retained good selectivity and affinity for DOR (Figure 5 and Table 2) [168]. In this scaffold, increasing the ring size by inserting an alkyl group into the disulfide bond resulted in the decrease of DOR activity and selectivity. JOM-6 (Tyr-c2,4(SCH2CH2S)[DCys-Phe-DPen]-NH2) with a dithioether ring was a selective MOR agonist. Various modifications at position 4 with constrained amino acids were not successful due to its important role except p-alkyl substituted or β-methyl substitution. Interestingly, [Dmt1/β-methyl-2′,6′-dimethyl tyrosine (Tmt)1, Phe(X)3]-DPDPE resulted in highly potent and selective MOR agonists with the improved drug-like property [20]. Conformational analysis of DPDPE and side chain topography led to the development of nonpeptidic or peptidic DOR ligand SL-3111 with 2000-fold selectivity (Ki = 8.4 nM) for the DOR [169].
DLTs (DLT A, -1 and 2) are highly selective and potent DOR agonists containing D-amino acid and anionic residues responsible for the distinctive biological and structural features [31]. Moreover, these peptides are enzymatically stable and permeable to the BBB due to their structural features [34]. Their message region, Tyr-DAla-Phe-, was shown to form a β-turn structure stabilized by the DAla, and the substitution of a Phe residue with a more constrained β-iPrPhe residue enhanced the conformation resulting in the exceptional selectivity (30,000 fold) over MOR [170]. The C-terminal address region was attributed to the DOR selectivity through the hydrophobic Val5-Val6 residues, which stabilize an optimal conformation for the binding rather than interact with the receptor [31,171]. The deletion of the anionic character at position 4 resulted in the loss of selectivity due to increased MOR activity. The application of 1-aminocycloalkane-1-carboxylic acids at positions 2, 3, and 4 retained its activity at the DOR, and deletion of the anionic character (Asp4) resulted in the enhancement of MOR activity, indicating unfavorable repulsion between negative ion and the MOR receptor [172]. Contrarily, positive net charges were favored for the MOR over DOR, as shown in DALDA [149]. Most cyclic analogs reversed the selectivity due to the marked decrease in DOR binding while some were MOR selective [31]. c[DCys2, Pen5]-DLT-1 was a highly selective DOR agonist, but c[DCys2, Cys5]-DLT-1 was a mixed MOR/DOR agonist [173].
N,N-diallylation of Leu-ENK analog resulted in the discovery of highly selective DOR antagonist (Ke = 22.0 nM, 227-fold selectivity over MOR), ICI174864 (N,N-dially-Tyr-Aib-Aib-Phe-Leu-OH) [174]. There are a greater number of specific DOR antagonists than MOR antagonists that exhibit potent bioactivities. H-Tyr-Tic-Phe-Phe-OH (TIPP) and TIP containing a constrained Phe analog, Tic, at position 2 are potent and selective DOR antagonists while DTic-substituted analogs act as a mixed MOR/DOR agonist, showing the configuration-dependent differential effect on selectivity and intrinsic activity [175]. These have been the molecule of interest for SAR studies to investigate the agonist and antagonist properties of the receptor as well as to improve metabolic stability [100,175,176]. Truncation of the C-terminal dipeptide led to small dipeptide analogs with diverse biological profiles, DOR agonist or antagonist, or MOR agonist/DOR antagonist. H-Tyr-Tic-NH-CH2-CH(Ph)2 and H-Dmt-Tic were a selective and potent DOR agonist and antagonist, respectively [100,176]. Simple modification of the C-terminal carboxylate of TIPP to the carboxamide produced a mixed MOR agonist/DOR antagonist activity, and further modification with Dmt1 resulted in a bifunctional activity in DIPP-NH2 with a potent analgesic effect. H-Dmt-Ticψ [CH2NH]Phe-Phe-NH2 (DIPP-NH2[ψ]) designed to reduce chemical degradation via diketopiperazine formation showed more potent i.c.v. analgesic effect in rat tail-flick test with no physical dependence and less tolerance than morphine [100,176,177].
Dmt-Tic is the shortest peptide derived from TIPP possessing a highly selective and potent DOR antagonist activity, and its modification showed a variety of MOR and/or DOR selectivity and agonist and/or antagonist activities [178,179,180]. (S,S)-isomer of Aba-Dmt moiety, where a Tic was replaced and the backbone direction was reversed, turned out to be a potent MOR antagonist (Ke = 39 nM) [181].

KOR Peptidomimetics

DYN A and its fragments, DYN A-(1-13) and DYN A-(1-11), are highly active at KOR, and the message and address regions are essential for adapting a helical conformation extending from Tyr1 and Arg9. The address region containing multiple basic amino acid residues was suggested to be responsible for the KOR efficacy. Although the relatively large size of the molecule and the extensive metabolism are more problematic, many studies demonstrated important roles of DYN A and its target receptor, KOR, in pain and other disorders such as depression, anxiety, and drug abuse, and more interest is being focused on the therapeutic potentials of KOR agonists and antagonists [182,183,184].
The low metabolic stability has been improved by various modifications such as D-amino acid residue and N-alkylated amino acid substitution, C-terminal amidation, and peptide backbone reduction. Nonetheless, low selectivity, [NMeTyr1, NMeArg7, DLeu8]-DYN A-(1–8)-NHEt (E-2078) were examples of successful modifications to impart metabolic stability in the shortest fragment together with ([DAla2, Arg6Ψ(CH2NH)Arg7]-DYN A-(1-8)-NH2 (SK-9709) (Figure 5 and Table 2) [185]. E-2078 and SK-9709 showed good antinociceptive effects mainly through the KOR in various pain models after systemic administration, indicating the ability to cross the BBB [186,187]. Highly constrained peptidomimetics of Dyn A-(1-8) with a cyclopropyl methyl normetazocine moiety, which mimics a phenolic ring, a basic nitrogen, and a phenyl ring, were potent KOR agonists with a high affinity and selectivity over MOR and DOR as well as a potent antinociceptive effect [19,188].
A KOR agonist, DPhe-DPhe-DNle-DArg-NH2, was discovered by combinatorial library tool and modified at the C-terminus to afford a peripherally acting analog, DPhe-DPhe-DNle-DArg-NH(4-picolyl) CR665 (FE200665) [189,190,191]. In various pain models, this analog showed potent peripheral analgesic and anti-inflammatory effects [192]. The modification of DArg4 with dimethyl-DLys4 produced an orally active peripherally restricted KOR agonist, CR845 [116]. It binds to nerves and immune cells in the periphery and blocks pain and inflammation without central opioid side effects.
KOR antagonistic peptides are shown to possess high therapeutic potential in treating drug abuse due to limited duration of action, unlike small molecules with an extremely long duration of action. Therefore, significant efforts have focused on the development of potent and selective antagonists using the DYN A scaffold. N-terminal monoalkylation of [D-Pro10]-DYN A-(1-11) resulted in a marked enhancement in KOR selectivity by decreasing MOR affinity and further N,N-dialkylation produced an antagonist activity [42,193]. The substitution of Pro at position 3 was also reported to be a highly selective KOR antagonist. Although a few key structural features were identified for the KOR antagonist activity, their applications to DYN A resulted in limited success with weak efficacy, low selectivity, and residual agonist property. Later, numerous peptide antagonists have emerged at the KOR, including the [(2S)-Mdp1]-DYN A-(1-11)-NH2 (Dynantin), [AcPhe1,Phe2,Phe3,Arg4,DAla8]-DYN A-(1–11)-NH2 (Arodyn), c2,5[N-BzlTyr1,DAsp5,Dap8]-DYN A-(1-11)-NH2 (Zyklophin), and c[Pro-Phe-Trp-Phe] (CJ-15208) by deleting N-terminal free amine group [18,19,43,184,194,195,196]. Nα-benzylation of a cyclic DYN A-(1-11) analog resulted in the discovery of a highly selective KOR antagonist with desirable therapeutic potentials such as systemic activity, proper duration of action, and the BBB penetration in the treatment of diseases such as depression, anxiety, and drug abuse [18,182]. Zyklophin is a selective KOR antagonist with the ability to inhibit the KOR selectively and high potential as a pharmacological tool and a therapeutic agent [18].
Figure 5. Structures of selective opioid peptidomimetic.
Figure 5. Structures of selective opioid peptidomimetic.
Biomolecules 12 01241 g005
Table 2. In vitro biological activities of representative opioid peptides and peptidomimetics at MOR, DOR, and KOR a.
Table 2. In vitro biological activities of representative opioid peptides and peptidomimetics at MOR, DOR, and KOR a.
Binding Affinity, IC50 (or Ki) [nM]Inhibitory Potency, IC50 [nM]Refs.
MORDORKORMOR
(GPI)
DOR
(MVD)
KOR
(GPI or RVD)
DAMGO1.8634560904.5032.8105[197]
EM-10.36150654304.03283-[21,39]
EM-20.69923352406.88344-[21,39]
[(R)-Nip2]-EM-20.04>1000-0.0002 b>1000 b-[144,145]
[Tmp3]-EM-20.441400-1.90>10,000-[97]
DER0.7673.5>10001.0723.2-[149,198]
ADAMB12.9>1000>1000---[85]
DALDA1.6919,2004230254781-[33,154]
Dmt-DALDA0.143210022.31.4123.1-[33,154]
Dmt-NMe-DAla-[(S)-Aba-Gly]-NH214.85.0>100,0000.00174 c0.016 c-[155]
JOM-60.2924.820002.93 d338 d-[199]
Morphiceptin79.4>1000-3184800-[159,200]
PL0172.94200-211250-[200]
PL0325.5>10,000-291510-[200]
H-Tyr-Pro-D-1-Nal-Pro-NH21.9>1000-9.5735.4-[160
H-Dmt-DAla-(R)-β3-1-Nal-Pro-NH20.3423817507.94inactiveinactive[108]
H-Dmt-DAla-(R)-β2-1-Nal-Pro-NH20.051.0411.23.5589.1224[108]
Met-ENK9.50.91444019019-[201,202]
Leu-ENK9.432.53821035.61.73550[154,197]
DADLE13.82.0616,0008.90.73134[197]
DPDPE7132.72>15,00030004.14>10,000[197]
DPLPE6592.80>15,00023502.77>10,000[197]
[Phe(p-F)4]-DPLPE16000.43-7400.016-[203]
JOM-1351.50.74-4604.2 [204]
DSLET313.8-3600.58-[205]
DLT-121400.60-28900.36-[173]
DLT-216800.73-31800.67-[173]
[(2S,3R)-β-iPrPhe3]-DLT-I63,0002.14-32,5001.23-[170]
c[DCys2, Pen5]-DLT-1 37602.2-11000.25-[173]
c[DCys2, Cys5]-DLT-1 5.150.87-2.980.23-[173]
[DAla2,Ac6c4]-DLT2.450.045-5580.52-[172]
β-END1.01.0523.52.1-[206,207]
biphalin1.42.6-8.827-[208]
DYN A5.042.540.232.522.512.6[209]
DYN A-(1-13)- NH21.294.070.151.77863.1[209]
DYN A-(1-11)-NH21.086.990.0777.51040.376[209,210]
[DAla3]-DYN A-(1-11)-NH2674070.36--2.38[210]
E-20784.5127.21.910.37.42.6[185]
CR665405020,3000.24--0.03 c[190]
CR845--0.32--0.16 c[211]
CTOP4.35600-426 e--[161]
CTAP2.15310-75.8 e--[161]
TIPP17201.22-inactive4.80 e-[100]
TIPP-NH278.83.00-170018.0 e-[100]
DIPP-NH21.190.118-18.20.209 e-[100]
DIPP-NH2[ψ] 0.9430.447-7.710.537 e-[100]
Dmt-Tic13601.84-inactive6.55 e-[100]
Dmt-Tic-NH-(CH2)3-Ph0.3860.0871-1021.69 e-[100]
[Pro3]-DYN A-(1-11)- NH2570088002.7--244 e[210]
N,N-diallyl-[DPro10]-DYN A-(1-11)31.71493.60228 e-97 e[42]
dynantin13535420.7925 e3220 e0.632 e[196]
Arodyn1740583010.0---[212]
Zyklophin5880>10,00030.3--84 (KB) c[213]
a The variability in the binding and functional data between various laboratories may be attributed to differences in the synaptosomal preparations, the concentration, type of radioligand, and specific activity used, and the method of reporting binding. Refer to the citations for the details. b Ca Assay, EC50; c cAMP Assay, EC50; d GTP assay, EC50; e Antagonist mode, Ke; GPI: guinea pig ilem; MVD: mouse vas deferens; RVD: rat vas deferens.

3.2.3. Tools for Mimetics

Peptide Backbone Modifications

Modification of the amide bonds that do not interact with the target is one of the most widely used strategies to optimize the metabolic stability of peptides in vivo, despite the risk of interrupting the conformational profile of the peptides. It is critical to retain the spatial disposition of side chains that are essential for the interaction with the receptors. Peptoids, poly Nα-substituted glycines are α-peptidomimetics, in which the side chains are attached to the amide nitrogen atom rather than to the α-carbon atom, thus causing a loss of the intra-backbone hydrogen bonding and chirality and a formal shift of the side chain position, but improving the resistance to enzymatic hydrolysis and cell permeability [214,215] (Figure 6). These are well suited for the high-throughput library synthesis, and CHIR4531, a peptoid trimer, was the first active opioid peptoid with high MOR affinity (Ki = 6 nM) discovered from a library of 5000 peptoids [214]. In most cases, however, peptoids or peptoid-peptide hybrid modifications were not successful in the well-known opioid peptides due to resulting conformational changes [216].
β-peptides, poly β-amino acids, possess greater structural diversity and high resistance to the enzymatic degradation due to the additional carbon atom in each amino acid residue, and their biological properties differ from those of parent peptides. Substitution of each amino acid with respective β-amino acids in Leu-ENK, DER, and DLT resulted in a loss of the affinities at MOR and DOR [217]. Other backbone modifications are retro (reversal of backbone), inverso (inversion of chirality of amino acid), and retro-inverso modifications applied to opioid peptides as a tool for structural alteration to improve metabolic stability [218]. These modifications of opioid peptides are not considered to be promising due to the interruption of the spatial disposition of key side chains, despite a few successful cases [218,219]. Overall, studies on backbone modifications of opioid peptides indicated that this approach is not efficient due to resulting structural variability.

Cyclization

Cyclization has been a major tool to develop conformationally constrained topographical structures for opioid receptors. Limiting the number of conformations enhances the ability to position the backbone and/or side chains properly for the receptors and improves receptor selectivity and specificity, bioavailability, and metabolic stability [220,221]. Cyclization decreases the number of hydrogen bonding and hydrodynamic radius in a solution but increases lipophilicity, which aids the molecule to cross the BBB to reach the CNS [45]. Constraints of cyclic peptides can be attained through the formation of a covalent bond, such as a lactone, lactam, disulfide, thioether, etc., that spans from the N-terminus to C-terminus, N/C-terminus to side chain, or side chain to side chain, which is the most common [128,222,223,224,225] (Figure 6). Typical scaffolds of opioid cyclic peptides are Tyr-c[Xxx-Gly-Phe-Yyy]-OH/NH2, Tyr-c[Xxx-Gly-Phe-Yyy], Tyr-c[Xxx-Phe-Phe-Yyy]-OH/NH2, Tyr-c[Xxx-Phe-Yyy]-OH/NH2, and Tyr-c[Xxx-Phe-Yyy] modified from ENKs, EMs, DER, DLTs, and morphiceptin, where Xxx and Yyy are residues to be consumed for the cyclization through a variety of linkers. However, the cyclization of those small sizes of opioid peptides is not straightforward due to the lack of folding ability in a compact structure.
Due to the essential role of a positively charged nitrogen atom in the receptor recognition, most modifications left Tyr1 unaffected. Cyclic EM-1 analogs in which the N-terminal amino group was consumed decreased MOR affinities dramatically to the micromolar range, although there were a few exceptional cases, such as cN,C[Tyr-DPro-DTrp-Phe-Gly] showing moderate binding affinity (Table 3) [147]. Dhp-c2,5[DOrn-2-Nal-DPro-Gly] was the first cyclic analog without a positively charged nitrogen atom yielding a weak MOR/DOR antagonist [41]. SAR studies on cyclic EM-2 analogs indicated optimal ring size for MOR affinity and selectivity to be 14 while ENK preferred to be 18 [226,227]. I.c.v. and i.p. administrations of cyclic EM-2 analogs (Tyr-c2,5[DLys-Phe(F/2,4-difluoro)-Tyr-Gly]) exhibited potent analgesic effects indicating the potential brain uptake, and the structural analyses using NMR spectroscopy suggested that a trans conformation of the backbone between positions 2 and 3 is critical for affinity, selectivity, and functional activity at the MOR. An extended structure of EM-1, ZH853 (Tyr-c[DLys-Trp-Phe-Glu]-Gly-NH2) is a recent success showing high potential as a therapeutic agent for various pain states, including neuropathic pain [28,30,228]. Another EM analog, Tyr-c2,5(-SCH2S-)[DCys-Phe-2-Nal-Cys]-NH2 showed a MOR/KOR agonist and DOR partial agonist/antagonist profile with a similar affinity at MOR and DOR [229]. Enhanced KOR activity was observed in a series of cyclic EM pentapeptides, Dmt-c[DLys-Xxx-Yyy-Asp]-NH2, in which Xxx and Yyy are Phe or NMePhe [195].
Interestingly, the substitution of a Tyr with a Dmt in cyclic peptides did not affect affinities, unlike linear peptides, but still increased the metabolic stability. Oral administration of a cyclic morphiceptin analog, Dmt-c[DLys-Phe-DPro-Asp]-NH2 (P-317), inhibited GI motility and produced a potent and long-lasting MOR-medicated antinociceptive effect due to enhanced enzymatic resistance and anti-diarrhoeal activity [230,231]. The small cyclic tetrapeptide, Tyr-c2,4[DLys-Phe-Ala], has the picomolar range of potency for MOR (0.11 nM) and DOR (0.54 nM) in the GPI and MVD assays, and its modification with Asp4 significantly reduced both activities, indicating the incompatible effect of the negative charge on both receptors [232]. Interestingly, Tyr-c[DLys-Phe-Asp]-NH2, in which the side chain of Asp4 was consumed for the cyclization, was a highly potent and selective MOR agonist [227].
Driven by the success of DPDPE, many modifications have been performed mainly to replace a disulfide bond susceptible to enzymatic degradation and nucleophilic, with a stable bond, such as a rigid olefin bond (Figure 6). Both cis and trans olefin analogs, Tyr-c2,5(-trans/cisCH=CH-)[DAla-Gly-Phe-DAla]-OH, were potent MOR/DOR agonists with a slight selectivity for DOR [224]. A methylamine-bridged ENK analog, Tyr-c[(NβCH3)-D-A2pr-Gly-Phe-NHCH2CH2-], was a potent MOR/DOR agonist showing substantial antinociceptive effects (ED50 = 27 ng) [233]. Lanthionine (thioether) bridge-substituted analogs, Tyr-c2,5(-S-)[DAla-Gly-Phe-D/LAla]-OH, were also potent MOR/DOR agonists with a subnanomolar analgesic potency (ED50 = 0.0015 and 0.0018 nM, respectively) [223]. A series of cyclic ENKs with a urea bridge showed mixed MOR/DOR agonist activity. Tyr-c2,5(-NHCONH-)[DLys-Gly-Phe-Dap]-NH2 was a very potent MOR/DOR agonist with the subnanomolar range of EC50 (0.21 and 0.65 nM in the GPI and MVD, respectively) and ED50 (0.0792 and 0.3526 nM in the hot-plate and tale-flick tests, respectively, i.c.v. in rats) [226,234]. Cyclic biphalin analogs through various linkers such as xylene were highly potent MOR/DOR agonists with prolonged analgesic effects after systemic administration [235,236]. Interestingly, a C-terminal-modified cyclic biphalin analog, c2,2′(Dmt-DCys-Gly-Phe)2-cystamine, was a very potent MOR/DOR/KOR agonist with the sub-nanomolar affinities and the same high efficacies with DAMGO, DADLE, and salvinorin A at MOR, DOR, and KOR, respectively [237].
DYN A was cyclized at various positions due to a relatively longer peptide chain, and the positioning and stereochemistry of the cyclization impacted KOR affinity and selectivity. cN,5[Trp3,Trp4,Glu5]-DYN A (1–11)-NH2 was the first cyclic KOR antagonist lacking the N-terminal amino group but retained the nanomolar affinity [238]. Two isomers of c2,5(-CH=CH-)[Ala2,Ala5]Dyn A (1–11)-NH2 exhibited a high KOR affinity but similar selectivity to DYN A-(1–11)-NH2 [239]. Incorporation of DAla at position 3 of c2,5[DAsp2,Dap5]-Dyn A (1–11)-NH2 resulted in the increase of KOR selectivity over MOR and DOR [195]. Two isomers of c5,8(-CH=CH-)[Ala5,Ala8]Dyn A (1–11)-NH2 having the address region in a cycle significantly increased KOR selectivity compared to Dyn A (1–11)-NH2 [239]. The cyclization of a partial KOR agonist, [BzlTyr1]-DYN A-(1-11)-NH2, at the same positions 5 and 8 through a lactam ring resulted in the development of a highly selective KOR antagonist, zyklophin, capable of crossing the BBB [18,194,213].

N- and/or C-Terminal Modifications: Amide (-NHMe, -NHMe, -NMe2), Ester (-OMe, -OEt, -OBu), Hydroxy (-OH), or Hydrazide (-NH-NH2)

In addition to the optimization of enzymatic stability, modifications of the N- and C-terminus of opioid peptides affect functional activity and receptor selectivity, respectively, based on the message–address concept [129,132]. Due to the essential role, N-terminal modifications were limited to the trials converting functional activity, from agonism to antagonism, while C-terminal modifications were performed frequently to optimize the selectivity and physicochemical property or conjugate with other pharmacophoric structure [39,43,93,245]. Transformations of the C-terminal carboxylate into a carboxamide or a carboxylate ester resulted in the reduction of DOR selectivity in ENK and DPDPE due to an enhanced MOR activity [57,167]. The attachment of alkyl urea moiety (-NHCH2CH2NH(C=O)NH2) at the C-terminus of ENK retained the same activity at MOR and DOR [246]. Various C-terminal amidations of TIPP transformed the selective DOR antagonist into a mixed MOR agonist/DOR antagonist [175]. Simple C-terminal modifications of EMs to an ester or a hydrazide showed significant analgesic effects after central (i.c.v.) and peripheral (s.c.) administrations with the reduced tolerance, cardiovascular, and constipation [60]. There was no explanation of how physicochemical properties modulated side effects. A recent study showed C-terminal amidated ENK reduced β-arrestin recruitment efficacies at MOR and KOR [247]. The role of the C-terminus is still not clear, due to the inconsistent results [60]. Further modification of Phe4 to -NH-X- (aromatic, heteroaromatic, and aliphatic groups) resulted in the loss of MOR activity [23].
Deletion and modification of the N-terminal amino group alter opioid agonist activities to antagonist activities as discussed in Section 3.1.2 [39,43]. Interestingly, incorporation of lipoamino acids to the N-terminus of EM retained MOR agonist activity and afforded a strong antinociceptive effect without causing constipation after oral administration in CCI rats [71]. Cationization of the N-terminus of morphiceptin and EM through guanidinylation (NH2-C(=NH2)-NH-) increased the brain uptake and metabolic stability, although decreased affinity at MOR occurred [71]. N-guanidinylated cyclic ureido DLT analogs exhibited the balanced nanomolar affinities at the MOR and DOR, and high enzymatic stability that mediates clinically relevant analgesia [248]. A recent study showed that modifications of the N- and C-terminus with a guanidine and a tetrazole could improve metabolic stability (t1/2 > 6 h) and lipophilicity (clogD = 0.44) without affecting their activities at MOR and DOR [52]. ADAMB is a good example to show long-lasting activity and strong oral antinociceptive activity after the substitutions at the N-terminus and position 4 [85]. Recent studies suggested that C-terminal modifications can regulate β-arrestin 2 recruitment and thus reduce MOR related side effects [123,124].

Peptide Linking

As shown in biphalin, linking two opioid peptides increases potency, efficacy, and bioavailability compared to their corresponding monomeric counterparts in the cases of receptor dimerization or cross-linking of neighboring receptors by triggering receptor-receptor interactions [94,249]. Dual agonist activity at MOR and DOR results in a synergistic antinociceptive effect that is advantageous in reducing dosage, thereby attenuating related side effects. Based on this advantage, peptide linking has been widely used by connecting two address regions through various linkers while retaining two message regions [94,235,250]. Biphalin showed a similar trend on modifications to ENK, and substitutions of Phe4 with Phe(X) and Nal residues affected its activity positively together with the linker variation. It was shown that analogs with a hydrophilic linker such as butanediol increased affinity at KOR whereas a piperazine linker produced a balanced high affinity for MOR and DOR [94]. [Dmt1]DALDA-NH-CH2-CH2-NH-TICP(ψ), in which two structural moieties were linked through an ethylenediamine, was a hetero-bivalent ligand displaying MOR agonist/DOR antagonist profile with the nanomolar binding affinities at both receptors [251]. Bitopic ligands, a new class of ligands that simultaneously target orthosteric and allosteric sites, are an extension of the bivalent ligand concept, and no bitopic opioid ligand has been identified yet. However, considering positive allosteric interactions between MOR and DOR, bivalent ligands may act as a bitopic ligand [252].

Local Backbone Modifications

Peptide backbone is defined by the φ, ψ, ω, and χ torsional angles [128]. Altering a peptide bond using various non-natural amino acids or isosteres has been widely used to amplify conformational constraints of the dihedral angles and thereby to reduce the rotational freedom of those side chains directly involved in the biological interaction (Figure 6).
Most modifications of ENK using alkylurea, trifluoroethylamine, oxymethylene, olefin, and bismethylene isosteres were not successful except in a few cases [57,253,254,255]. Among them, olefine bonds from α,β-unsaturated α-amino acids conserve the planarity of the peptide bond but no H-bonding because of the deletion of an oxygen atom, an acceptor. A fluoroalkene substituted-ENK, [Tyr1-ψ[(Z)CF=CH]-Gly2]-Leu-ENK, was a successful case increasing the stability as an orally active CNS-distributed peptide probe [166,256]. Its modifications at the C-terminus reduced β-arrestin recruitment through the DOR and MOR while retaining affinity and cAMP potency [124]. Another DOR analogs, [Gly2-ψ[(Z)CF=CH]-Phe3]-Leu-ENK, [Gly2-ψ[C(=O]O]-Phe3]-Leu-ENK, [Gly2-ψ[C(=S)NH]-Phe3]-Leu-ENK, retained the agonist activity, indicating the amide bonds can be replaced by a fluoroalkene, an ester, a thioamide possessing high dipole moment and rotational barrier without a substantial loss of activity [165]. Positional scanning of the Leu-ENK with an ester bond and N-methylation demonstrated that the fourth amide bond is replaceable [164]. The SAR results of two analogs, [Gly4-ψ[C(=O]O]-Leu5]-ENK and [NMeLeu5]-ENK indicated that an H-bonding acceptor is critical [164]. Trials to enhance the systemic analgesic effect of [Dmt1]-DPDPE using oxymethylene (-C-O-), trans-double bond (-C=C-), or bismethylene (-C-C-) isosteres were not successful [57]. β-sulfonamide analog of EM-2 in which Pro2 was replaced with 3-pyrrolidinemethanesulphonic acid (βPrs) retained a high affinity and selectivity at MOR (Ki = 19 nM) [140]. Recently, aza-pipecolyl (azaPip) substituted EM and morphiceptin analogs were developed to study the conformational requirements, whether cis-amide geometry and β-turn conformation are critical for the receptor interactions [257]. Its application in H-Dmt-azaPip-DPhe-Pro-NH2 improved potency and affinity for MOR and DOR, indicating the requisite for natural Tyr-Pro cis form for the receptor binding [257].
As a local variation of parent amide backbone, β-amino acids, γ-amino acids, β-dipeptidomimetics, and the other conformationally constrained non-natural amino acids (Figure 7) have been simply applied to opioid peptides [51,258]. Insertion of one (β-amino acids) or two additional carbon atoms (γ-amino acids) into a peptide backbone changed the conformational structure, afforded the structural diversity by increasing diastereomers, and enhanced metabolic stability for prolonged in vivo effect [258]. Incorporation of cyclic amino acids further produced successful peptidomimetics for degradation resistance and restricted conformation. For these reasons, many studies have focused on incorporating a variety of β-, γ-, or cyclic β-amino acids into opioid peptides.
Position 2 of EM has been modified successfully without damaging the affinity but improving the enzymatic hydrolysis resistance despite the reduction of MOR selectivity over DOR [111,136,142,240,259]. Likewise, many studies showed that the substitutions of various β-amino acids including acyclic β-amino acids into EMs were well tolerated at position 2, unlike the other positions [136,141,143]. Substitution of a Pro residue with β2-(R)-Pro or β3-(S)-Pro increased or retained affinity at the MOR and exhibited good resistance to the proteolytic enzyme and effective peripheral antinociception [111,260]. Dmt-(1S,2R)Achc-Phe-Phe(p-F)-NH2, in which cis-2-aminocyclohexane carboxylic acid (Achc) was substituted at position 2 was a potent MOR agonist (EC50 = 51 nM) with high efficacy (176% compared to DAMGO) and enhanced proteolytic stability (t1/2 > 20 h in rat brain membrane) [136]. Incorporation of β2-, β3-hPhe, or β3-hTic at positions 3 and 4 resulted in the decrease of MOR selectivity by increasing DOR activity while retaining MOR activity, and NMR and molecular modeling studies indicated their predominant turn structures [261,262]. A series of more constrained β-amino acids, 2-methylene-3-aminopropanoic acids (MAP), were incorporated into positions 3 and 4 of EM [263]. H-Tyr-Pro-Trp-(2-Furyl)Map-NH2 was a very potent MOR agonist with the picomolar range of binding affinity (Ki = 0.22 nM), 430-fold increased agonist activity (EC50 = 0.0334 nM, Emax = 97%) in the cAMP test, high metabolic stability, and enhanced antinociceptive activity in the mice tail-flick test [263]. β-dipeptidomimetics (-β3hPhe-Tbac-, -β3hPhe-Tia-, -β3hPhe-β2hPhe-, -β3hPhe-β3hPhe-, -β2hPhe-β2hPhe-) replacing the native backbone of Phe3-Phe4 residue in EM-2 retained MOR affinity due to the prevalent extended conformation [262]. Substitution of a Phe residue with chiral α-hydroxy-β-phenylalanine (AHPBA, βAtc) in EMs and [Dmt1,Tic2]-EMs retained the biological activities of the parent peptides and improved half-life slightly in most cases [264]. Among those, Dmt-Tic-(2R,3S)AHPBA-Phe-NH2 significantly increased the half-life (>2 h) and retained a strong antinociception (ED50 = 2.0 nM, i.c.v.) without causing a motor effect and acute tolerance due to the potential role of DOR antagonist involved.
In morphiceptin analogs that are similar to the structure of EMs, incorporation of cis-2-aminocyclopentane carboxylic acid (cis-2-Acpc) into position 2 retained the activity and selectivity of parent peptides for MOR due to enough separation of Tyr1 and Phe3[129,141,157]. Contrarily, the incorporation of cis-2-Achc into DER lost MOR activity, indicating its unfavorable conformation for the receptor and distinct receptor interaction from morphiceptin [265]. In multiply modified analogs, the substitution of β2- or β3-amino acids affected the selectivity depending on their configurations and resulted in the development of an agonist, H-Dmt-DAla-β2-(R)-1-Nal-Pro-NH2, for MOR, DOR, and KOR (Ki = 0.05, 1.04, and 11.2 nM, respectively) with high metabolic stability (t1/2 > 50 h) [108]. Incorporations of β3-homo amino acids into the message region of DLT decreased binding affinity at DOR except positions 2 and 3 [217].
Many studies have focused on restricting conformational structures to an active one using constrained amino acid analogs [266]. Tic, 7-hydroxy-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid (Htc), and β-carboline-3-carboxylic acid (Tcc) are a constrained cyclic analog of Phe, Tyr, and Trp, respectively, which are key structural elements of most opioid peptides. and Tic has been successfully applied to the design of opioid peptidomimetics (Figure 8). Dmt-Tic is the shortest peptide pharmacophore for DOR antagonism and is being widely used as a template due to the remarkable alterations in selectivity and activity [180,264,267,268]. 1,2,3,4-tetrahydroquinoline (Thq) derivative coupled with a Dmt residue to resemble JOM-13 structure showed potent MOR agonist activity [269]. Further modifications at C-6 and C-8 positions of Thq with various aromatic, cycloalkyl, heterocyclic, and alkyl groups transferred its activity to a mixed opioid activity with optimized pharmacological properties reducing tolerance and dependence as well as improving the BBB permeability [91,270,271]. The modification with 4-hydroxy 2-methylindanyl group at C-6 significantly increased efficacy at KOR (Ki = 0.77 nM, EC50 = 25 nM, 92% stimulation) and MOR (Ki = 0.18 nM, EC50 = 4.9 nM, 66% stimulation) resulting in a MOR/KOR agonist [272].
Replacement of Pro2 of EMs with a more constrained cyclic amino acid such as (S)-azetidine-2 carboxylic acid (Aze) and 3,4-dehydro-(S)-proline (Δ3Pro) improved the potency but resulted in the loss of selectivity for MOR (Figure 8) [143,273]. In nature, diverse structures of α,β-dehydroamino acids exist, and dehydrophenylalanine is the most widely used one. The substitution at position 4 of EMs retained the same high potency at MOR as β-MePhe substituted analog [273,274]. A 4-imdidazolidone moiety that contains a methylene bridge between two amide bonds was well tolerated for positions 4 and 5 of ENKs showing a similar range of affinities at MOR and DOR with a slight MOR selectivity [275]. Replacement of Gly2-Gly3-Phe4 moiety with an aminobutyl-substituted pyrazinone ring diminished MOR and DOR activities, whereas 6-(Dmt-NHBu)- and bis-(Dmt-NHBu)-substituted pyrazinone rings showed very potent MOR agonist and weak DOR antagonist properties [58,276,277]. Recently, triazole rings are receiving attention thanks to the simple synthetic procedure, click chemistry, and the ability to adjust flexibility on all the rotatable bonds [278]. However, the positional scanning of ENKs with a 1,4-disubstituted triazole, 1,5-disubstituted tetrazole, 4-imidazolidinone, and cyclopropane ring was not successful, with an overall loss of activity in most cases. It was suggested that natural amide bonds are critical for receptor interactions, particularly through hydrogen bondings between oxygen and hydrogen atoms [275,278,279,280]. Another positional scanning of DER-tetrapeptide with piperazin-2-one moieties, N,N′-ethylene-bridged dipeptides showed slightly different SAR depending on the chirality and backbone [281].
Bicyclic conformational constraints are the most common way of building β-turn mimics, as established by the β-turn-dipeptide. Various heterocyclic scaffolds have been applied to induce the β-turn structure of opioid peptides. Atypical MOR agonist without a cationic amino group, in which Ri, Ri+2, and Ri+3 are 4-hydroxyphenethyl, phenethyl, phenethyl, respectively, was identified through combinational libraries of β-turn peptidomimetics of EM [128,139]. Another β-turn opioid peptidomimetic, thiazolidinone bicyclo [2,3]-Leu-ENK, turned out to be a weak MOR agonist/DOR antagonist [282]. Azepinone scaffold has been utilized for the formation of constrained dipeptidic moieties such as Aba, 8-hydroxy-4-amino-1,2,4,5-tetrahydro-2-benzazepin-3-one (Hba), and 4-mino-1,2,4,5-tetrahydro-indolo [2,3-c]azepin-3-one (Aia). Replacement of Tic-Gly in Dmt-Tic-Gly analogs with dipeptidic moieties, Aba-Gly and DAia-Gly, shifted affinity and selectivity to afford a potent MOR agonist and DOR antagonist, respectively [283,284,285,286]. The same replacement effect was observed in DER tetrapeptide when Phe3-Gly4 was substituted by Aba-Gly [287]. Recent application of a bicyclic scaffold rendered atypical cyclic EM peptidomimetic [155]. H-Dmt-NMe-DAla-[(S)-Aba-Gly]-NH2 behaves as a mixed MOR/DOR agonist with high potency (EC50 = 1.7 pM, 16 pM, respectively). Functional electrophysiological in vitro screening using primary cortical and spinal cord networks indicated that this was the most potent, showing potential for neuropathic pain.

Non-Natural Amino Acid Replacements without the Alteration of a Peptide Bond

The simplest approach to accomplish conformational changes is to replace one or several positions with non-natural amino acids, such as D-amino acids, α-, β-, or γ-substituted amino acids, and Nα-alkylated amino acids [162]. Introduction of a D-amino acid alters the spatial orientation of a side chain and results in the change of the conformation of the peptide backbone. The configuration-dependent differential effects on the opioid receptor selectivity and intrinsic activity have been well evidenced [162,175]. Nα-substituted amino acids also influence the conformational freedom of both the backbone and side chain of an adjacent residue and increase the peptide’s lipophilicity with the reduced number of hydrogen bondings, resulting in the enhancement of the BBB permeability. The simple introduction of a NMePhe residue at position 4 of ENK retained a good DOR activity along with the increase of lipophilicity and stability [164]. [DCys2, NMe-DCys5]ENK turned out to be the most potent cyclic MOR/DOR agonist (IC50 = 1.3, 0.02 nM in GPI and MVD, respectively), possessing a good analgesic effect [288]. FK33-824 (Tyr-DAla-NMePhe-Met(O)-ol) and Metkephamid (Tyr-DAla-Gly-NMeMet-NH2) were representative analogs showing central analgesic effects through systemic administrations, which indicate the increased BBB permeability [98].
Cα-substituted amino acids have been used to limit the number of conformations by reducing rotational freedom around a peptide bond. The 2-aminoisobutyric acid (Aib) is a common residue to elicit a turn-structure due to the extra methyl group-induced steric hindrance [128]. Incorporation of an Aib residue into ENK (position 3) and DLT (positions 2–4) retained a similar activity at DOR without detrimental effects [167,289,290]. ICI174864 is an example showing high DOR selectivity with the Aib substitutions [174]. Substitutions of Cα,α-dialkyl cyclic amino acids in DLT showed that the D-chirality of the 2nd position is not critical for the DOR selectivity, and the aromatic ring of the 3rd position is replaceable with a hydrophobic ring [172]. The replacement of Phe3 of DLT with azidomethyl- or piperidinylmethyl-substituted Phe residues (S-form) retained the same range of affinity and selectivity for DOR, whereas other derivatives lost affinities [134].
Cβ-substituted amino acids restrict the Cα-Cβ bond without affecting backbone angles and select one rotamer of aromatic amino acids (Figure 9) [128]. For this reason, Cβ-substituted amino acids have been frequently used to produce a constrained structure optimized for biological activity and physicochemical property [291]. β-MePhe, β-iPrPhe, β-methyl-2,6-dimethylphenylalanine, and Tmt are the most widely used in opioid peptidomimetics [292,293]. The favorable side chain rotamer for DOR was the trans form (2S, 3R), and [(2S, 3R)-β-Tmt1]DPDPE and [(2S, 3R)-β-iPrPhe3]DLT increased DOR selectivity with retained affinity [170,293]. D,L-amino-β-mercapto-β,β-pentamethylenepropionic acid (Apmp) has a cyclohexyl ring in place of the geminal dimethyl group of Pen, and its substitution at position 2 in DPDPE resulted in the loss of DOR affinity and selectivity due to the steric hindrance and hydrophobic property while the substitution at position 5 was tolerated [294].
Sterically constrained aromatic amino acids have been substituted for Phe and Tyr residues in opioid peptides successfully: 2′,6′-dimethylphenylalanine (Dmp), Emp, Tmp, 2′,6′-dimethyl tyrosine Dmt, and Tmt [97]. Among these, Dmt is the most successfully applied in many opioid peptides to afford [100,267]. The β-methyl group on the side chain of the amino acid basically “locks” the χ1 and χ2 conformation of the molecule [57,128]. The substitution in DPDPE and DPDPE-OMe resulted in the increase of potency at DOR and MOR but the decrease of DOR selectivity because of the enhanced MOR activity. The substituted analogs showed a significant antinociceptive activity following systemic administration [55,57]. Another trial to utilize the SAR at the MOR resulted in the discovery of a molecule with a highly balanced MOR agonist/DOR antagonist property, DIPP[ψ]-NH2 [100,176,177]. The substitution in DALDA resulted in the discovery of [Dmt1]-DALDA showing the drug-like property, long duration of antinociceptive effects in vivo after i.th. or s.c. injections, high resistance to enzymatic degradation, and slow clearance along with oral delivery potential [61,295,296,297]. Studies on the binding mode indicated that the two methyl groups on the aromatic ring of Dmt1 in opioid peptides participated in additional lipophilic contacts with MOR residue as well as steric influence [298]. Other constrained aromatic amino acids, Dmp, Tmp, and Emp, were introduced to the EM scaffold to lead distinct bifunctional activities such as MOR/DOR agonists or MOR agonists/DOR antagonists [97]. The alteration of EM-2 activity by Dmt1 and alkylated Phe3 residues produced a dual MOR/DOR agonist activity in Dmt-Pro-Emp-Phe-NH2 and a mixed MOR agonist/DOR antagonist in Dmt-Pro-Tmp-Phe-NH2 [97]. Substitution of a cyclic EM, Dmt-c[Lys-Phe-Phe-Asp]-NH2, with 2′,3′, or 4′-MePhe4 improved the activity slightly at DOR and dramatically at MOR, and exhibited remarkable antinociceptive effects after i.c.v. administration through potent MOR and KOR agonist activities [107].
Naphtylalanine (Nal) analogs contain a more extended aromatic ring system, like a Trp residue, and possess higher π-electron density. These bulkier residues have been successfully substituted for a Phe residue in ENKs, cyclic ENK analogs, and other opioid peptides to afford potent mixed-efficacy for opioid receptors [229,299,300]. Tyr-c[DOrn-2-Nal-DPro-Gly] was the first potent analog with mixed MOR agonist/DOR antagonist property after less potent TIPP-NH2 [300]. Structural analysis indicated that the enhanced MOR activities in ENK analogs came from the relatively extended structure [299,301]. A variety of p-substituted phenylalanine analogs were also introduced to multiple opioid peptides including ENKs, DPDPE, and DSLET as a surrogate of Tyr1, and (S)-4-carboxamidophenylalanine (Cpa) and (S)-4-carboxamido-2,6-dimethylphenylalanine (Cdp) exhibited the equivalent affinity for DOR compared to parent compounds [302]. The Cdp with 2.6-dimethyl group enhanced affinity for the MOR more than Cpa analog.
In most opioid peptides, the substitution of halogenated Phe residues, Phe(p-X), produced an enhanced opioid activity unlike other derivatives such as Phe(p-NH2), Phe(p-NCS), Phe(p-NHCOCH2Br) [93,303,304,305]. However, the substitution of 4-bis(2-chloroethyl)amino-L-phenylalanine (Mel) in Met-ENK-Arg-Phe was not tolerated at MOR and DOR, whereas C-terminal substitution was well tolerated at KOR [306]. The fifth residue in the address region of Leu-ENK has been modified by various analogs, including branched alkyl, cycloalkyl, or Nα-Cα cyclized moiety to result in a small effect on biological activity due to the lack of role in the receptor interaction (Figure 10). Interestingly, aza-β-homoleucine or cycloleucine substituted ENK turned out to be biased with lower β-arrestin recruitment, suggesting the potential role of the 5th position in producing G-protein biased signaling [123].
Besides the modification of a backbone and the replacement of amino acid residue, a combinatorial library has also been used to identify new privileged peptidomimetic structures for opioid receptors. Acetalins (AC-RFMWMT/R/K-NH2) was the first potent MOR antagonists identified through the peptide library, and Tyr-DNve-Gly-Nal-NH2 (Ki = 0.4 nM), Trp-DTyr-Aba-Arg-NH2 (Ki = 7 nM), and DPhe-DPhe-DNle-DArg-NH2 (Ki = 1 nM) were the most potent MOR, DOR, and KOR agonists, respectively [192,307,308]. First peptidomimetic library using nonnatural oligo-N-substituted glycines identified a MOR ligand CHIR4531 (Ki = 6 nM) [214].

4. Summary and Prospective Aspects

Opioid analgesics such as morphine have been the most commonly used for severe pain despite serious side effects with long-term administration, and there are still unmet needs to solve the problems. Because the efforts to develop a new molecular entity for a therapeutic purpose have not been successful, and there are currently a decreasing number of approved drugs, it is likely important to find an alternative for the increase of productivity. In general, the lack of efficacy, toxicity, and clinical safety are the most common causes of failure in the drug discovery process. Therefore, utilization of natural opioid peptides, including EMs, DER, DLTs, ENKs, DYNs, and END, which do not possess intrinsic toxicity in the body, may provide an efficient model for novel analgesics with enhanced therapeutic advantages. In this light, opioid peptidomimetics have been studied deeply to afford applications in the therapeutic field, although the results are still insignificant: one in medical use.
Undesirable physicochemical properties and the resulting low metabolic stability and BBB permeability are the main drawbacks of opioid peptides as a drug candidate. The strategy of opioid peptidomimetic serving as a bridge between the respective characters of peptides and small molecules has solved the problems and enhanced the biological activities of natural opioid peptides as well in many cases. Mimicking natural opioid peptides is still dynamic, and there remains much room for a simplified and synthetically straightforward structure while maintaining the key structural features for the receptors. For the rational design of an opioid peptidomimetic, a full understanding of metabolic degradation pathways and SAR, and its application for pharmacokinetic and pharmacodynamic properties are significant.
Over the decades, various synthetic strategies have been developed for the preparation of non-natural amino acid derivatives or isosteres of which the substitution can modulate the intrinsic flexibility of peptide backbone. As discussed in numerous cases, opioid peptidomimetics have successfully advanced the research in opioid and pain treatment along with the advances of new technologies in structural analysis, molecular mechanics, and synthetic methodology. Systemic administrations of a few opioid peptidomimetics, including oral administration, have been shown to produce central analgesic effects through enhanced metabolic stability and BBB permeability. On the other hand, a few studies have also proven that peripherally acting opioid peptidomimetics are safe analgesic drugs with minimal side effects in the CNS. Further efforts, however, in transforming the molecules to opioid peptide-based drugs with improved safety profiles are still critical, and new formulations or routes of administration may quicken the process. Taken together, the small size of opioid peptidomimetics is likely to have a high potential of being a potent systemic and oral drug and also to provide unanticipated opportunities for the development of a novel analgesic drug that is free from opioid side effects.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Farmer, P.; Ariens, E. Speculations on the design of nonpeptidic peptidomimetics. Trends Pharmacol. Sci. 1982, 3, 362–365. [Google Scholar] [CrossRef]
  2. Wang, L.; Wang, N.; Zhang, W.; Cheng, X.; Yan, Z.; Shao, G.; Wang, X.; Wang, R.; Fu, C. Therapeutic peptides: Current applications and future directions. Sign. Trans. Targeted Therap. 2022, 7, 1–27. [Google Scholar]
  3. Eguchi, M. Recent advances in selective opioid receptor agonists and antagonists. Med. Res. Rev. 2004, 24, 182–212. [Google Scholar]
  4. Onogi, T.; Minami, M.; Katao, Y.; Nakagawa, T.; Aoki, Y.; Toya, T.; Katsumata, S.; Satoh, M. DAMGO, a μ-opioid receptor selective agonist, distinguishes between μ-and δ-opioid receptors around their first extracellular loops. FEBS Lett. 1995, 357, 93–97. [Google Scholar] [CrossRef]
  5. Johnson, K.V.-A.; Dunbar, R.I. Pain tolerance predicts human social network size. Sci. Rep. 2016, 6, 1–5. [Google Scholar]
  6. Lemaire, S.; Berube, A.; Derome, G.; Lemaire, I.; Magnan, J.; Regoli, D.; St. Pierre, S. Synthesis and biological activity of. beta.-endorphin and analogs. Additional evidence for multiple opiate receptors. J. Med. Chem. 1978, 21, 1232–1235. [Google Scholar] [CrossRef]
  7. Dickenson, A.H. Plasticity: Implications for opioid and other pharmacological interventions in specific pain states. Behav. Brain. Sci. 1997, 20, 392–403. [Google Scholar] [CrossRef]
  8. Cowan, A.; Zhu, X.; Mosberg, H.; Omnaas, J.; Porreca, F. Direct dependence studies in rats with agents selective for different types of opioid receptor. J. Pharmacol. Exp. Therap. 1988, 246, 950–955. [Google Scholar]
  9. Cheng, P.Y.; Wu, D.; Decena, J.; Soong, Y.; McCabe, S.; Szeto, H.H. Opioid-induced stimulation of fetal respiratory activity by [D-Ala2] deltorphin I. Eur. J. Pharmacol. 1993, 230, 85–88. [Google Scholar] [CrossRef]
  10. Akiyama, K.; Gee, K.W.; Mosberg, H.I.; Hruby, V.J.; Yamamura, H.I. Characterization of [3H][2-D-penicillamine, 5-D-penicillamine]-enkephalin binding to delta opiate receptors in the rat brain and neuroblastoma--glioma hybrid cell line (NG 108-15). Proc. Natl. Acad. Sci. USA 1985, 82, 2543–2547. [Google Scholar] [CrossRef]
  11. Rigter, H.; Hanna, T.J.; Messing, R.B.; Martinez, J.L., Jr.; Vasquez, B.J.; Jensen, R.A.; Veliquette, J.; McGaugh, J.L. Enkephalins interfere with acquisition of an active avoidance response. Life Sci. 1980, 26, 337–345. [Google Scholar] [CrossRef]
  12. David, M.; Moisand, C.; Meunier, J.-C.; Morgat, J.-L.; Gacel, G.; Roques, B.P. [3H] Tyr-d-Ser-Gly-Phe-Leu-Thr: A specific probe for the δ-opiate receptor subtype in brain membranes. Eur. J. Pharmacol. 1982, 78, 385–387. [Google Scholar] [CrossRef]
  13. Handa, B.K.; Land, A.C.; Lord, J.A.; Morgan, B.A.; Rance, M.J.; Smith, C.F. Analogues of beta-LPH61-64 possessing selective agonist activity at mu-opiate receptors. Eur. J. Pharmacol. 1981, 70, 531–540. [Google Scholar] [CrossRef]
  14. Wee, S.; Koob, G.F. The role of the dynorphin–κ opioid system in the reinforcing effects of drugs of abuse. Psychopharmacology 2010, 210, 121–135. [Google Scholar] [CrossRef] [Green Version]
  15. Bruchas, M.R.; Land, B.B.; Chavkin, C. The dynorphin/kappa opioid system as a modulator of stress-induced and pro-addictive behaviors. Brain. Res. 2010, 1314, 44–55. [Google Scholar] [CrossRef]
  16. Van’t Veer, A.; Carlezon, W.A. Role of kappa-opioid receptors in stress and anxiety-related behavior. Psychopharmacology 2013, 229, 435–452. [Google Scholar] [CrossRef]
  17. Schwarzer, C. 30 years of dynorphins--new insights on their functions in neuropsychiatric diseases. Pharmacol. Ther. 2009, 123, 353–370. [Google Scholar] [CrossRef]
  18. Aldrich, J.V.; Patkar, K.A.; McLaughlin, J.P. Zyklophin, a systemically active selective kappa opioid receptor peptide antagonist with short duration of action. Proc. Natl. Acad. Sci. USA 2009, 106, 18396–18401. [Google Scholar] [CrossRef]
  19. Ronsisvalle, G.; Pasquinucci, L.; Pittalà, V.; Marrazzo, A.; Prezzavento, O.; Di Toro, R.; Falcucci, B.; Spampinato, S. Nonpeptide Analogues of Dynorphin A (1−8): Design, Synthesis, and Pharmacological Evaluation of κ-Selective Agonists. J. Med. Chem. 2000, 43, 2992–3004. [Google Scholar] [CrossRef]
  20. Hruby, V.J.; Agnes, R.S. Conformation-activity relationships of opioid peptides with selective activities at opioid receptors. Biopolymers 1999, 51, 391–410. [Google Scholar] [CrossRef]
  21. Zadina, J.E.; Hackler, L.; Ge, L.J.; Kastin, A.J. A potent and selective endogenous agonist for the mu-opiate receptor. Nature 1997, 386, 499–502. [Google Scholar] [CrossRef]
  22. Zadina, J.E.; MARTIN-SCHILD, S.; Gerall, A.A.; Kastin, A.J.; Hackler, L.; GE, L.J.; Zhang, X. Endomorphins: Novel endogenous μ-opiate receptor agonists in regions of high μ-opiate receptor density. Ann. N. Y. Acad. Sci. 1999, 897, 136–144. [Google Scholar] [CrossRef]
  23. Liu, W.X.; Wang, R. Endomorphins: Potential roles and therapeutic indications in the development of opioid peptide analgesic drugs. Med. Res. Rev. 2012, 32, 536–580. [Google Scholar] [CrossRef]
  24. Wilson, A.M.; Soignier, R.D.; Zadina, J.E.; Kastin, A.J.; Nores, W.L.; Olson, R.D.; Olson, G.A. Dissociation of analgesic and rewarding effects of endomorphin-1 in rats. Peptides 2000, 21, 1871–1874. [Google Scholar] [CrossRef]
  25. Czapla, M.A.; Gozal, D.; Alea, O.A.; Beckerman, R.C.; Zadina, J.E. Differential cardiorespiratory effects of endomorphin 1, endomorphin 2, DAMGO, and morphine. Am. J. Respir. Crit. Care Med. 2000, 162, 994–999. [Google Scholar] [CrossRef]
  26. Janecka, A.; Staniszewska, R.; Gach, K.; Fichna, J. Enzymatic degradation of endomorphins. Peptides 2008, 29, 2066–2073. [Google Scholar] [CrossRef]
  27. De Marco, R.; Janecka, A. Strategies to improve bioavailability and in vivo efficacy of the endogenous opioid peptides endomorphin-1 and endomorphin-2. Curr. Top. Med. Chem. 2016, 16, 141–155. [Google Scholar] [CrossRef]
  28. Feehan, A.K.; Morgenweck, J.; Zhang, X.; Amgott-Kwan, A.T.; Zadina, J.E. Novel Endomorphin Analogs Are More Potent and Longer-Lasting Analgesics in Neuropathic, Inflammatory, Postoperative, and Visceral Pain Relative to Morphine. J. Pain 2017, 18, 1526–1541. [Google Scholar] [CrossRef]
  29. Zhang, Y.-Z.; Wang, M.-M.; Wang, S.-Y.; Wang, X.-F.; Yang, W.-J.; Zhao, Y.-N.; Han, F.-T.; Zhang, Y.; Gu, N.; Wang, C.-L. Novel Cyclic Endomorphin Analogues with Multiple Modifications and Oligoarginine Vector Exhibit Potent Antinociception with Reduced Opioid-like Side Effects. J. Med. Chem. 2021, 64, 16801–16819. [Google Scholar] [CrossRef]
  30. Nilges, M.R.; Laurent, M.; Cable, C.; Arens, L.; Vafiades, J.; Zadina, J.E. Discriminative Stimulus and Low Abuse Liability Effects of Novel Endomorphin Analogs Suggest a Potential Treatment Indication for Opioid Use Disorder. J. Pharmacol. Exp. Ther. 2019, 370, 369–379. [Google Scholar] [CrossRef]
  31. Lazarus, L.H.; Bryant, S.D.; Cooper, P.S.; Salvadori, S. What peptides these deltorphins be. Prog. Neurobiol. 1999, 57, 377–420. [Google Scholar] [CrossRef]
  32. Lazarus, L.; Wilson, W.; Guglietta, A.; De Castiglione, R. Dermorphin interaction with rat brain opioid receptors: Involvement of hydrophobic sites in the binding domain. Mol. Pharmacol. 1990, 37, 886–892. [Google Scholar]
  33. Schiller, P.W.; Nguyen, T.M.; Chung, N.N.; Lemieux, C. Dermorphin analogues carrying an increased positive net charge in their “message” domain display extremely high mu opioid receptor selectivity. J. Med. Chem. 1989, 32, 698–703. [Google Scholar] [CrossRef]
  34. Fiori, A.; Cardelli, P.; Negri, L.; Savi, M.R.; Strom, R.; Erspamer, V. Deltorphin transport across the blood-brain barrier. Proc. Natl. Acad. Sci. USA 1997, 94, 9469–9474. [Google Scholar] [CrossRef]
  35. Koehl, A.; Hu, H.; Maeda, S.; Zhang, Y.; Qu, Q.; Paggi, J.M.; Latorraca, N.R.; Hilger, D.; Dawson, R.; Matile, H.; et al. Structure of the µ-opioid receptor–Gi protein complex. Nature 2018, 558, 547–552. [Google Scholar] [CrossRef]
  36. Mosberg, H.I.; Hurst, R.; Hruby, V.J.; Gee, K.; Yamamura, H.I.; Galligan, J.J.; Burks, T.F. Bis-penicillamine enkephalins possess highly improved specificity toward delta opioid receptors. Proc. Natl. Acad. Sci. USA 1983, 80, 5871–5874. [Google Scholar] [CrossRef]
  37. Wang, Y.; Kuczera, K. Molecular dynamics simulations of cyclic and linear DPDPE: Influence of the disulfide bond on peptide flexibility. J. Phys. Chem. 1996, 100, 2555–2563. [Google Scholar] [CrossRef]
  38. Lung, F.D.; Meyer, J.P.; Lou, B.S.; Xiang, L.; Li, G.; Davis, P.; DeLeon, I.A.; Yamamura, H.I.; Porreca, F.; Hruby, V.J. Effects of modifications of residues in position 3 of dynorphin A(1-11)-NH2 on kappa receptor selectivity and potency. J. Med. Chem. 1996, 39, 2456–2460. [Google Scholar] [CrossRef]
  39. Li, T.; Jinsmaa, Y.; Nedachi, M.; Miyazaki, A.; Tsuda, Y.; Ambo, A.; Sasaki, Y.; Bryant, S.D.; Marczak, E.; Li, Q.; et al. Transformation of mu-opioid receptor agonists into biologically potent mu-opioid receptor antagonists. Bioorg. Med. Chem. 2007, 15, 1237–1251. [Google Scholar] [CrossRef]
  40. Fichna, J.; do-Rego, J.-C.; Kosson, P.; Costentin, J.; Janecka, A. Characterization of antinociceptive activity of novel endomorphin-2 and morphiceptin analogs modified in the third position. Biochem. Pharmacol. 2005, 69, 179–185. [Google Scholar] [CrossRef]
  41. Schiller, P.W.; Berezowska, I.; Nguyen, T.M.-D.; Schmidt, R.; Lemieux, C.; Chung, N.N.; Falcone-Hindley, M.L.; Yao, W.; Liu, J.; Iwama, S. Novel ligands lacking a positive charge for the δ-and μ-opioid receptors. J. Med. Chem. 2000, 43, 551–559. [Google Scholar] [CrossRef] [PubMed]
  42. Gairin, J.E.; Mazarguil, H.; Alvinerie, P.; Botanch, C.; Cros, J.; Meunier, J.C. N,N-diallyl-tyrosyl substitution confers antagonist properties on the kappa-selective opioid peptide [D-Pro10]dynorphin A(1-11). Br. J. Pharmacol. 1988, 95, 1023–1030. [Google Scholar] [CrossRef] [PubMed]
  43. Schiller, P.W.; Weltrowska, G.; Nguyen, T.M.-D.; Lemieux, C.; Chung, N.N.; Lu, Y. Conversion of δ-, κ-and μ-Receptor Selective Opioid Peptide Agonists into δ-, κ-and μ-Selective Antagonists. Life Sci. 2003, 73, 691–698. [Google Scholar] [CrossRef]
  44. Tancredi, T.; Salvadori, S.; Amodeo, P.; Picone, D.; Lazarus, L.H.; Bryant, S.D.; Guerrini, R.; Marzola, G.; Temussi, P.A. Conversion of enkephalin and dermorphin into delta-selective opioid antagonists by single-residue substitution. Eur. J. Biochem. 1994, 224, 241–247. [Google Scholar] [CrossRef]
  45. Witt, K.A.; Davis, T.P. CNS drug delivery: Opioid peptides and the blood-brain barrier. AAPS J. 2006, 8, E76–E88. [Google Scholar] [CrossRef] [PubMed]
  46. Lau, J.L.; Dunn, M.K. Therapeutic peptides: Historical perspectives, current development trends, and future directions. Bioorg. Med. Chem. 2018, 26, 2700–2707. [Google Scholar] [CrossRef] [PubMed]
  47. Gentilucci, L.; De Marco, R.; Cerisoli, L. Chemical modifications designed to improve peptide stability: Incorporation of non-natural amino acids, pseudo-peptide bonds, and cyclization. Curr. Pharm. Des. 2010, 16, 3185–3203. [Google Scholar] [CrossRef] [PubMed]
  48. Pert, C.B.; Pert, A.; Chang, J.K.; Fong, B.T. (D-Ala2)-Met-enkephalinamide: A potent, long-lasting synthetic pentapeptide analgesic. Science 1976, 194, 330–332. [Google Scholar] [CrossRef]
  49. Avan, I.; Hall, C.D.; Katritzky, A.R. Peptidomimetics via modifications of amino acids and peptide bonds. Chem. Soc. Rev. 2014, 43, 3575–3594. [Google Scholar] [CrossRef]
  50. Roberts, L.R.; Brady, K.; Brown, A.; Davey, D.; Feng, L.; Huang, H.; Liu, D.; Malet, L.; McMurray, G.; Phelan, A.; et al. Kappa agonist CovX-Bodies. Bioorg. Med. Chem. Lett. 2012, 22, 4173–4178. [Google Scholar] [CrossRef]
  51. Kropotova, E.S.; Ivleva, I.S.; Karpenko, M.N.; Mosevitsky, M.I. Design of enkephalin modifications protected from brain extracellular peptidases providing long-term analgesia. Bioorg. Med. Chem. 2020, 28, 115184. [Google Scholar] [CrossRef]
  52. Beaudeau, J.L.; Blais, V.; Holleran, B.J.; Bergeron, A.; Pineyro, G.; Guerin, B.; Gendron, L.; Dory, Y.L. N-Guanidyl and C-Tetrazole Leu-Enkephalin Derivatives: Efficient Mu and Delta Opioid Receptor Agonists with Improved Pharmacological Properties. ACS Chem. Neurosci. 2019, 10, 1615–1626. [Google Scholar] [CrossRef]
  53. Weber, S.J.; Abbruscato, T.J.; Brownson, E.A.; Lipkowski, A.W.; Polt, R.; Misicka, A.; Haaseth, R.C.; Bartosz, H.; Hruby, V.J.; Davis, T.P. Assessment of an in vitro blood-brain barrier model using several [Met5]enkephalin opioid analogs. J. Pharmacol. Exp. Ther. 1993, 266, 1649–1655. [Google Scholar]
  54. Banks, W.A.; Kastin, A.J. Peptides and the blood-brain barrier: Lipophilicity as a predictor of permeability. Brain Res. Bullet. 1985, 15, 287–292. [Google Scholar] [CrossRef]
  55. Hansen, D.W., Jr.; Stapelfeld, A.; Savage, M.A.; Reichman, M.; Hammond, D.L.; Haaseth, R.C.; Mosberg, H.I. Systemic analgesic activity and delta-opioid selectivity in [2,6-dimethyl-Tyr1,D-Pen2,D-Pen5]enkephalin. J. Med. Chem. 1992, 35, 684–687. [Google Scholar] [CrossRef]
  56. Weber, S.J.; Greene, D.L.; Sharma, S.D.; Yamamura, H.I.; Kramer, T.H.; Burks, T.F.; Hruby, V.J.; Hersh, L.B.; Davis, T.P. Distribution and analgesia of [3H][D-Pen2, D-Pen5]enkephalin and two halogenated analogs after intravenous administration. J. Pharmacol. Exp. Ther. 1991, 259, 1109–1117. [Google Scholar]
  57. Chandrakumar, N.S.; Stapelfeld, A.; Beardsley, P.M.; Lopez, O.T.; Drury, B.; Anthony, E.; Savage, M.A.; Williamson, L.N.; Reichman, M. Analogs of the. delta. opioid receptor selective cyclic peptide [cyclic][2-D-penicillamine, 5-D-penicillamine]-enkephalin: 2’, 6’-dimethyltyrosine and Gly3-Phe4 amide bond isostere substitutions. J. Med. Chem. 1992, 35, 2928–2938. [Google Scholar] [CrossRef]
  58. Jinsmaa, Y.; Miyazaki, A.; Fujita, Y.; Li, T.; Fujisawa, Y.; Shiotani, K.; Tsuda, Y.; Yokoi, T.; Ambo, A.; Sasaki, Y.; et al. Oral bioavailability of a new class of micro-opioid receptor agonists containing 3,6-bis[Dmt-NH(CH(2))(n)]-2(1H)-pyrazinone with central-mediated analgesia. J. Med. Chem. 2004, 47, 2599–2610. [Google Scholar] [CrossRef]
  59. Okada, Y.; Tsuda, Y.; Fujita, Y.; Yokoi, T.; Sasaki, Y.; Ambo, A.; Konishi, R.; Nagata, M.; Salvadori, S.; Jinsmaa, Y. Unique high-affinity synthetic μ-opioid receptor agonists with central-and systemic-mediated analgesia. J. Med. Chem. 2003, 46, 3201–3209. [Google Scholar] [CrossRef]
  60. Wang, C.L.; Qiu, T.T.; Yang, D.J.; Yuan, B.Y.; Han, F.T.; Li, L.; Gu, N. Endomorphin-2 analogs with C-terminal esterification produce potent systemic antinociception with reduced tolerance and gastrointestinal side effects. Neuropharmacology 2017, 116, 98–109. [Google Scholar] [CrossRef]
  61. Zhao, K.; Luo, G.; Zhao, G.M.; Schiller, P.W.; Szeto, H.H. Transcellular transport of a highly polar 3+ net charge opioid tetrapeptide. J. Pharmacol. Exp. Ther. 2003, 304, 425–432. [Google Scholar] [CrossRef]
  62. Terasaki, T.; Hirai, K.; Sato, H.; Kang, Y.; Tsuji, A. Absorptive-mediated endocytosis of a dynorphin-like analgesic peptide, E-2078 into the blood-brain barrier. J. Pharmacol. Exp. Therap. 1989, 251, 351–357. [Google Scholar]
  63. Deguchi, Y.; Miyakawa, Y.; Sakurada, S.; Naito, Y.; Morimoto, K.; Ohtsuki, S.; Hosoya, K.I.; Terasaki, T. Blood–brain barrier transport of a novel µ1-specific opioid peptide, H-Tyr-d-Arg-Phe-β-Ala-OH (TAPA). J. Neurochem. 2003, 84, 1154–1161. [Google Scholar] [CrossRef]
  64. Deguchi, Y.; Naito, Y.; Ohtsuki, S.; Miyakawa, Y.; Morimoto, K.; Hosoya, K.-I.; Sakurada, S.; Terasaki, T. Blood-brain barrier permeability of novel [D-arg2] dermorphin (1–4) analogs: Transport property is related to the slow onset of antinociceptive activity in the central nervous system. J. Pharmacol. Exp. Therap. 2004, 310, 177–184. [Google Scholar] [CrossRef]
  65. Polt, R.; Dhanasekaran, M.; Keyari, C.M. Glycosylated neuropeptides: A new vista for neuropsychopharmacology? Med. Res. Rev. 2005, 25, 557–585. [Google Scholar] [CrossRef]
  66. Egleton, R.D.; Mitchell, S.A.; Huber, J.D.; Janders, J.; Stropova, D.; Polt, R.; Yamamura, H.I.; Hruby, V.J.; Davis, T.P. Improved bioavailability to the brain of glycosylated Met-enkephalin analogs. Brain Res. 2000, 881, 37–46. [Google Scholar] [CrossRef]
  67. Egleton, R.D.; Mitchell, S.A.; Huber, J.D.; Palian, M.M.; Polt, R.; Davis, T.P. Improved blood-brain barrier penetration and enhanced analgesia of an opioid peptide by glycosylation. J. Pharmacol. Exp. Ther. 2001, 299, 967–972. [Google Scholar]
  68. Li, Y.; Lefever, M.R.; Muthu, D.; Bidlack, J.M.; Bilsky, E.J.; Polt, R. Opioid glycopeptide analgesics derived from endogenous enkephalins and endorphins. Fut. Med. Chem. 2012, 4, 205–226. [Google Scholar] [CrossRef]
  69. Fichna, J.; Mazur, M.; Grzywacz, D.; Kamysz, W.; Perlikowska, R.; Piekielna, J.; Sobczak, M.; Salaga, M.; Toth, G.; Janecka, A.; et al. Novel glycosylated endomorphin-2 analog produces potent centrally-mediated antinociception in mice after peripheral administration. Bioorg. Med. Chem. Lett. 2013, 23, 6673–6676. [Google Scholar] [CrossRef]
  70. Ballet, S.; Betti, C.; Novoa, A.; Tomboly, C.; Nielsen, C.U.; Helms, H.C.; Lesniak, A.; Kleczkowska, P.; Chung, N.N.; Lipkowski, A.W.; et al. In Vitro Membrane Permeation Studies and in Vivo Antinociception of Glycosylated Dmt(1)-DALDA Analogues. ACS Med. Chem. Lett. 2014, 5, 352–357. [Google Scholar] [CrossRef]
  71. Varamini, P.; Mansfeld, F.M.; Blanchfield, J.T.; Wyse, B.D.; Smith, M.T.; Toth, I. Synthesis and biological evaluation of an orally active glycosylated endomorphin-1. J. Med. Chem. 2012, 55, 5859–5867. [Google Scholar] [CrossRef]
  72. Tomatis, R.; Marastoni, M.; Balboni, G.; Guerrini, R.; Capasso, A.; Sorrentino, L.; Santagada, V.; Caliendo, G.; Lazarus, L.H.; Salvadori, S. Synthesis and pharmacological activity of deltorphin and dermorphin-related glycopeptides. J. Med. Chem. 1997, 40, 2948–2952. [Google Scholar] [CrossRef]
  73. Negri, L.; Lattanzi, R.; Tabacco, F.; Orru, L.; Severini, C.; Scolaro, B.; Rocchi, R. Dermorphin and deltorphin glycosylated analogues: Synthesis and antinociceptive activity after systemic administration. J. Med. Chem. 1999, 42, 400–404. [Google Scholar] [CrossRef]
  74. Mosberg, H.I.; Yeomans, L.; Anand, J.P.; Porter, V.; Sobczyk-Kojiro, K.; Traynor, J.R.; Jutkiewicz, E.M. Development of a bioavailable mu opioid receptor (MOPr) agonist, delta opioid receptor (DOPr) antagonist peptide that evokes antinociception without development of acute tolerance. J. Med. Chem. 2014, 57, 3148–3153. [Google Scholar] [CrossRef]
  75. Palian, M.M.; Boguslavsky, V.I.; O’Brien, D.F.; Polt, R. Glycopeptide-membrane interactions: Glycosyl enkephalin analogues adopt turn conformations by NMR and CD in amphipathic media. J. Am. Chem. Soc. 2003, 125, 5823–5831. [Google Scholar] [CrossRef]
  76. Witt, K.A.; Huber, J.D.; Egleton, R.D.; Roberts, M.J.; Bentley, M.D.; Guo, L.; Wei, H.; Yamamura, H.I.; Davis, T.P. Pharmacodynamic and pharmacokinetic characterization of poly(ethylene glycol) conjugation to met-enkephalin analog [D-Pen2, D-Pen5]-enkephalin (DPDPE). J. Pharmacol. Exp. Ther. 2001, 298, 848–856. [Google Scholar]
  77. Lindqvist, A.; Rip, J.; Gaillard, P.J.; Bjorkman, S.; Hammarlund-Udenaes, M. Enhanced brain delivery of the opioid peptide DAMGO in glutathione pegylated liposomes: A microdialysis study. Mol. Pharm. 2013, 10, 1533–1541. [Google Scholar] [CrossRef] [PubMed]
  78. Greene, D.L.; Hau, V.S.; Abbruscato, T.J.; Bartosz, H.; Misicka, A.; Lipkowski, A.W.; Hom, S.; Gillespie, T.J.; Hruby, V.J.; Davis, T.P. Enkephalin analog prodrugs: Assessment of in vitro conversion, enzyme cleavage characterization and blood-brain barrier permeability. J. Pharmacol. Exp. Ther. 1996, 277, 1366–1375. [Google Scholar] [PubMed]
  79. Ouyang, H.; Tang, F.; Siahaan, T.J.; Borchardt, R.T. A modified coumarinic acid-based cyclic prodrug of an opioid peptide: Its enzymatic and chemical stability and cell permeation characteristics. Pharm. Res. 2002, 19, 794–801. [Google Scholar] [CrossRef]
  80. Yang, J.Z.; Chen, W.; Borchardt, R.T. In vitro stability and in vivo pharmacokinetic studies of a model opioid peptide, H-Tyr-D-Ala-Gly-Phe-D-Leu-OH (DADLE), and its cyclic prodrugs. J. Pharmacol. Exp. Ther. 2002, 303, 840–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Prokai-Tatrai, K.; Kim, H.-S.; Prokai, L. The utility of oligopeptidase in brain-targeting delivery of an enkephalin analogue by prodrug design. Open Med. Chem. J. 2008, 2, 97. [Google Scholar] [CrossRef]
  82. Liederer, B.M.; Borchardt, R.T. Stability of oxymethyl-modified coumarinic acid cyclic prodrugs of diastereomeric opioid peptides in biological media from various animal species including human. J. Pharm. Sci. 2005, 94, 2198–2206. [Google Scholar] [CrossRef]
  83. Wang, J.; Hogenkamp, D.J.; Tran, M.; Li, W.-Y.; Yoshimura, R.F.; Johnstone, T.B.; Shen, W.-C.; Gee, K.W. Reversible lipidization for the oral delivery of leu-enkephalin. J. Drug Target. 2006, 14, 127–136. [Google Scholar] [CrossRef]
  84. Ogawa, T.; Araki, M.; Miyamae, T.; Okayama, T.; Hagiwara, M.; Sakurada, S.; Morikawa, T. Synthesis and antinociceptive activity of orally active opioid peptides: Improvement of oral bioavailability by esterification. Chem. Pharm. Bull. 2003, 51, 759–771. [Google Scholar] [CrossRef]
  85. Ogawa, T.; Miyamae, T.; Murayama, K.; Okuyama, K.; Okayama, T.; Hagiwara, M.; Sakurada, S.; Morikawa, T. Synthesis and structure-activity relationships of an orally available and long-acting analgesic peptide, N(alpha)-amidino-Tyr-D-Arg-Phe-MebetaAla-OH (ADAMB). J. Med. Chem. 2002, 45, 5081–5089. [Google Scholar] [CrossRef]
  86. Machelska, H.; Celik, M.O. Advances in Achieving Opioid Analgesia Without Side Effects. Front. Pharmacol. 2018, 9, 1388. [Google Scholar] [CrossRef]
  87. Anand, J.P.; Montgomery, D. Multifunctional Opioid Ligands. Handb. Exp. Pharmacol. 2018, 247, 21–51. [Google Scholar] [CrossRef]
  88. Hruby, V.J. Multivalent peptide and peptidomimetic ligands for the treatment of pain without toxicities and addiction. Peptides 2019, 116, 63–67. [Google Scholar] [CrossRef]
  89. Dietis, N.; Guerrini, R.; Calo, G.; Salvadori, S.; Rowbotham, D.J.; Lambert, D.G. Simultaneous targeting of multiple opioid receptors: A strategy to improve side-effect profile. Br. J. Anaesth 2009, 103, 38–49. [Google Scholar] [CrossRef]
  90. Zhu, Y.; King, M.A.; Schuller, A.G.; Nitsche, J.F.; Reidl, M.; Elde, R.P.; Unterwald, E.; Pasternak, G.W.; Pintar, J.E. Retention of supraspinal delta-like analgesia and loss of morphine tolerance in delta opioid receptor knockout mice. Neuron 1999, 24, 243–252. [Google Scholar] [CrossRef] [Green Version]
  91. Mosberg, H.I.; Yeomans, L.; Harland, A.A.; Bender, A.M.; Sobczyk-Kojiro, K.; Anand, J.P.; Clark, M.J.; Jutkiewicz, E.M.; Traynor, J.R. Opioid peptidomimetics: Leads for the design of bioavailable mixed efficacy mu opioid receptor (MOR) agonist/delta opioid receptor (DOR) antagonist ligands. J. Med. Chem. 2013, 56, 2139–2149. [Google Scholar] [CrossRef]
  92. Henry, S.; Anand, J.P.; Twarozynski, J.J.; Brinkel, A.C.; Pogozheva, I.D.; Sears, B.F.; Jutkiewicz, E.M.; Traynor, J.R.; Mosberg, H.I. Aromatic–Amine Pendants Produce Highly Potent and Efficacious Mixed Efficacy μ-Opioid Receptor (MOR)/δ-Opioid Receptor (DOR) Peptidomimetics with Enhanced Metabolic Stability. J. Med. Chem. 2020, 63, 1671–1683. [Google Scholar] [CrossRef]
  93. Lee, Y.S.; Kulkarani, V.; Cowell, S.M.; Ma, S.W.; Davis, P.; Hanlon, K.E.; Vanderah, T.W.; Lai, J.; Porreca, F.; Vardanyan, R.; et al. Development of potent mu and delta opioid agonists with high lipophilicity. J. Med. Chem. 2011, 54, 382–386. [Google Scholar] [CrossRef]
  94. Cowell, S.M.; Lee, Y.S. Biphalin: The Foundation of Bivalent Ligands. Curr. Med. Chem. 2016, 23, 3267–3284. [Google Scholar] [CrossRef]
  95. Horan, P.J.; Mattia, A.; Bilsky, E.J.; Weber, S.; Davis, T.P.; Yamamura, H.I.; Malatynska, E.; Appleyard, S.M.; Slaninova, J.; Misicka, A.; et al. Antinociceptive profile of biphalin, a dimeric enkephalin analog. J. Pharmacol. Exp. Ther. 1993, 265, 1446–1454. [Google Scholar]
  96. Lowery, J.J.; Raymond, T.J.; Giuvelis, D.; Bidlack, J.M.; Polt, R.; Bilsky, E.J. In vivo characterization of MMP-2200, a mixed delta/mu opioid agonist, in mice. J. Pharmacol. Exp. Ther. 2011, 336, 767–778. [Google Scholar] [CrossRef]
  97. Li, T.; Shiotani, K.; Miyazaki, A.; Tsuda, Y.; Ambo, A.; Sasaki, Y.; Jinsmaa, Y.; Marczak, E.; Bryant, S.D.; Lazarus, L.H.; et al. Bifunctional [2′,6′-dimethyl-L-tyrosine1]endomorphin-2 analogues substituted at position 3 with alkylated phenylalanine derivatives yield potent mixed mu-agonist/delta-antagonist and dual mu-agonist/delta-agonist opioid ligands. J. Med. Chem. 2007, 50, 2753–2766. [Google Scholar] [CrossRef]
  98. Frederickson, R.C.; Smithwick, E.L.; Shuman, R.; Bemis, K.G. Metkephamid, a systemically active analog of methionine enkephalin with potent opioid alpha-receptor activity. Science 1981, 211, 603–605. [Google Scholar] [CrossRef]
  99. Burkhardt, C.; Frederickson, R.C.; Pasternak, G.W. Metkephamid (Tyr-D-ala-Gly-Phe-N(Me)Met-NH2), a potent opioid peptide: Receptor binding and analgesic properties. Peptides 1982, 3, 869–871. [Google Scholar] [CrossRef]
  100. Schiller, P.W.; Schmidt, R.; Weltrowska, G.; Berezowska, I.; Nguyen, T.M.-D.; Dupuis, S.; Chung, N.N.; Lemieux, C.; Wilkes, B.C.; Carpenter, K.A. Conformationally constrained opioid peptide analogs with novel activity profiles. Lett. Peptide Sci. 1998, 5, 209–214. [Google Scholar] [CrossRef]
  101. Dietis, N.; McDonald, J.; Molinari, S.; Calo, G.; Guerrini, R.; Rowbotham, D.J.; Lambert, D.G. Pharmacological characterization of the bifunctional opioid ligand H-Dmt-Tic-Gly-NH-Bzl (UFP-505). Br. J. Anaesth 2012, 108, 262–270. [Google Scholar] [CrossRef] [Green Version]
  102. Purington, L.C.; Sobczyk-Kojiro, K.; Pogozheva, I.D.; Traynor, J.R.; Mosberg, H.I. Development and in vitro characterization of a novel bifunctional mu-agonist/delta-antagonist opioid tetrapeptide. ACS Chem. Biol. 2011, 6, 1375–1381. [Google Scholar] [CrossRef]
  103. Harland, A.A.; Yeomans, L.; Griggs, N.W.; Anand, J.P.; Pogozheva, I.D.; Jutkiewicz, E.M.; Traynor, J.R.; Mosberg, H.I. Further optimization and evaluation of bioavailable, mixed-efficacy μ-opioid receptor (MOR) agonists/δ-opioid receptor (DOR) antagonists: Balancing MOR and DOR affinities. J. Med. Chem. 2015, 58, 8952–8969. [Google Scholar] [CrossRef]
  104. Henry, S.P.; Fernandez, T.J.; Anand, J.P.; Griggs, N.W.; Traynor, J.R.; Mosberg, H.I. Structural Simplification of a Tetrahydroquinoline-Core Peptidomimetic mu-Opioid Receptor (MOR) Agonist/delta-Opioid Receptor (DOR) Antagonist Produces Improved Metabolic Stability. J. Med. Chem. 2019, 62, 4142–4157. [Google Scholar] [CrossRef]
  105. Bender, A.M.; Clark, M.J.; Agius, M.P.; Traynor, J.R.; Mosberg, H.I. Synthesis and evaluation of 4-substituted piperidines and piperazines as balanced affinity mu opioid receptor (MOR) agonist/delta opioid receptor (DOR) antagonist ligands. Bioorg. Med. Chem. Lett. 2014, 24, 548–551. [Google Scholar] [CrossRef]
  106. Lee, Y.S.; Remesic, M.; Ramos-Colon, C.; Wu, Z.; LaVigne, J.; Molnar, G.; Tymecka, D.; Misicka, A.; Streicher, J.M.; Hruby, V.J. Multifunctional Enkephalin Analogs with a New Biological Profile: MOR/DOR Agonism and KOR Antagonism. Biomedicines 2021, 9, 625. [Google Scholar] [CrossRef]
  107. Perlikowska, R.; Malfacini, D.; Cerlesi, M.C.; Piekielna, J.; Floriot, L.; Henry, T.; Do-Rego, J.C.; Tömböly, C.; Kluczyk, A.; Janecka, A. Pharmacological characterization of endomorphin-2-based cyclic pentapeptides with methylated phenylalanine residues. Peptides 2014, 55, 145–150. [Google Scholar] [CrossRef]
  108. Gach-Janczak, K.; Piekielna-Ciesielska, J.; Adamska-Bartłomiejczyk, A.; Perlikowska, R.; Kruszyński, R.; Kluczyk, A.; Krzywik, J.; Sukiennik, J.; Cerlesi, M.C.; Calo, G. Synthesis and activity of opioid peptidomimetics with β2-and β3-amino acids. Peptides 2017, 95, 116–123. [Google Scholar] [CrossRef]
  109. Martinez, V.; Abalo, R. Peripherally acting opioid analgesics and peripherally-induced analgesia. Behav. Pharmacol. 2020, 31, 136–158. [Google Scholar] [CrossRef]
  110. Zaitseva, N.; Galan, S.; Pavlova, L. Prospects of a search for kappa-opioid receptor agonists with analgesic activity. Pharm. Chem. J. 2018, 51, 843–851. [Google Scholar] [CrossRef]
  111. Spampinato, S.; Qasem, A.R.; Calienni, M.; Murari, G.; Gentilucci, L.; Tolomelli, A.; Cardillo, G. Antinociception by a peripherally administered novel endomorphin-1 analogue containing beta-proline. Eur. J. Pharmacol. 2003, 469, 89–95. [Google Scholar] [CrossRef]
  112. Hesselink, J.M.K. CR845 (Difelikefalin), A Kappa Receptors Agonist In Phase III By CARA Therapeutics: A Case Of ‘Spin’In Scientific Writing? J. Pharmacol. Clin. Res. 2017, 2, 555588. [Google Scholar] [CrossRef]
  113. Wallace, M.S.; Moulin, D.; Clark, A.; Wasserman, R.; Neale, A.; Morley-Forster, P.; Castaigne, J.-P.; Teichman, S. A Phase II, multicenter, randomized, double-blind, placebo-controlled crossover study of CJC-1008—a long-acting, parenteral opioid analgesic—in the treatment of postherpetic neuralgia. J. Opioid. Manag. 2006, 2, 167–173. [Google Scholar] [CrossRef]
  114. Tiwari, V.; Yang, F.; He, S.Q.; Shechter, R.; Zhang, C.; Shu, B.; Zhang, T.; Tiwari, V.; Wang, Y.; Dong, X.; et al. Activation of Peripheral mu-opioid Receptors by Dermorphin [D-Arg2, Lys4] (1-4) Amide Leads to Modality-preferred Inhibition of Neuropathic Pain. Anesthesiology 2016, 124, 706–720. [Google Scholar] [CrossRef]
  115. Posner, J.; Moody, S.; Peck, A.; Rutter, D.; Telekes, A. Analgesic, central, cardiovascular and endocrine effects of the enkephalin analogue Tyr-D. Arg-Gly-Phe (4NO 2)-Pro-NH 2 (443C81) in healthy volunteers. Eur. J. Clin. Pharmacol. 1990, 38, 213–218. [Google Scholar] [CrossRef]
  116. Deeks, E.D. Difelikefalin: First Approval. Drugs 2021, 81, 1937–1944. [Google Scholar] [CrossRef]
  117. DeWire, S.M.; Yamashita, D.S.; Rominger, D.H.; Liu, G.; Cowan, C.L.; Graczyk, T.M.; Chen, X.T.; Pitis, P.M.; Gotchev, D.; Yuan, C.; et al. A G protein-biased ligand at the mu-opioid receptor is potently analgesic with reduced gastrointestinal and respiratory dysfunction compared with morphine. J. Pharmacol. Exp. Ther. 2013, 344, 708–717. [Google Scholar] [CrossRef]
  118. Piekielna-Ciesielska, J.; Wtorek, K.; Janecka, A. Biased Agonism as an Emerging Strategy in the Search for Better Opioid Analgesics. Curr. Med. Chem. 2020, 27, 1562–1575. [Google Scholar] [CrossRef]
  119. He, L.; Gooding, S.W.; Lewis, E.; Felth, L.C.; Gaur, A.; Whistler, J.L. Pharmacological and genetic manipulations at the µ-opioid receptor reveal arrestin-3 engagement limits analgesic tolerance and does not exacerbate respiratory depression in mice. Neuropsychopharmacology 2021, 46, 2241–2249. [Google Scholar] [CrossRef]
  120. Faouzi, A.; Varga, B.R.; Majumdar, S. Biased opioid ligands. Molecules 2020, 25, 4257. [Google Scholar] [CrossRef]
  121. Madariaga-Mazón, A.; Marmolejo-Valencia, A.F.; Li, Y.; Toll, L.; Houghten, R.A.; Martinez-Mayorga, K. Mu-Opioid receptor biased ligands: A safer and painless discovery of analgesics? Drug Discov. Today 2017, 22, 1719–1729. [Google Scholar] [CrossRef]
  122. Piekielna-Ciesielska, J.; Ferrari, F.; Calo, G.; Janecka, A. Cyclopeptide Dmt-[D-Lys-p-CF3-Phe-Phe-Asp]NH2, a novel G protein-biased agonist of the mu opioid receptor. Peptides 2018, 101, 227–233. [Google Scholar] [CrossRef]
  123. Bella Ndong, D.; Blais, V.; Holleran, B.J.; Proteau-Gagné, A.; Cantin-Savoie, I.; Robert, W.; Nadon, J.F.; Beauchemin, S.; Leduc, R.; Piñeyro, G. Exploration of the fifth position of leu-enkephalin and its role in binding and activating delta (DOP) and mu (MOP) opioid receptors. Peptide Sci. 2019, 111, e24070. [Google Scholar] [CrossRef]
  124. Sharma, K.K.; Cassell, R.J.; Su, H.; Blaine, A.T. Modulating β Arrestin-2 Recruitment at the δ- and µ-Opioid Receptors; Cambridge Open Engage: Cambridge, UK, 2020. [Google Scholar]
  125. Cassell, R.J.; Sharma, K.K.; Su, H.; Cummins, B.R.; Cui, H.; Mores, K.L.; Blaine, A.T.; Altman, R.A.; van Rijn, R.M. The Meta-Position of Phe4 in Leu-enkephalin Regulates Potency, Selectivity, Functional Activity, and Signaling Bias at the Delta and Mu Opioid Receptors. Molecules 2019, 24, 4542. [Google Scholar] [CrossRef]
  126. Kandasamy, R.; Hillhouse, T.M.; Livingston, K.E.; Kochan, K.E.; Meurice, C.; Eans, S.O.; Li, M.-H.; White, A.D.; Roques, B.P.; McLaughlin, J.P. Positive allosteric modulation of the mu-opioid receptor produces analgesia with reduced side effects. Proc. Natl. Acad. Sci. USA 2021, 118, e2000017118. [Google Scholar] [CrossRef]
  127. Remesic, M.; Hruby, V.J.; Porreca, F.; Lee, Y.S. Recent Advances in the Realm of Allosteric Modulators for Opioid Receptors for Future Therapeutics. ACS Chem. Neurosci. 2017, 8, 1147–1158. [Google Scholar] [CrossRef]
  128. Cowell, S.M.; Lee, Y.S.; Cain, J.P.; Hruby, V.J. Exploring Ramachandran and chi space: Conformationally constrained amino acids and peptides in the design of bioactive polypeptide ligands. Curr. Med. Chem. 2004, 11, 2785–2798. [Google Scholar] [CrossRef]
  129. Yamazaki, T.; Ro, S.; Goodman, M.; Chung, N.N.; Schiller, P.W. A topochemical approach to explain morphiceptin bioactivity. J. Med. Chem. 1993, 36, 708–719. [Google Scholar] [CrossRef]
  130. Granier, S.; Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Weis, W.I.; Kobilka, B.K. Structure of the delta-opioid receptor bound to naltrindole. Nature 2012, 485, 400–404. [Google Scholar] [CrossRef]
  131. Chavkin, C.; Goldstein, A. Specific receptor for the opioid peptide dynorphin: Structure--activity relationships. Proc. Natl. Acad. Sci. USA 1981, 78, 6543–6547. [Google Scholar] [CrossRef]
  132. Portoghese, P.S.; Sultana, M.; Takemori, A.E. Design of peptidomimetic delta opioid receptor antagonists using the message-address concept. J. Med. Chem. 1990, 33, 1714–1720. [Google Scholar] [CrossRef] [PubMed]
  133. Borics, A.; Toth, G. Structural comparison of mu-opioid receptor selective peptides confirmed four parameters of bioactivity. J. Mol. Graph Model 2010, 28, 495–505. [Google Scholar] [CrossRef]
  134. Lasota, A.; Frączak, O.; Muchowska, A.; Nowakowski, M.; Maciejczyk, M.; Ejchart, A.; Olma, A. Synthesis, Biological Activity, and NMR-Based Structural Studies of Deltorphin I Analogs Modified in Message Domain with a New α, α-Disubstituted Glycines. Chem. Biol. Drug Design 2016, 87, 824–832. [Google Scholar] [CrossRef] [PubMed]
  135. Gentilucci, L.; Tolomelli, A. Recent advances in the investigation of the bioactive conformation of peptides active at the μ-opioid receptor. Conformational analysis of endomorphins. Curr. Top. Med. Chem. 2004, 4, 105–121. [Google Scholar] [CrossRef] [PubMed]
  136. Mallareddy, J.R.; Borics, A.; Keresztes, A.; Kover, K.E.; Tourwe, D.; Toth, G. Design, synthesis, pharmacological evaluation, and structure-activity study of novel endomorphin analogues with multiple structural modifications. J. Med. Chem. 2011, 54, 1462–1472. [Google Scholar] [CrossRef]
  137. Keller, M.; Boissard, C.; Patiny, L.; Chung, N.N.; Lemieux, C.; Mutter, M.; Schiller, P.W. Pseudoproline-containing analogues of morphiceptin and endomorphin-2: Evidence for a cis Tyr-Pro amide bond in the bioactive conformation. J. Med. Chem. 2001, 44, 3896–3903. [Google Scholar] [CrossRef] [PubMed]
  138. Doi, M.; Asano, A.; Komura, E.; Ueda, Y. The structure of an endomorphin analogue incorporating 1-aminocyclohexane-1-carboxlylic acid for proline is similar to the beta-turn of Leu-enkephalin. Biochem. Biophys. Res. Commun. 2002, 297, 138–142. [Google Scholar] [CrossRef]
  139. Eguchi, M.; Shen, R.Y.; Shea, J.P.; Lee, M.S.; Kahn, M. Design, synthesis, and evaluation of opioid analogues with non-peptidic beta-turn scaffold: Enkephalin and endomorphin mimetics. J. Med. Chem. 2002, 45, 1395–1398. [Google Scholar] [CrossRef]
  140. Giordano, C.; Sansone, A.; Masi, A.; Lucente, G.; Punzi, P.; Mollica, A.; Pinnen, F.; Feliciani, F.; Cacciatore, I.; Davis, P.; et al. Synthesis and activity of endomorphin-2 and morphiceptin analogues with proline surrogates in position 2. Eur. J. Med. Chem. 2010, 45, 4594–4600. [Google Scholar] [CrossRef]
  141. Keresztes, A.; Szucs, M.; Borics, A.; Kover, K.E.; Forro, E.; Fulop, F.; Tomboly, C.; Peter, A.; Pahi, A.; Fabian, G.; et al. New endomorphin analogues containing alicyclic beta-amino acids: Influence on bioactive conformation and pharmacological profile. J. Med. Chem. 2008, 51, 4270–4279. [Google Scholar] [CrossRef]
  142. Perlikowska, R.; Fichna, J.; Wyrebska, A.; Poels, J.; Vanden Broeck, J.; Toth, G.; Storr, M.; do Rego, J.C.; Janecka, A. Design, synthesis and pharmacological characterization of endomorphin analogues with non-cyclic amino acid residues in position 2. Basic Clin. Pharmacol. Toxicol. 2010, 106, 106–113. [Google Scholar] [CrossRef]
  143. Torino, D.; Mollica, A.; Pinnen, F.; Lucente, G.; Feliciani, F.; Davis, P.; Lai, J.; Ma, S.W.; Porreca, F.; Hruby, V.J. Synthesis and evaluation of new endomorphin analogues modified at the Pro(2) residue. Bioorg. Med. Chem. Lett. 2009, 19, 4115–4118. [Google Scholar] [CrossRef]
  144. Staniszewska, R.; Fichna, J.; Gach, K.; Toth, G.; Poels, J.; Vanden Broeck, J.; Janecka, A. Synthesis and biological activity of endomorphin-2 analogs incorporating piperidine-2-, 3- or 4-carboxylic acids instead of proline in position 2. Chem. Biol. Drug Des. 2008, 72, 91–94. [Google Scholar] [CrossRef]
  145. Perlikowska, R.; Gach, K.; Fichna, J.; Toth, G.; Walkowiak, B.; do-Rego, J.C.; Janecka, A. Biological activity of endomorphin and [Dmt1]endomorphin analogs with six-membered proline surrogates in position 2. Bioorg. Med. Chem. 2009, 17, 3789–3794. [Google Scholar] [CrossRef]
  146. Fujita, Y.; Tsuda, Y.; Li, T.; Motoyama, T.; Takahashi, M.; Shimizu, Y.; Yokoi, T.; Sasaki, Y.; Ambo, A.; Kita, A.; et al. Development of potent bifunctional endomorphin-2 analogues with mixed mu-/delta-opioid agonist and delta-opioid antagonist properties. J. Med. Chem. 2004, 47, 3591–3599. [Google Scholar] [CrossRef]
  147. Cardillo, G.; Gentilucci, L.; Tolomelli, A.; Spinosa, R.; Calienni, M.; Qasem, A.R.; Spampinato, S. Synthesis and evaluation of the affinity toward mu-opioid receptors of atypical, lipophilic ligands based on the sequence c[-Tyr-Pro-Trp-Phe-Gly-]. J. Med. Chem. 2004, 47, 5198–5203. [Google Scholar] [CrossRef]
  148. Honda, T.; Shirasu, N.; Isozaki, K.; Kawano, M.; Shigehiro, D.; Chuman, Y.; Fujita, T.; Nose, T.; Shimohigashi, Y. Differential receptor binding characteristics of consecutive phenylalanines in micro-opioid specific peptide ligand endomorphin-2. Bioorg. Med. Chem. 2007, 15, 3883–3888. [Google Scholar] [CrossRef]
  149. Mizoguchi, H.; Bagetta, G.; Sakurada, T.; Sakurada, S. Dermorphin tetrapeptide analogs as potent and long-lasting analgesics with pharmacological profiles distinct from morphine. Peptides 2011, 32, 421–427. [Google Scholar] [CrossRef]
  150. Negri, L.; Melchiorri, P.; Lattanzi, R. Pharmacology of amphibian opiate peptides. Peptides 2000, 21, 1639–1647. [Google Scholar] [CrossRef]
  151. Chaki, K.; Kawamura, S.; Kisara, K.; Sakurada, S.; Sakurada, T.; Sasaki, Y.; Sato, T.; Susuki, K. Antinociception and physical dependence produced by [D-Arg2] dermorphin tetrapeptide analogues and morphine in rats. Br. J. Pharmacol. 1988, 95, 15–22. [Google Scholar] [CrossRef]
  152. Sato, T.; Sakurada, S.; Sakurada, T.; Furuta, S.; Nakata, N.; Kisara, K.; Sasaki, Y.; Suzuki, K. Comparison of the antinociceptive effect between D-Arg containing dipeptides and tetrapeptides in mice. Neuropeptides 1984, 4, 269–279. [Google Scholar] [CrossRef]
  153. Sakurada, S.; Takeda, S.; Sato, T.; Hayashi, T.; Yuki, M.; Kutsuwa, M.; Tan-No, K.; Sakurada, C.; Kisara, K.; Sakurada, T. Selective antagonism by naloxonazine of antinociception by Tyr-D-Arg-Phe-beta-Ala, a novel dermorphin analogue with high affinity at mu-opioid receptors. Eur. J. Pharmacol. 2000, 395, 107–112. [Google Scholar] [CrossRef]
  154. Schiller, P.W.; Nguyen, T.M.-D.; Berezowska, I.; Dupuis, S.; Weltrowska, G.; Chung, N.N.; Lemieux, C. Synthesis and in vitro opioid activity profiles of DALDA analogues. Eur. J. Med. Chem. 2000, 35, 895–901. [Google Scholar] [CrossRef]
  155. Vandormael, B.; Fourla, D.-D.; Gramowski-Voß, A.; Kosson, P.; Weiss, D.G.; Schröder, O.H.-U.; Lipkowski, A.; Georgoussi, Z.; Tourwé, D. Superpotent [Dmt1] Dermorphin tetrapeptides containing the 4-aminotetrahydro-2-benzazepin-3-one scaffold with mixed μ/δ opioid receptor agonistic properties. J. Med. Chem. 2011, 54, 7848–7859. [Google Scholar] [CrossRef]
  156. Chang, K.J.; Lillian, A.; Hazum, E.; Cuatrecasas, P.; Chang, J.K. Morphiceptin (NH4-tyr-pro-phe-pro-COHN2): A potent and specific agonist for morphine (mu) receptors. Science 1981, 212, 75–77. [Google Scholar] [CrossRef]
  157. Yamazaki, T.; Probsti, A.; Schiller, P.W.; Goodman, M. Biological and conformational studies of [Val4]morphiceptin and [D-Val4]morphiceptin analogs incorporating cis-2-aminocyclopentane carboxylic acid as a peptidomimetic for proline. Int. J. Pept Protein Res. 1991, 37, 364–381. [Google Scholar] [CrossRef]
  158. Chang, K.J.; Wei, E.T.; Killian, A.; Chang, J.K. Potent morphiceptin analogs: Structure activity relationships and morphine-like activities. J. Pharmacol. Exp. Ther. 1983, 227, 403–408. [Google Scholar]
  159. Fichna, J.; do-Rego, J.-C.; Costentin, J.; Chung, N.N.; Schiller, P.W.; Kosson, P.; Janecka, A. Opioid receptor binding and in vivo antinociceptive activity of position 3-substituted morphiceptin analogs. Biochem. Biophys. Res. Commun. 2004, 320, 531–536. [Google Scholar] [CrossRef]
  160. Ambo, A.; Terashima, T.; Sasaki, Y. Novel [D-Arg2]dermorphin(1-4) analogs with mu-opioid receptor antagonist activity. Chem. Pharm. Bull. 2002, 50, 1401–1403. [Google Scholar] [CrossRef]
  161. Kazmierski, W.; Wire, W.S.; Lui, G.K.; Knapp, R.J.; Shook, J.E.; Burks, T.F.; Yamamura, H.I.; Hruby, V.J. Design and synthesis of somatostatin analogues with topographical properties that lead to highly potent and specific mu opioid receptor antagonists with greatly reduced binding at somatostatin receptors. J. Med. Chem. 1988, 31, 2170–2177. [Google Scholar] [CrossRef]
  162. Hruby, V.J.; Gehrig, C.A. Recent developments in the design of receptor specific opioid peptides. Med. Res. Rev. 1989, 9, 343–401. [Google Scholar] [CrossRef] [PubMed]
  163. Fournie-Zaluski, M.C.; Gacel, G.; Maigret, B.; Premilat, S.; Roques, B.P. Structural requirements for specific recognition of mu or delta opiate receptors. Mol. Pharmacol. 1981, 20, 484–491. [Google Scholar] [PubMed]
  164. Rochon, K.; Proteau-Gagné, A.; Bourassa, P.; Nadon, J.-F.o.; Côté, J.; Bournival, V.r.; Gobeil, F., Jr.; Guérin, B.; Dory, Y.L.; Gendron, L. Preparation and evaluation at the delta opioid receptor of a series of linear leu-enkephalin analogues obtained by systematic replacement of the amides. ACS Chem. Neurosci. 2013, 4, 1204–1216. [Google Scholar] [CrossRef] [PubMed]
  165. Nadon, J.-F.o.; Rochon, K.; Grastilleur, S.b.; Langlois, G.; Dao, T.T.H.; Blais, V.r.; Guérin, B.; Gendron, L.; Dory, Y.L. Synthesis of Gly-ψ [(Z) CF= CH]-Phe, a Fluoroalkene Dipeptide Isostere, and Its Incorporation into a Leu-enkephalin Peptidomimetic. ACS Chem. Neurosci. 2017, 8, 40–49. [Google Scholar] [CrossRef]
  166. Karad, S.N.; Pal, M.; Crowley, R.S.; Prisinzano, T.E.; Altman, R.A. Synthesis and Opioid Activity of Tyr1-ψ [(Z) CF= CH]-Gly2 and Tyr1-ψ [(S)/(R)-CF3CH-NH]-Gly2 Leu-enkephalin Fluorinated Peptidomimetics. ChemMedChem 2017, 12, 571–576. [Google Scholar] [CrossRef]
  167. Haaseth, R.C.; Sobczyk-Kojiro, K.; Medzihradsky, F.; Smith, C.B.; Mosberg, H.I. Single residue modifications of the delta opioid receptor selective peptide,[d-Pen2, d-Pen5]-enkephalin (DPDPE) Correlation of pharmacological effects with structural and conformational features. Int. J. Peptide Protein Res. 1990, 36, 139–146. [Google Scholar] [CrossRef]
  168. Mosberg, H.I.; Omnaas, J.R.; Medzihradsky, F.; Smith, C.B. Cyclic, disulfide-and dithioether-containing opioid tetrapeptides: Development of a ligard with high delta opioid receptor selectivity and affinity. Life Sci. 1988, 43, 1013–1020. [Google Scholar] [CrossRef]
  169. Liao, S.; Alfaro-Lopez, J.; Shenderovich, M.D.; Hosohata, K.; Lin, J.; Li, X.; Stropova, D.; Davis, P.; Jernigan, K.A.; Porreca, F. De novo design, synthesis, and biological activities of high-affinity and selective non-peptide agonists of the δ-opioid receptor. J. Med. Chem. 1998, 41, 4767–4776. [Google Scholar] [CrossRef]
  170. Liao, S.; Shenderovich, M.; Kover, K.E.; Zhang, Z.; Hosohata, K.; Davis, P.; Porreca, F.; Yamamura, H.I.; Hurby, V.J. Synthesis, biology, NMR and conformation studies of the topographically constrained delta-opioid selective peptide analogs of [beta-iPrPhe(3)]deltorphin I. J. Pept. Res. 2001, 57, 257–276. [Google Scholar] [CrossRef]
  171. Schullery, S.E.; Rodgers, D.W.; Tripathy, S.; Jayamaha, D.E.; Sanvordekar, M.D.; Renganathan, K.; Mousigian, C.; Heyl, D.L. The role of backbone conformation in deltorphin II binding: A QSAR study of new analogues modified in the 5-, 6-positions of the address domain. Bioorg. Med. Chem. 2001, 9, 2633–2642. [Google Scholar] [CrossRef]
  172. Breveglieri, A.; Guerrini, R.; Salvadori, S.; Bianchi, C.; Bryant, S.D.; Attila, M.; Lazarus, L.H. Design and synthesis of 1-aminocycloalkane-1-carboxylic acid-substituted deltorphin analogues: Unique delta and mu opioid activity in modified peptides. J. Med. Chem. 1996, 39, 773–780. [Google Scholar] [CrossRef]
  173. Misicka, A.; Lipkowski, A.W.; Horvath, R.; Davis, P.; Yamamura, H.I.; Porreca, F.; Hruby, V.J. Design of cyclic deltorphins and dermenkephalins with a disulfide bridge leads to analogs with high selectivity for. delta.-opioid receptors. J. Med. Chem. 1994, 37, 141–145. [Google Scholar] [CrossRef]
  174. Cotton, R.; Giles, M.G.; Miller, L.; Shaw, J.S.; Timms, D. ICI 174864: A highly selective antagonist for the opioid delta-receptor. Eur. J. Pharmacol. 1984, 97, 331–332. [Google Scholar] [CrossRef]
  175. Schiller, P.W.; Nguyen, T.M.; Weltrowska, G.; Wilkes, B.C.; Marsden, B.J.; Lemieux, C.; Chung, N.N. Differential stereochemical requirements of mu vs. delta opioid receptors for ligand binding and signal transduction: Development of a class of potent and highly delta-selective peptide antagonists. Proc. Natl. Acad. Sci. USA 1992, 89, 11871–11875. [Google Scholar] [CrossRef]
  176. Schiller, P.W.; Weltrowska, G.; Berezowska, I.; Nguyen, T.M.D.; Wilkes, B.C.; Lemieux, C.; Chung, N.N. The TIPP opioid peptide family: Development of δ antagonists, δ agonists, and mixed μ agonist/δ antagonists. Peptide Sci. 1999, 51, 411–425. [Google Scholar] [CrossRef]
  177. Schiller, P.W.; Fundytus, M.E.; Merovitz, L.; Weltrowska, G.; Nguyen, T.M.-D.; Lemieux, C.; Chung, N.N.; Coderre, T.J. The opioid μ agonist/δ antagonist DIPP-NH2 [Ψ] produces a potent analgesic effect, no physical dependence, and less tolerance than morphine in rats. J. Med. Chem. 1999, 42, 3520–3526. [Google Scholar] [CrossRef]
  178. Pagé, D.; Naismith, A.; Schmidt, R.; Coupal, M.; Labarre, M.; Gosselin, M.; Bellemare, D.; Payza, K.; Brown, W. Novel C-terminus modifications of the Dmt-Tic motif: A new class of dipeptide analogues showing altered pharmacological profiles toward the opioid receptors. J. Med. Chem. 2001, 44, 2387–2390. [Google Scholar] [CrossRef]
  179. Salvadori, S.; Attila, M.; Balboni, G.; Bianchi, C.; Bryant, S.D.; Crescenzi, O.; Guerrini, R.; Picone, D.; Tancredi, T.; Temussi, P.A. δ Opioidmimetic antagonists: Prototypes for designing a new generation of ultraselective opioid peptides. Mol. Med. 1995, 1, 678–689. [Google Scholar] [CrossRef]
  180. Lee, Y.S.; Qu, H.C.; Davis, P.; Ma, S.W.; Vardanyan, R.; Lai, J.; Porreca, F.; Hruby, V.J. Chiral Effect of a Phe Residue in Position 3 of the Dmt(1)-L(or D)-Tic(2) Analogues on Opioid Functional Activities. ACS Med. Chem. Lett. 2013, 4, 656–659. [Google Scholar] [CrossRef]
  181. Van den Eynde, I.; Laus, G.; Schiller, P.W.; Kosson, P.; Chung, N.N.; Lipkowski, A.W.; Tourwé, D. A new structural motif for μ-opioid antagonists. J. Med. Chem. 2005, 48, 3644–3648. [Google Scholar] [CrossRef]
  182. Aldrich, J.V.; McLaughlin, J.P. Peptide kappa opioid receptor ligands: Potential for drug development. AAPS J. 2009, 11, 312–322. [Google Scholar] [CrossRef]
  183. Dalefield, M.L.; Scouller, B.; Bibi, R.; Kivell, B.M. The Kappa Opioid Receptor: A Promising Therapeutic Target for Multiple Pathologies. Front. Pharmacol. 2022, 13, 837671. [Google Scholar] [CrossRef]
  184. Aldrich, J.V.; McLaughlin, J.P. Peptide Kappa opioid receptor ligands and their potential for drug development. In The Kappa Opioid Receptor; Liu-Chen, L.Y., Inan, S., Eds.; Springer: Berlin, Germany, 2021; Volume 271. [Google Scholar]
  185. Yoshino, H.; Nakazawa, T.; Arakawa, Y.; Kaneko, T.; Tsuchiya, Y.; Matsunaga, M.; Araki, S.; Ikeda, M.; Yamatsu, K.; Tachibana, S. Synthesis and structure-activity relationships of dynorphin A-(1-8) amide analogs. J. Med. Chem. 1990, 33, 206–212. [Google Scholar] [CrossRef]
  186. Nakazawa, T.; Furuya, Y.; Kaneko, T.; Yamatsu, K.; Yoshino, H.; Tachibana, S. Analgesia produced by E-2078, a systemically active dynorphin analog, in mice. J. Pharmacol. Exp. Ther. 1990, 252, 1247–1254. [Google Scholar]
  187. Hiramatsu, M.; Inoue, K.; Ambo, A.; Sasaki, Y.; Kameyama, T. Long-lasting antinociceptive effects of a novel dynorphin analogue, Tyr-D-Ala-Phe-Leu-Arg ψ (CH2NH) Arg-NH2, in mice. Br. J. Pharmacol. 2001, 132, 1948–1956. [Google Scholar] [CrossRef]
  188. Ronsisvalle, G.; Pappalardo, M.; Carboni, L.; Vittorio, F.; Pasquinucci, L.; Marrazzo, A.; Cacciaguerra, S.; Spampinato, S. Peptidomimetics of the κ-opioid receptor. A hybrid MPCB/peptide ligand (MPCB-RRI) binds κ cloned receptor with nanomolar affinity. Analgesia 1996, 2, 283–286. [Google Scholar]
  189. Binder, W.; Machelska, H.; Mousa, S.; Schmitt, T.; Rivière, P.J.; Junien, J.-L.; Stein, C.; Schäfer, M. Analgesic and antiinflammatory effects of two novel κ-opioid peptides. Anesth. J. Am. Soc. Anesthesiol. 2001, 94, 1034–1044. [Google Scholar] [CrossRef]
  190. Vanderah, T.W.; Largent-Milnes, T.; Lai, J.; Porreca, F.; Houghten, R.A.; Menzaghi, F.; Wisniewski, K.; Stalewski, J.; Sueiras-Diaz, J.; Galyean, R.; et al. Novel D-amino acid tetrapeptides produce potent antinociception by selectively acting at peripheral kappa-opioid receptors. Eur. J. Pharmacol. 2008, 583, 62–72. [Google Scholar] [CrossRef]
  191. Arendt-Nielsen, L.; Olesen, A.E.; Staahl, C.; Menzaghi, F.; Kell, S.; Wong, G.Y.; Drewes, A.M. Analgesic Efficacy of Peripheral κ-Opioid Receptor Agonist CR665 Compared to Oxycodone in a Multi-modal, Multi-tissue Experimental Human Pain ModelSelective Effect on Visceral Pain. Anesth. J. Am. Soc. Anesthesiol. 2009, 111, 616–624. [Google Scholar] [CrossRef]
  192. Dooley, C.T.; Ny, P.; Bidlack, J.M.; Houghten, R.A. Selective ligands for the μ, δ, and κ opioid receptors identified from a single mixture based tetrapeptide positional scanning combinatorial library. J. Biol. Chem. 1998, 273, 18848–18856. [Google Scholar] [CrossRef]
  193. Choi, H.; Murray, T.F.; DeLander, G.E.; Schmidt, W.K.; Aldrich, J.V. Synthesis and opioid activity of [D-Pro10]dynorphin A-(1-11) analogues with N-terminal alkyl substitution. J. Med. Chem. 1997, 40, 2733–2739. [Google Scholar] [CrossRef]
  194. Patkar, K.A.; Murray, T.F.; Aldrich, J.V. The Effects of C-Terminal Modifications on the Opioid Activity of [N-BenzylTyr1] Dynorphin A-(1− 11) Analogues. J. Med. Chem. 2009, 52, 6814–6821. [Google Scholar] [CrossRef]
  195. Vig, B.S.; Murray, T.F.; Aldrich, J.V. Synthesis and opioid activity of side-chain-to-side-chain cyclic dynorphin A-(1− 11) amide analogues cyclized between positions 2 and 5. 1. Substitutions in position 3. J. Med. Chem. 2004, 47, 446–455. [Google Scholar] [CrossRef]
  196. Lu, Y.; Nguyen, T.M.; Weltrowska, G.; Berezowska, I.; Lemieux, C.; Chung, N.N.; Schiller, P.W. [2′,6′-Dimethyltyrosine]dynorphin A(1-11)-NH2 analogues lacking an N-terminal amino group: Potent and selective kappa opioid antagonists. J. Med. Chem. 2001, 44, 3048–3053. [Google Scholar] [CrossRef]
  197. Corbett, A.D.; Gillan, M.G.; Kosterlitz, H.W.; McKnight, A.T.; Paterson, S.J.; Robson, L.E. Selectivities of opioid peptide analogues as agonists and antagonists at the delta-receptor. Br. J. Pharmacol. 1984, 83, 271–279. [Google Scholar] [CrossRef]
  198. Kreil, G.; Barra, D.; Simmaco, M.; Erspamer, V.; Erspamer, G.F.; Negri, L.; Severini, C.; Corsi, R.; Melchiorri, P. Deltorphin, a novel amphibian skin peptide with high selectivity and affinity for δ opioid receptors. Eur. J. Pharmacol. 1989, 162, 123–128. [Google Scholar] [CrossRef]
  199. McFadyen, I.; Ho, J.; Mosberg, H.I.; Traynor, J.R. Modifications of the cyclic mu receptor selective tetrapeptide Tyr-c [d-Cys-Phe-d-Pen] NH2 (Et): Effects on opioid receptor binding and activation. J. Peptide Res. 2000, 55, 255–261. [Google Scholar] [CrossRef]
  200. Yaksh, T.L.; Malmberg, A.B.; Ro, S.; Schiller, P.; Goodman, M. Characterization of the spinal antinociceptive activity of constrained peptidomimetic opioids. J. Pharmacol. Exp. Ther. 1995, 275, 63–72. [Google Scholar]
  201. Brantl, V.; Pfeiffer, A.; Herz, A.; Henschen, A.; Lottspeich, F. Antinociceptive potencies of β-casomorphin analogs as compared to their affinities towards μ and δ opiate receptor sites in brain and periphery. Peptides 1982, 3, 793–797. [Google Scholar] [CrossRef]
  202. Weber, E.; Esch, F.; Böhlen, P.; Paterson, S.; Corbett, A.; McKnight, A.; Kosterlitz, H.; Barchas, J.; Evans, C. Metorphamide: Isolation, structure, and biologic activity of an amidated opioid octapeptide from bovine brain. Proc. Natl. Acad. Sci. USA 1983, 80, 7362–7366. [Google Scholar] [CrossRef]
  203. Hruby, V.J.; Bartosz-Bechowski, H.; Davis, P.; Slaninova, J.; Zalewska, T.; Stropova, D.; Porreca, F.; Yamamura, H.I. Cyclic enkephalin analogues with exceptional potency and selectivity for δ-opioid receptors. J. Med. Chem. 1997, 40, 3957–3962. [Google Scholar] [CrossRef]
  204. Mosberg, H.I.; Lomize, A.L.; Wang, C.; Kroona, H.; Heyl, D.L.; Sobczyk-Kojiro, K.; Ma, W.; Mousigian, C.; Porreca, F. Development of a Model for the. delta. Opioid Receptor Pharmacophore. 1. Conformationally Restricted Tyr1 Replacements in the Cyclic. delta. Receptor Selective Tetrapeptide Tyr-c [D-Cys-Phe-D-Pen] OH (JOM-13). J. Med. Chem. 1994, 37, 4371–4383. [Google Scholar] [CrossRef]
  205. Gacel, G.; Fournie-Zaluski, M.-C.; Roques, B.P. D-Tyr—Ser—Gly—Phe—Leu—Thr, a highly preferential ligand for δ-opiate receptors. FEBS Lett. 1980, 118, 245–247. [Google Scholar] [CrossRef]
  206. Raynor, K.; Kong, H.; Chen, Y.; Yasuda, K.; Yu, L.; Bell, G.I.; Reisine, T. Pharmacological characterization of the cloned kappa-, delta-, and mu-opioid receptors. Mol. Pharmacol. 1994, 45, 330–334. [Google Scholar]
  207. Sanchez-Blazquez, P.; Garzon, J.; Lee, N. [Leu5] enkephalin containing peptides derived from a common precursor: Evaluation of opioid activity in in vitro bioassays. Eur. J. Pharmacol. 1984, 98, 389–396. [Google Scholar] [CrossRef]
  208. Lipkowski, A.W.; Misicka, A.; Davis, P.; Stropova, D.; Janders, J.; Lachwa, M.; Porreca, F.; Yamamura, H.I.; Hruby, V.J. Biological activity of fragments and analogues of the potent dimeric opioid peptide, biphalin. Bioorg. Med. Chem. Lett. 1999, 9, 2763–2766. [Google Scholar] [CrossRef]
  209. Kawasaki, A.M.; Knapp, R.J.; Walton, A.; Wire, W.S.; Zalewska, T.; Yamamura, H.I.; Porreca, F.; Burks, T.F.; Hruby, V.J. Syntheses, opioid binding affinities, and potencies of dynorphin A analogues substituted in positions 1, 6, 7, 8 and 10. Int. J. Peptide Protein Res. 1993, 42, 411–419. [Google Scholar] [CrossRef]
  210. Schlechtingen, G.; Zhang, L.; Maycock, A.; DeHaven, R.N.; Daubert, J.D.; Cassel, J.; Chung, N.N.; Schiller, P.W.; Goodman, M. [Pro3] Dyn A (1−11)-NH2: A Dynorphin Analogue with High Selectivity for the κ Opioid Receptor. J. Med. Chem. 2000, 43, 2698–2702. [Google Scholar] [CrossRef]
  211. Albert-Vartanian, A.; Boyd, M.R.; Hall, A.L.; Morgado, S.J.; Nguyen, E.; Nguyen, V.P.H.; Patel, S.P.; Russo, L.J.; Shao, A.J.; Raffa, R.B. Will peripherally restricted kappa-opioid receptor agonists (pKORA s) relieve pain with less opioid adverse effects and abuse potential? J. Clin. Pharm. Ther. 2016, 41, 371–382. [Google Scholar] [CrossRef]
  212. Bennett, M.A.; Murray, T.F.; Aldrich, J.V. Identification of arodyn, a novel acetylated dynorphin A-(1− 11) analogue, as a κ opioid receptor antagonist. J. Med. Chem. 2002, 45, 5617–5619. [Google Scholar] [CrossRef]
  213. Patkar, K.A.; Yan, X.; Murray, T.F.; Aldrich, J.V. [N α-BenzylTyr, c yclo (d-Asp5, Dap8)]-dynorphin A-(1− 11) NH2 Cyclized in the “Address” Domain Is a Novel κ-Opioid Receptor Antagonist. J. Med. Chem. 2005, 48, 4500–4503. [Google Scholar] [CrossRef] [PubMed]
  214. Zuckermann, R.N.; Martin, E.J.; Spellmeyer, D.C.; Stauber, G.B.; Shoemaker, K.R.; Kerr, J.M.; Figliozzi, G.M.; Goff, D.A.; Siani, M.A.; Simon, R.J.; et al. Discovery of nanomolar ligands for 7-transmembrane G-protein-coupled receptors from a diverse N-(substituted)glycine peptoid library. J. Med. Chem. 1994, 37, 2678–2685. [Google Scholar] [CrossRef] [PubMed]
  215. Simon, R.J.; Kania, R.S.; Zuckermann, R.N.; Huebner, V.D.; Jewell, D.A.; Banville, S.; Ng, S.; Wang, L.; Rosenberg, S.; Marlowe, C.K.; et al. Peptoids: A modular approach to drug discovery. Proc. Natl. Acad. Sci. USA 1992, 89, 9367–9371. [Google Scholar] [CrossRef] [PubMed]
  216. Biondi, L.; Giannini, E.; Filira, F.; Gobbo, M.; Marastoni, M.; Negri, L.; Scolaro, B.; Tomatisc, R.; Rocchi, R. Synthesis, conformation and biological activity of dermorphin and deltorphin I analogues containing N-alkylglycine in place of residues in position 1, 3, 5 and 6. J. Pept. Sci. 2003, 9, 638–648. [Google Scholar] [CrossRef] [PubMed]
  217. Wilczyńska, D.; Kosson, P.; Kwasiborska, M.; Ejchart, A.; Olma, A. Synthesis and receptor binding of opioid peptide analogues containing β3-homo-amino acids. J. Peptide Sci. 2009, 15, 777–782. [Google Scholar] [CrossRef] [PubMed]
  218. Chorev, M. The partial retro–inverso modification: A road traveled together. Pept. Sci. Orig. Res. Biomol. 2005, 80, 67–84. [Google Scholar] [CrossRef] [PubMed]
  219. Salvadori, S.; Marastoni, M.; Balboni, G.; Sarto, G.P.; Tomatis, R. Synthesis and pharmacological activity of partially modified retro-inverso dermorphin tetrapeptides. J. Med. Chem. 1985, 28, 769–774. [Google Scholar] [CrossRef]
  220. Remesic, M.; Lee, Y.S.; Hruby, V.J. Cyclic Opioid Peptides. Curr. Med. Chem. 2016, 23, 1288–1303. [Google Scholar] [CrossRef]
  221. Piekielna, J.; Perlikowska, R.; Gach, K.; Janecka, A. Cyclization in opioid peptides. Curr Drug Targets 2013, 14, 798–816. [Google Scholar] [CrossRef]
  222. Pencheva, N.; Milanov, P.; Vezenkov, L.; Pajpanova, T.; Naydenova, E. Opioid profiles of Cys2-containing enkephalin analogues. Eur. J. Pharmacol. 2004, 498, 249–256. [Google Scholar] [CrossRef]
  223. Rew, Y.; Malkmus, S.; Svensson, C.; Yaksh, T.L.; Chung, N.N.; Schiller, P.W.; Cassel, J.A.; DeHaven, R.N.; Taulane, J.P.; Goodman, M. Synthesis and biological activities of cyclic lanthionine enkephalin analogues: Delta-opioid receptor selective ligands. J. Med. Chem. 2002, 45, 3746–3754. [Google Scholar] [CrossRef]
  224. Mollica, A.; Guardiani, G.; Davis, P.; Ma, S.W.; Porreca, F.; Lai, J.; Mannina, L.; Sobolev, A.P.; Hruby, V.J. Synthesis of stable and potent delta/mu opioid peptides: Analogues of H-Tyr-c[D-Cys-Gly-Phe-D-Cys]-OH by ring-closing metathesis. J. Med. Chem. 2007, 50, 3138–3142. [Google Scholar] [CrossRef]
  225. Berezowska, I.; Chung, N.N.; Lemieux, C.; Wilkes, B.C.; Schiller, P.W. Dicarba analogues of the cyclic enkephalin peptides H-Tyr-c[D-Cys-Gly-Phe-D(or L)-Cys]NH(2) retain high opioid activity. J. Med. Chem. 2007, 50, 1414–1417. [Google Scholar] [CrossRef]
  226. Pawlak, D.; Oleszczuk, M.; Wójcik, J.; Pachulska, M.; Chung, N.N.; Schiller, P.W.; Izdebski, J. Highly potent side-chain to side-chain cyclized enkephalin analogues containing a carbonyl bridge: Synthesis, biology and conformation. J. Peptide Sci. Off. Publ. Eur. Peptide Soc. 2001, 7, 128–140. [Google Scholar] [CrossRef]
  227. Piekielna, J.; Kluczyk, A.; Gentilucci, L.; Cerlesi, M.C.; Tomböly, C.; Łapiński, K.; Janecki, T.; Janecka, A. Ring size in cyclic endomorphin-2 analogs modulates receptor binding affinity and selectivity. Organ. Biomol. Chem. 2015, 13, 6039–6046. [Google Scholar] [CrossRef]
  228. Zadina, J.E.; Nilges, M.R.; Morgenweck, J.; Zhang, X.; Hackler, L.; Fasold, M.B. Endomorphin analog analgesics with reduced abuse liability, respiratory depression, motor impairment, tolerance, and glial activation relative to morphine. Neuropharmacology 2016, 105, 215–227. [Google Scholar] [CrossRef]
  229. Purington, L.C.; Pogozheva, I.D.; Traynor, J.R.; Mosberg, H.I. Pentapeptides displaying μ opioid receptor agonist and δ opioid receptor partial agonist/antagonist properties. J. Med. Chem. 2009, 52, 7724–7731. [Google Scholar] [CrossRef]
  230. Fichna, J.; Perlikowska, R.; Wyrebska, A.; Gach, K.; Piekielna, J.; do-Rego, J.C.; Toth, G.; Kluczyk, A.; Janecki, T.; Janecka, A. Effect of 2′,6′-dimethyl-L-tyrosine (Dmt) on pharmacological activity of cyclic endomorphin-2 and morphiceptin analogs. Bioorg. Med. Chem. 2011, 19, 6977–6981. [Google Scholar] [CrossRef]
  231. Zielińska, M.; Chen, C.; Mokrowiecka, A.; Cygankiewicz, A.I.; Zakrzewski, P.K.; Sałaga, M.; Małecka-Panas, E.; Wlaź, P.; Krajewska, W.M.; Fichna, J. Orally administered novel cyclic pentapeptide P-317 alleviates symptoms of diarrhoea-predominant irritable bowel syndrome. J. Pharm. Pharmacol. 2015, 67, 244–254. [Google Scholar] [CrossRef]
  232. Burden, J.E.; Davis, P.; Porreca, F.; Spatola, A.F. Synthesis and biological activities of YkFA analogues: Effects of position 4 substitutions and altered ring size on in vitro opioid activity. Bioorg. Med. Chem. Lett. 2002, 12, 213–216. [Google Scholar] [CrossRef]
  233. Shreder, K.; Zhang, L.; Dang, T.; Yaksh, T.L.; Umeno, H.; DeHaven, R.; Daubert, J.; Goodman, M. Synthesis and biological activity of a novel methylamine-bridged enkephalin analogue (MABE): A new route to cyclic peptides and peptidomimetics. J. Med. Chem. 1998, 41, 2631–2635. [Google Scholar] [CrossRef]
  234. Kotlinska, J.; Bochenski, M.; Lagowska-Lenard, M.; Gibula-Bruzda, E.; Witkowska, E.; Izdebski, J. Enkephalin derivative, cyclo [Nε, Nβ-carbonyl-D-Lys2, Dap5] enkephalinamide (CUENK6), induces a highly potent antinociception in rats. Neuropeptides 2009, 43, 221–228. [Google Scholar] [CrossRef]
  235. Stefanucci, A.; Carotenuto, A.; Macedonio, G.; Novellino, E.; Pieretti, S.; Marzoli, F.; Szucs, E.; Erdei, A.I.; Zador, F.; Benyhe, S.; et al. Cyclic Biphalin Analogues Incorporating a Xylene Bridge: Synthesis, Characterization, and Biological Profile. ACS Med. Chem. Lett. 2017, 8, 858–863. [Google Scholar] [CrossRef]
  236. Mollica, A.; Davis, P.; Ma, S.-W.; Porreca, F.; Lai, J.; Hruby, V.J. Synthesis and biological activity of the first cyclic biphalin analogues. Bioorg. Med. Chem. letters 2006, 16, 367–372. [Google Scholar] [CrossRef]
  237. Remesic, M.; Macedonio, G.; Mollica, A.; Porreca, F.; Hruby, V.; Lee, Y.S. Cyclic biphalin analogues with a novel linker lead to potent agonist activities at mu, delta, and kappa opioid receptors. Bioorg. Med. Chem. 2018, 26, 3664–3667. [Google Scholar] [CrossRef]
  238. Vig, B.S.; Murray, T.F.; Aldrich, J.V. A novel N-terminal cyclic dynorphin A analogue cyclo(N,5)[Trp(3),Trp(4),Glu(5)] dynorphin A-(1-11)NH(2) that lacks the basic N-terminus. J. Med. Chem. 2003, 46, 1279–1282. [Google Scholar] [CrossRef]
  239. Fang, W.J.; Cui, Y.; Murray, T.F.; Aldrich, J.V. Design, synthesis, and pharmacological activities of dynorphin A analogues cyclized by ring-closing metathesis. J. Med. Chem. 2009, 52, 5619–5625. [Google Scholar] [CrossRef]
  240. Adamska-Bartłomiejczyk, A.; Lipiński, P.F.; Piekielna-Ciesielska, J.; Kluczyk, A.; Janecka, A. Pharmacological profile and molecular modeling of cyclic opioid analogs incorporating various phenylalanine derivatives. ChemMedChem 2020, 15, 1322–1329. [Google Scholar] [CrossRef]
  241. Przydzial, M.; Pogozheva, I.; Ho, J.; Bosse, K.; Sawyer, E.; Traynor, J.R.; Mosberg, H.I. Design of high affinity cyclic pentapeptide ligands for κ-opioid receptors. J. Peptide Res. 2005, 66, 255–262. [Google Scholar] [CrossRef]
  242. Piekielna, J.; Gentilucci, L.; De Marco, R.; Perlikowska, R.; Adamska, A.; Olczak, J.; Mazur, M.; Artali, R.; Modranka, J.; Janecki, T. Cyclic side-chain-linked opioid analogs utilizing cis-and trans-4-aminocyclohexyl-D-alanine. Bioorg. Med. Chem. 2014, 22, 6545–6551. [Google Scholar] [CrossRef]
  243. Touati-Jallabe, Y.; Bojnik, E.; Legrand, B.; Mauchauffée, E.; Chung, N.N.; Schiller, P.W.; Benyhe, S.; Averlant-Petit, M.-C.; Martinez, J.; Hernandez, J.-F.o. Cyclic enkephalins with a diversely substituted guanidine bridge or a thiourea bridge: Synthesis, biological and structural evaluations. J. Med. Chem. 2013, 56, 5964–5973. [Google Scholar] [CrossRef]
  244. Lu, Y.; Lum, T.K.; Leow Augustine, Y.W.; Weltrowska, G.; Nguyen, T.M.-D.; Lemieux, C.; Chung, N.N.; Schiller, P.W. Replacement of the N-terminal tyrosine residue in opioid peptides with 3-(2, 6-dimethyl-4-carbamoylphenyl) propanoic acid (Dcp) results in novel opioid antagonists. J. Med. Chem. 2006, 49, 5382–5385. [Google Scholar] [CrossRef]
  245. Lee, Y.S.; Petrov, R.; Park, C.K.; Ma, S.-w.; Davis, P.; Lai, J.; Porreca, F.; Vardanyan, R.; Hruby, V.J. Development of novel enkephalin analogues that have enhanced opioid activities at both μ and δ opioid receptors. J. Med. Chem. 2007, 50, 5528–5532. [Google Scholar] [CrossRef]
  246. Wiszniewska, A.; Kunce, D.; Chung, N.N.; Schiller, P.W.; Izdebski, J. p-Nitrophenoxycarbonyl derivatives of Boc-protected diaminoalkanes in the synthesis of enkephalin peptidomimetics. J. Pept. Sci. 2005, 11, 579–583. [Google Scholar] [CrossRef]
  247. Sharma, K.K.; Cassell, R.J.; Meqbil, Y.J.; Su, H.; Blaine, A.T.; Cummins, B.R.; Mores, K.L.; Johnson, D.K.; van Rijn, R.M.; Altman, R.A. Modulating β-arrestin 2 recruitment at the δ-and μ-opioid receptors using peptidomimetic ligands. RSC Med. Chem. 2021, 12, 1958–1967. [Google Scholar] [CrossRef]
  248. Bankowski, K.; Michalak, O.M.; Lesniak, A.; Filip, K.E.; Cmoch, P.; Szewczuk, Z.; Stefanowicz, P.; Izdebski, J. N-terminal guanidinylation of the cyclic 1,4-ureido-deltorphin analogues: The synthesis, receptor binding studies, and resistance to proteolytic digestion. J. Pept. Sci. 2015, 21, 467–475. [Google Scholar] [CrossRef]
  249. Hazum, E.; Chang, K.-J.; Leighton, H.; Lever, O.W., Jr.; Cuatrecasas, P. Increased biological activity of dimers of oxymorphone and enkephalin: Possible role of receptor crosslinking. Biochem. Biophys. Res. Commun. 1982, 104, 347–353. [Google Scholar] [CrossRef]
  250. Dyniewicz, J.; Lipinski, P.F.; Kosson, P.; Lesniak, A.; Bochynska-Czyz, M.; Muchowska, A.; Tourwe, D.; Ballet, S.; Misicka, A.; Lipkowski, A.W. Hydrazone Linker as a Useful Tool for Preparing Chimeric Peptide/Nonpeptide Bifunctional Compounds. ACS Med. Chem. Lett. 2017, 8, 73–77. [Google Scholar] [CrossRef]
  251. Weltrowska, G.; Lemieux, C.; Chung, N.; Schiller, P. A chimeric opioid peptide with mixed μ agonist/δ antagonist properties. J. Peptide Res. 2004, 63, 63–68. [Google Scholar] [CrossRef]
  252. Gomes, I.; IJzerman, A.P.; Ye, K.; Maillet, E.L.; Devi, L.A. G protein-coupled receptor heteromerization: A role in allosteric modulation of ligand binding. Mol. Pharmacol. 2011, 79, 1044–1052. [Google Scholar] [CrossRef]
  253. Wiszniewska, A.; Kunce, D.; Chung, N.N.; Schiller, P.W.; Izdebski, J. Synthesis of peptidomimetics: An evaluation of p-nitrophenyl carbamate of ethylenediamine. Lett. Peptide Sci. 2003, 10, 33–39. [Google Scholar] [CrossRef]
  254. Sinisi, R.; Ghilardi, A.; Ruiu, S.; Lazzari, P.; Malpezzi, L.; Sani, M.; Pani, L.; Zanda, M. Synthesis and in vitro evaluation of trifluoroethylamine analogues of enkephalins. ChemMedChem 2009, 4, 1416–1420. [Google Scholar] [CrossRef] [PubMed]
  255. Blomberg, D.; Kreye, P.; Fowler, C.; Brickmann, K.; Kihlberg, J. Synthesis and biological evaluation of leucine enkephalin turn mimetics. Organ. Biomol. Chem. 2006, 4, 416–423. [Google Scholar] [CrossRef] [PubMed]
  256. Altman, R.A.; Sharma, K.K.; Rajewski, L.G.; Toren, P.C.; Baltezor, M.J.; Pal, M.; Karad, S.N. Tyr1-ψ [(Z) CF═ CH]-Gly2 Fluorinated Peptidomimetic Improves Distribution and Metabolism Properties of Leu-Enkephalin. ACS Chem. Neurosci. 2018, 9, 1735–1742. [Google Scholar] [CrossRef]
  257. Chingle, R.; Mulumba, M.; Chung, N.N.; Nguyen, T.M.-D.; Ong, H.; Ballet, S.; Schiller, P.W.; Lubell, W.D. Solid-Phase Azopeptide Diels–Alder Chemistry for Aza-pipecolyl Residue Synthesis To Study Peptide Conformation. J. Organ. Chem. 2019, 84, 6006–6016. [Google Scholar] [CrossRef]
  258. Cabrele, C.; Martinek, T.s.A.; Reiser, O.; Berlicki, Ł. Peptides containing β-amino acid patterns: Challenges and successes in medicinal chemistry. J. Med. Chem. 2014, 57, 9718–9739. [Google Scholar] [CrossRef]
  259. Cardillo, G.; Gentilucci, L.; Qasem, A.R.; Sgarzi, F.; Spampinato, S. Endomorphin-1 analogues containing beta-proline are mu-opioid receptor agonists and display enhanced enzymatic hydrolysis resistance. J. Med. Chem. 2002, 45, 2571–2578. [Google Scholar] [CrossRef]
  260. Cardillo, G.; Gentilucci, L.; Melchiorre, P.; Spampinato, S. Synthesis and binding activity of endomorphin-1 analogues containing beta-amino acids. Bioorg. Med. Chem. Lett. 2000, 10, 2755–2758. [Google Scholar] [CrossRef]
  261. Lesma, G.; Salvadori, S.; Airaghi, F.; Bojnik, E.; Borsodi, A.; Recca, T.; Sacchetti, A.; Balboni, G.; Silvani, A. Synthesis, pharmacological evaluation and conformational investigation of endomorphin-2 hybrid analogues. Mol. Divers 2013, 17, 19–31. [Google Scholar] [CrossRef]
  262. Lesma, G.; Salvadori, S.; Airaghi, F.; Murray, T.F.; Recca, T.; Sacchetti, A.; Balboni, G.; Silvani, A. Structural and Biological Exploration of Phe3–Phe4-Modified Endomorphin-2 Peptidomimetics. Med. Chem. Lett. 2013, 4, 795–799. [Google Scholar] [CrossRef]
  263. Wang, Y.; Xing, Y.; Liu, X.; Ji, H.; Kai, M.; Chen, Z.; Yu, J.; Zhao, D.; Ren, H.; Wang, R. A new class of highly potent and selective endomorphin-1 analogues containing alpha-methylene-beta-aminopropanoic acids (map). J. Med. Chem. 2012, 55, 6224–6236. [Google Scholar] [CrossRef]
  264. Hu, M.; Giulianotti, M.A.; McLaughlin, J.P.; Shao, J.; Debevec, G.; Maida, L.E.; Geer, P.; Cazares, M.; Misler, J.; Li, L.; et al. Synthesis and biological evaluations of novel endomorphin analogues containing alpha-hydroxy-beta-phenylalanine (AHPBA) displaying mixed mu/delta opioid receptor agonist and delta opioid receptor antagonist activities. Eur. J. Med. Chem. 2015, 92, 270–281. [Google Scholar] [CrossRef]
  265. Bozu, B.; Fulop, F.; Toth, G.K.; Toth, G.; Szucs, M. Synthesis and opioid binding activity of dermorphin analogues containing cyclic beta-amino acids. Neuropeptides 1997, 31, 367–372. [Google Scholar] [CrossRef]
  266. Janecka, A.; Kruszynski, R. Conformationally restricted peptides as tools in opioid receptor studies. Curr. Med. Chem. 2005, 12, 471–481. [Google Scholar] [CrossRef]
  267. Bryant, S.D.; Jinsmaa, Y.; Salvadori, S.; Okada, Y.; Lazarus, L.H. Dmt and opioid peptides: A potent alliance. Peptide Sci. 2003, 71, 86–102. [Google Scholar] [CrossRef]
  268. Montgomery, D.; Anand, J.P.; Griggs, N.W.; Fernandez, T.J.; Hartman, J.G.; Sánchez-Santiago, A.A.; Pogozheva, I.D.; Traynor, J.R.; Mosberg, H.I. Novel Dimethyltyrosine–Tetrahydroisoquinoline Peptidomimetics with Aromatic Tetrahydroisoquinoline Substitutions Show in Vitro Kappa and Mu Opioid Receptor Agonism. ACS Chem. Neurosci. 2019, 10, 3682–3689. [Google Scholar] [CrossRef]
  269. Wang, C.; McFadyen, I.J.; Traynor, J.R.; Mosberg, H.I. Design of a high affinity peptidomimetic opioid agonist from peptide pharmacophore models. Bioorg. Med. Chem. Lett. 1998, 8, 2685–2688. [Google Scholar] [CrossRef]
  270. Nastase, A.F.; Griggs, N.W.; Anand, J.P.; Fernandez, T.J.; Harland, A.A.; Trask, T.J.; Jutkiewicz, E.M.; Traynor, J.R.; Mosberg, H.I. Synthesis and Pharmacological Evaluation of Novel C-8 Substituted Tetrahydroquinolines as Balanced-Affinity Mu/Delta Opioid Ligands for the Treatment of Pain. ACS Chem. Neurosci. 2018, 9, 1840–1848. [Google Scholar] [CrossRef]
  271. Bender, A.M.; Griggs, N.W.; Anand, J.P.; Traynor, J.R.; Jutkiewicz, E.M.; Mosberg, H.I. Asymmetric synthesis and in vitro and in vivo activity of tetrahydroquinolines featuring a diverse set of polar substitutions at the 6 position as mixed-efficacy mu opioid receptor/delta opioid receptor ligands. ACS Chem. Neurosci. 2015, 6, 1428–1435. [Google Scholar] [CrossRef]
  272. Harland, A.A.; Pogozheva, I.D.; Griggs, N.W.; Trask, T.J.; Traynor, J.R.; Mosberg, H.I. Placement of Hydroxy Moiety on Pendant of Peptidomimetic Scaffold Modulates Mu and Kappa Opioid Receptor Efficacy. ACS Chem. Neurosci. 2017, 8, 2549–2557. [Google Scholar] [CrossRef]
  273. Torino, D.; Mollica, A.; Pinnen, F.; Feliciani, F.; Lucente, G.; Fabrizi, G.; Portalone, G.; Davis, P.; Lai, J.; Ma, S.W.; et al. Synthesis and evaluation of new endomorphin-2 analogues containing (Z)-alpha,beta-didehydrophenylalanine (Delta(Z)Phe) residues. J. Med. Chem. 2010, 53, 4550–4554. [Google Scholar] [CrossRef]
  274. Siodłak, D. α, β-Dehydroamino acids in naturally occurring peptides. Amino Acids 2015, 47, 1–17. [Google Scholar] [CrossRef]
  275. Rinnova, M.; Nefzi, A.; Houghten, R.A. Opioid activity of 4-imidazolidinone positional analogues of Leu-Enkephalin. Bioorg. Med. Chem. Lett. 2002, 12, 3175–3178. [Google Scholar] [CrossRef]
  276. Shiotani, K.; Li, T.; Miyazaki, A.; Tsuda, Y.; Yokoi, T.; Ambo, A.; Sasaki, Y.; Bryant, S.D.; Lazarus, L.H.; Okada, Y. Design and synthesis of opioidmimetics containing 2’,6’-dimethyl-L-tyrosine and a pyrazinone-ring platform. Bioorg. Med. Chem. Lett. 2007, 17, 5768–5771. [Google Scholar] [CrossRef]
  277. Okada, Y.; Tsukatani, M.; Taguchi, H.; Yokoi, T.; Bryant, S.D.; Lazarus, L.H. Amino acids and peptides. LII. Design and synthesis of opioid mimetics containing a pyrazinone ring and examination of their opioid receptor binding activity. Chem. Pharm. Bull. 1998, 46, 1374–1382. [Google Scholar] [CrossRef]
  278. Proteau-Gagné, A.; Rochon, K.; Roy, M.; Albert, P.-J.; Guérin, B.; Gendron, L.; Dory, Y.L. Systematic replacement of amides by 1, 4-disubstituted [1-3] triazoles in Leu-enkephalin and the impact on the delta opioid receptor activity. Bioorg. Med. Chem. Lett. 2013, 23, 5267–5269. [Google Scholar] [CrossRef] [Green Version]
  279. Olczak, J.; Kaczmarek, K.; Maszczyńska, I.; Lisowski, M.; Stropova, D.; Hruby, V.J.; Yamamura, H.I.; Lipkowski, A.W.; Zabrocki, J. Consequences of cis-amide bond simulation in opioid peptides. Lett. Peptide Sci. 1998, 5, 437–440. [Google Scholar] [CrossRef]
  280. Martin, S.F.; Dwyer, M.P.; Hartmann, B.; Knight, K.S. Cyclopropane-derived peptidomimetics. Design, synthesis, and evaluation of novel enkephalin analogues. J. Org. Chem. 2000, 65, 1305–1318. [Google Scholar] [CrossRef] [PubMed]
  281. Yamashita, T.; Tsuru, E.; Banjyo, E.; Doe, M.; Shibata, K.; Yasuda, M.; Gemba, M. Synthesis and opiate activity of pseudo-tetrapeptides containing chiral piperazin-2-one and piperazine derivatives. Chem. Pharm. Bull. 1997, 45, 1940–1944. [Google Scholar] [CrossRef] [PubMed]
  282. Gu, X.; Ying, J.; Min, B.; Cain, J.P.; Davis, P.; Willey, P.; Navratilova, E.; Yamamura, H.I.; Porreca, F.; Hruby, V.J. Parallel synthesis and biological evaluation of different sizes of bicyclo [2,3]-Leu-enkephalin analogues. Biopolymers 2005, 80, 151–163. [Google Scholar] [CrossRef] [PubMed]
  283. Ballet, S.; Salvadori, S.; Trapella, C.; Bryant, S.D.; Jinsmaa, Y.; Lazarus, L.H.; Negri, L.; Giannini, E.; Lattanzi, R.; Tourwe, D.; et al. New 2′,6′-dimethyl-L-tyrosine (Dmt) opioid peptidomimetics based on the Aba-Gly scaffold. Development of unique mu-opioid receptor ligands. J. Med. Chem. 2006, 49, 3990–3993. [Google Scholar] [CrossRef]
  284. Ballet, S.; Guillemyn, K.; Van der Poorten, O.; Schurgers, B.; Verniest, G.; Tourwé, D. Azepinone-Constrained Amino Acids in Peptide and Peptidomimetic Design. In Peptidomimetics I; Springer: Amsterdam, The Netherlands, 2015; pp. 177–209. [Google Scholar]
  285. Ballet, S.; Feytens, D.; De Wachter, R.; De Vlaeminck, M.; Marczak, E.D.; Salvadori, S.; de Graaf, C.; Rognan, D.; Negri, L.; Lattanzi, R. Conformationally constrained opioid ligands: The Dmt-Aba and Dmt-Aia versus Dmt-Tic scaffold. Bioorg. Med. Chem. Lett. 2009, 19, 433–437. [Google Scholar] [CrossRef]
  286. Ballet, S.; Marczak, E.D.; Feytens, D.; Salvadori, S.; Sasaki, Y.; Abell, A.D.; Lazarus, L.H.; Balboni, G.; Tourwé, D. Novel multiple opioid ligands based on 4-aminobenzazepinone (Aba), azepinoindole (Aia) and tetrahydroisoquinoline (Tic) scaffolds. Bioorg. Med. Chem. Lett. 2010, 20, 1610–1613. [Google Scholar] [CrossRef]
  287. Ballet, S.; Frycia, A.; Piron, J.; Chung, N.; Schiller, P.; Kosson, P.; Lipkowski, A.; Tourwé, D. Synthesis and biological evaluation of constrained analogues of the opioid peptide H-Tyr-d-Ala-Phe-Gly-NH2 using the 4-amino-2-benzazepin-3-one scaffold. J. Peptide Res. 2005, 66, 222–230. [Google Scholar] [CrossRef]
  288. Goodman, M.; Cai, W.; Smith, N.D. The bold legacy of Emil Fischer. J. Pept. Sci. 2003, 9, 594–603. [Google Scholar] [CrossRef]
  289. Schmidt, R.; Menard, D.; Mrestani-Klaus, C.; Chung, N.N.; Lemieux, C.; Schiller, P.W. Structural modifications of the N-terminal tetrapeptide segment of [D-Ala2] deltorphin I: Effects on opioid receptor affinities and activities in vitro and on antinociceptive potency. Peptides 1997, 18, 1615–1621. [Google Scholar] [CrossRef]
  290. Bryant, S.D.; Guerrini, R.; Salvadori, S.; Bianchi, C.; Tomatis, R.; Attila, M.; Lazarus, L.H. Helix-inducing α-aminoisobutyric acid in opioid mimetic deltorphin C analogues. J. Med. Chem. 1997, 40, 2579–2587. [Google Scholar] [CrossRef]
  291. Witt, K.A.; Slate, C.A.; Egleton, R.D.; Huber, J.D.; Yamamura, H.I.; Hruby, V.J.; Davis, T.P. Assessment of stereoselectivity of trimethylphenylalanine analogues of delta-opioid [D-Pen(2),D-Pen(5)]-enkephalin. J. Neurochem. 2000, 75, 424–435. [Google Scholar] [CrossRef]
  292. Bilsky, E.J.; Qian, X.; Hruby, V.J.; Porreca, F. Antinociceptive activity of [beta-methyl-2′, 6′-dimethyltyrosine(1)]-substituted cyclic [D-Pen(2), D-Pen(5)]Enkephalin and [D-Ala(2),Asp(4)]Deltorphin analogs. J. Pharmacol. Exp. Ther. 2000, 293, 151–158. [Google Scholar]
  293. Qian, X.; Shenderovich, M.D.; Kövér, K.E.; Davis, P.; Horváth, R.; Zalewska, T.; Yamamura, H.I.; Porreca, F.; Hruby, V.J. Probing the stereochemical requirements for receptor recognition of δ opioid agonists through topographic modifications in position 1. J. Am. Chem. Soc. 1996, 118, 7280–7290. [Google Scholar] [CrossRef]
  294. Bryan, W.M.; Callahan, J.F.; Codd, E.E.; Lemieux, C.; Moore, M.L.; Schiller, P.W.; Walker, R.F.; Huffman, W.F. Cyclic enkephalin analogues containing alpha-amino-beta-mercapto-beta, beta-pentamethylenepropionic acid at positions 2 or 5. J. Med. Chem. 1989, 32, 302–304. [Google Scholar] [CrossRef]
  295. Cai, Y.; Lu, D.; Chen, Z.; Ding, Y.; Chung, N.N.; Li, T.; Schiller, P.W. [Dmt1] DALDA analogues modified with tyrosine analogues at position 1. Bioorg. Med. Chem. Lett. 2016, 26, 3629–3631. [Google Scholar] [CrossRef]
  296. Shimoyama, M.; Shimoyama, N.; Zhao, G.M.; Schiller, P.W.; Szeto, H.H. Antinociceptive and respiratory effects of intrathecal H-Tyr-D-Arg-Phe-Lys-NH2 (DALDA) and [Dmt1] DALDA. J. Pharmacol. Exp. Ther. 2001, 297, 364–371. [Google Scholar]
  297. Neilan, C.L.; Nguyen, T.M.; Schiller, P.W.; Pasternak, G.W. Pharmacological characterization of the dermorphin analog [Dmt(1)]DALDA, a highly potent and selective mu-opioid peptide. Eur. J. Pharmacol. 2001, 419, 15–23. [Google Scholar] [CrossRef]
  298. Dumitrascuta, M.; Bermudez, M.; Ballet, S.; Wolber, G.; Spetea, M. Mechanistic Understanding of Peptide Analogues, DALDA,[Dmt1] DALDA, and KGOP01, Binding to the Mu Opioid Receptor. Molecules 2020, 25, 2087. [Google Scholar] [CrossRef]
  299. Mierke, D.F.; Said-Nejad, O.E.; Schiller, P.W.; Goodman, M. Enkephalin analogues containing β-naphthylalanine at the fourth position. Biopolym. Ori. Res. Biomol. 1990, 29, 179–196. [Google Scholar] [CrossRef]
  300. Schmidt, R.; Vogel, D.; Mrestani-Klaus, C.; Brandt, W.; Neubert, K.; Chung, N.N.; Lemieux, C.; Schiller, P.W. Cyclic. beta.-Casomorphin Analogs with Mixed. mu. Agonist/.delta. Antagonist Properties: Synthesis, Pharmacological Characterization, and Conformational Aspects. J. Med. Chem. 1994, 37, 1136–1144. [Google Scholar] [CrossRef]
  301. Yamazaki, T.; Said-Nejad, O.E.; Schiller, P.W.; Goodman, M. Conformational studies of stereoisomeric 14-membered cyclic enkephalin analogues containing 1-naphthylalanine at the fourth position: Chirality effect of leucine at the fifth position on biological activity and receptor selectivity. Biopolymers 1991, 31, 877–898. [Google Scholar] [CrossRef]
  302. Dolle, R.E.; Michaut, M.; Martinez-Teipel, B.; Belanger, S.; Graczyk, T.M.; DeHaven, R.N. Further studies of tyrosine surrogates in opioid receptor peptide ligands. Bioorg. Med. Chem. Lett. 2007, 17, 2656–2660. [Google Scholar] [CrossRef]
  303. Choi, H.; Murray, T.F.; Aldrich, J.V. Synthesis and evaluation of derivatives of leucine enkephalin as potential affinity labels for delta opioid receptors. Biopolymers 2003, 71, 552–557. [Google Scholar] [CrossRef]
  304. Choi, H.; Murray, T.F.; Aldrich, J.V. Dermorphin-based potential affinity labels for mu-opioid receptors. J. Pept. Res. 2003, 61, 40–45. [Google Scholar] [CrossRef] [PubMed]
  305. Choi, H.; Murray, T.F.; Aldrich, J.V. Synthesis and evaluation of potential affinity labels. derived from endomorphin-2. J. Pept. Res. 2003, 61, 58–62. [Google Scholar] [CrossRef] [PubMed]
  306. Sartania, N.; Szatmári, I.; Orosz, G.; Rónai, A.Z.; Medzihradszky, K.; Borsodi, A.; Benyhe, S. Irreversible labelling of the opioid receptors by a melphalan-substituted [Met5] enkephalin-Arg-Phe derivative. Eur. J. Pharmacol. 1999, 373, 241–249. [Google Scholar] [CrossRef]
  307. Dooley, C.; Chung, N.; Schiller, P.; Houghten, R. Acetalins: Opioid receptor antagonists determined through the use of synthetic peptide combinatorial libraries. Proc. Natl. Acad. Sci. USA 1993, 90, 10811–10815. [Google Scholar] [CrossRef]
  308. Dooley, C.T.; Houghten, R.A. New opioid peptides, peptidomimetics, and heterocyclic compounds from combinatorial libraries. Biopolymers 1999, 51, 379–390. [Google Scholar] [CrossRef]
Figure 1. Subtype selective opioid peptidomimetics developed from endogenous opioid peptides.
Figure 1. Subtype selective opioid peptidomimetics developed from endogenous opioid peptides.
Biomolecules 12 01241 g001
Figure 2. Various modifications to optimize metabolic stability, BBB permeability, and oral bioavailability.
Figure 2. Various modifications to optimize metabolic stability, BBB permeability, and oral bioavailability.
Biomolecules 12 01241 g002
Figure 3. Multifunctional opioid peptidomimetics: MOR/DOR agonist, MOR agonist/DOR antagonist.
Figure 3. Multifunctional opioid peptidomimetics: MOR/DOR agonist, MOR agonist/DOR antagonist.
Biomolecules 12 01241 g003
Figure 4. Message and address regions of natural opioid peptides.
Figure 4. Message and address regions of natural opioid peptides.
Biomolecules 12 01241 g004
Figure 6. Various tools for peptidomimetics.
Figure 6. Various tools for peptidomimetics.
Biomolecules 12 01241 g006
Figure 7. Structures of β-amino acids (upper) and alicyclic β-amino acids (lower). * Chiral.
Figure 7. Structures of β-amino acids (upper) and alicyclic β-amino acids (lower). * Chiral.
Biomolecules 12 01241 g007
Figure 8. Structures of constrained amino acids and dipeptidic scaffolds. * Chiral.
Figure 8. Structures of constrained amino acids and dipeptidic scaffolds. * Chiral.
Biomolecules 12 01241 g008
Figure 9. Cβ-substituted amino acids. * Chiral.
Figure 9. Cβ-substituted amino acids. * Chiral.
Biomolecules 12 01241 g009
Figure 10. Structures of Leucine derivatives.
Figure 10. Structures of Leucine derivatives.
Biomolecules 12 01241 g010
Table 1. Natural opioid peptides and their selectivities for the opioid receptors.
Table 1. Natural opioid peptides and their selectivities for the opioid receptors.
PeptidesStructureSelectivity
β-Endorphin (END)YGGFMTSEKSQTPLVTLFKNAIIKNAYKKGEMOR > DOR
Enkephalins (ENKs)YGGFLDOR > MOR
YGGFMDOR > MOR
YGGFMRFMOR > DOR > KOR
YGGFMRGLMOR > DOR > KOR
Dynorphin (DYN) AYGGFLRRIRPKLKWDNQKOR > MOR > DOR
DYN BYGGFLRRQFKVVTKOR > MOR > DOR
Endomorphin (EM)-1YPWF-NH2MOR
EM-2YPFF-NH2MOR
Dermorphines (DERs)YaFGYPS-NH2MOR
YaFGYPKMOR
YaFWYPNMOR
Deltorphine (DLT) AYmFHLMD-NH2DOR
DLT-1YaFDVVG-NH2DOR
DLT-2YaFEVVG-NH2DOR
Table 3. Binding affinities of potent cyclic opioid peptidomimetics at MOR, DOR, and KOR a.
Table 3. Binding affinities of potent cyclic opioid peptidomimetics at MOR, DOR, and KOR a.
StructureKi or IC50 (nM)Ref.
MORDORKOR
Tyr-c2,4[DLys-Phe-Asp]-NH20.21461684[227]
Tyr-c2,5[DLys-Phe-Phe-Asp]-NH20.351711.12[240]
Tyr-c2,5[DCys-Phe-Phe-DCys]-NH20.050.41.6[241]
Tyr-c2,4[DDap-Phe-Phe-Asp]-NH20.51>1000>1000[227]
Tyr-c2,5[cis-DACAla-Phe-Phe-Asp]-NH23.2>1000-[242]
Dmt-c2,5[cis-DACAla-Phe-Phe-Asp]-NH20.04>1000>1000[242]
Dmt-c[DLys-Phe-DPro-Asp]-NH2 (P-317)
Tyr-c2,5(-SCH2S-)[DCys-Phe-2-Nal-Cys]-NH20.470.481.3[229]
cN,C[Tyr-DPro-DTrp-Phe-Gly]34--[147]
Tyr-c2,5(-S-)[DVal-Gly-Phe-DAla]-OH6300.931600[223]
Tyr-c2,5(-S-)[DAla-Gly-Phe-DAla]-OH2.02.01600[223]
Tyr-c2,5(-CH2CH2-)[DAla-Gly-Phe-Ala]-NH22.35.9309[225]
Tyr c2,5(-cisCH=CH-) [DAla-Gly-Phe-DAla]-OH1.350.43576[224]
Tyr-c2,5(-NHCSNH-)[DDap-Gly-Phe-Dap]-NH20.45.4-[243]
Dcp-c2,5[DCys-Gly-Phe(4-NO2)-DCys]-NH22.84 b25.8980[244]
c2,2′(Tyr-DCys-Gly-Phe-NH)20.600.87-[236]
c2,2′(Dmt-DCys-Gly-Phe)2-cystamine0.270.360.87[237]
c2,5[DAsp2,DAla3,Dap5]-Dyn A (1–11)-NH23.891390.21[195]
c5,8(-transCH=CH-)[Ala5,Ala8]Dyn A (1–11)-NH236.04602.46[239]
cN,5[Trp3,Trp4,Glu5]-DYN A (1–11)-NH2331>890026.8[238]
a The variability in the binding data (Ki or IC50) between various laboratories may be attributed to differences in the synaptosomal preparations, the concentration, type of radioligand, and specific activity used, and the method of reporting binding. Refer to the citations for the details. b Antagonist.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, Y.S. Peptidomimetics and Their Applications for Opioid Peptide Drug Discovery. Biomolecules 2022, 12, 1241. https://doi.org/10.3390/biom12091241

AMA Style

Lee YS. Peptidomimetics and Their Applications for Opioid Peptide Drug Discovery. Biomolecules. 2022; 12(9):1241. https://doi.org/10.3390/biom12091241

Chicago/Turabian Style

Lee, Yeon Sun. 2022. "Peptidomimetics and Their Applications for Opioid Peptide Drug Discovery" Biomolecules 12, no. 9: 1241. https://doi.org/10.3390/biom12091241

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