• Open Access
    Original Article

    Late-stage diversification strategy for the synthesis of peptide acids and amides using hydrazides

    Shoko Tanaka 1
    Mizuki Kanno 1
    Yosuke Tashiro 1,2,3
    Tetsuo Narumi 1,2,3,4
    Nobuyuki Mase 1,2,3,4
    Kohei Sato 1,2,3,4*

    Explor Drug Sci. 2023;1:322–335 DOI: https://doi.org/10.37349/eds.2023.00023

    Received: June 29, 2023 Accepted: August 03, 2023 Published: October 09, 2023

    Academic Editor: Fernando Albericio, University of KwaZulu-Natal, South Africa; University of Barcelona, Spain

    This article belongs to the special issue Bioactive Peptides discovery and development

    Abstract

    Aim:

    Modification of the C-terminus of a peptide to improve its properties, particularly after constructing the peptide chain, has great promise in the development of peptide therapeutics. This study discusses the development of a late-stage diversification method for synthesizing peptide acids and amides from hydrazides which can serve as a common precursor.

    Methods:

    Peptide hydrazides were synthesized solely by using conventional solid-phase peptide synthesis (SPPS). Hydrazides were subjected to oxidation by potassium peroxymonosulfate (Oxone) to afford carboxylic acids. Azidation of hydrazides using sodium nitrite (NaNO2) under acidic conditions, followed by the addition of β-mercaptoethanol (BME), could also be used to generate carboxylic acids. For the preparation of peptide amides, azides that can be prepared from hydrazides were reacted with ammonium acetate (NH4OAc) or tris(2-carboxyethyl)phosphine (TCEP)∙hydrochloride (HCl) to develop the products through ammonolysis or a Staudinger reaction, which produces iminophosphorane from an azide and a phosphine. The antimicrobial activity of modelin-5 derivatives synthesized from the corresponding hydrazides was evaluated by the colony count of Escherichia coli (E. coli) after treatment with the peptides.

    Results:

    Oxone oxidation yielded the corresponding acids rapidly although oxidation-prone amino acids were incompatible. Azidation and subsequent treatment with BME afforded peptide acids an acceptable yield even in sequences containing amino acids that are prone to oxidation. Both methods for conversion of hydrazides to amides were found to afford the desired products in good yield and compatibility. The conditions that were developed were adapted to the synthesis of modelin-5 derivatives from the corresponding hydrazides, yielding late-stage production of the desired peptides. The amides of the resulting peptide showed more potent activity against E. coli than the acid form, and the most potent activity was observed from the hydrazide.

    Conclusions:

    The developed protocols allow hydrazides to be converted to acids or amides, enabling late-stage diversification of peptide C-terminal residues.

    Keywords

    Late-stage diversification, peptide hydrazide, peptide acid, peptide amide, antimicrobial peptide

    Introduction

    Peptide therapeutics play a role in drug development because of the unique properties of peptides, which include high specificity, good efficacy, and low immunogenicity when compared to small molecule drugs and biologics [1, 2]. To develop a peptide-based drug, chemical peptide synthesis and subsequent modifications on the sequence and terminal structures are steps that are essential to improve its pharmacological properties after the first step, the discovery of potential peptide therapeutics.

    Modification of the C-terminus of a peptide can significantly affect its bioactivity, stability, or three-dimensional conformations [3]. Even relatively minor structural changes, such as α-amidation of the C-terminal moiety, often significantly affect the peptide properties. For this reason, many strategies to access C-terminally modified peptides have been developed. The most widely used strategy is based on the selection of linkers prior to solid-phase peptide synthesis (SPPS) [4, 5]. This is a reliable strategy but the desired functional groups must be incorporated at the beginning of the synthesis because the peptides are usually synthesized beginning at the C-terminus (Figure 1A). Therefore, this type of strategy requires a significant synthetic effort to create a peptide library containing diverse C-terminal residues.

    Synthesis of C-terminally modified peptides. (A) Synthesis of C-terminally modified peptides using different linkers; (B) late-stage diversification of C-terminally modified peptides; (C) hydrazide-mediated late-stage synthesis of carboxylic acids or amides

    Late-stage modification strategies, which can lead to new analogs of a parent structure, have been recognized as useful methods that may allow the performance of only a single SPPS to construct the peptide sequence and complete the subsequent transformation of its C-terminus to generate various analogs (Figure 1B). The key to the success of this strategy in peptides is achieving the chemoselective and site-selective transformation, particularly in the absence of any protecting groups on reactive functional groups. Consequently, hydrazide was used as a chemoselective handle at the C-terminus (Figure 1C). This study focused on transforming hydrazides to carboxylic acids and amides because both of these C-terminal structures are common in bioactive peptides, often significantly affecting their structure and activity. Oxidation of a hydrazide will convert it to a carboxylic acid [610]. Treatment with β-mercaptoethanol (BME) can generate peptide acids from the corresponding thioesters, which can be prepared from unprotected hydrazides [1113]. Ammonolysis [14] or the Staudinger reaction [15] of the peptide azides which in turn can be prepared from hydrazides are two potential methods for the preparation of peptide amides. Here a systematic and comprehensive evaluation of the relevant transformations is described and their utility by synthesizing modelin-5 derivatives including an artificial antimicrobial peptide is demonstrated [16].

    Materials and methods

    Materials

    All commercially available reagents and protected amino acids were purchased and used without further purification. Rink amide ChemMatrix, trityl (Trt)-OH ChemMatrix, and HMPB ChemMatrix were purchased from Biotage Japan Ltd. Dry dichloromethane (DCM), dry N,N-dimethylformamide (DMF), methanol (MeOH), sodium nitrite (NaNO2), disodium hydrogen phosphate, copper (II) sulfate, 5-hydrate (CuSO4∙5H2O), N,N-diisopropylethylamine (DIPEA), and ammonium acetate (NH4OAc) were purchased from KANTO Chemical Co., Inc. 9-Fluorenylmethyloxycarbonyl (Fmoc) carbazate, 2-iodoxybenzoic acid (IBX), N-bromosuccinimide (NBS), trifluoroacetic acid (TFA), triisopropylsilane (TIS), BME, 4-dimethylaminopyridine (DMAP), piperidine, and m-cresol were purchased from Tokyo Chemical Industry Co., Ltd. DMF, diethyl ether (Et2O), acetonitrile (CH3CN), potassium peroxymonosulfate (Oxone) and tris(2-carboxyethyl)phosphine (TCEP)∙hydrochloride (HCl) were purchased from Fuji Film Wako Pure Chemical Industries, Ltd. Thioanisole, lysogeny broth (LB), Miller, and guanidine (Gn)∙HCl were purchased from NACALAI Tesque, Inc. N,N’-Diisopropylcarbodiimide (DIPCI) and Fmoc-D-alanine (Ala)-OH∙H2O were purchased from Watanabe Chemical Industries Ltd. Ethyl cyanohydroxyiminoacetate (OxymaPure) was purchased from Merck KGaA. Thionyl chloride was purchased from Kishida Chemical Co., Ltd. Fmoc-L-Ala-OH∙H2O and Fmoc-methionine (Met)-OH were purchased from Bachem AG. Fmoc-cysteine (Cys)(Trt)-OH, Fmoc-glutamic acid (Glu)(Ot-Bu)-OH, Fmoc-phenylalanine (Phe)-OH, Fmoc-glycine (Gly)-OH, Fmoc-histidine (His)tert-butoxycarbonyl (Boc)-OH, Fmoc-isoleucine (Ile)-OH, Fmoc-lysine (Lys)(Boc)-OH, Fmoc-leucine (Leu)-OH, Fmoc-proline (Pro)-OH, Fmoc-arginine (Arg)2,2,4,6,7-pentamethyldihydrobenzofuran-5-sulfonyl (Pbf)-OH, Fmoc-serine (Ser)(t-Bu)-OH, Fmoc-threonine (Thr)(t-Bu)-OH, Fmoc-valine (Val)-OH, Fmoc-tyrosine (Tyr)(t-Bu)-OH, and Fmoc-tryptophan (Trp)(Boc)-OH were purchased from CEM corporation. Fmoc-Cys acetamidomethyl (Acm)-OH was purchased from Peptide Institute, Inc.

    LCMS and preparative HPLC

    Liquid chromatography-mass spectrometry (LCMS) analyses were carried out on a Waters Alliance high-performance liquid chromatography (HPLC) system with Acquity single quadrupole (SQ) Detector (ESI-SQ). For HPLC separations, a Cosmosil 5C18-AR-II analytical column (Nacalai Tesque, 4.6 mm × 250 mm, flow rate 1.0 mL/min) and a Cosmosil 5C18-AR-II preparative column (Nacalai Tesque, 20 mm × 250 mm, flow rate 8.0 mL/min) were employed, and the eluted products were detected by ultra-violet (UV) at 220 nm. A solvent system consisting of 0.1% TFA aqueous solution (aq; v/v, solvent A) and 0.1% TFA in CH3CN (v/v, solvent B) was used for HPLC elution.

    Synthesis of model peptide hydrazides, H-LYRA-Xaa-NHNH2 (1a–q)

    The hydrazine-incorporated resin was prepared using previously reported methods [17]. Trt-OH ChemMatrix resin was swollen in dry DCM and treated with thionyl chloride (10 equiv) at room temperature (rt) for 2 h. After being washed with dry DCM and dry DMF, the resulting resin was swollen in dry DCM. To the resin was added a solution of Fmoc carbazate (4 equiv) and DIPEA (10 equiv) in dry DMF (same volume as DCM) dropwise at 0℃. The reaction mixture was shaken at rt overnight, then MeOH was added. After 10 min, the resin was filtered and washed successively with DMF, water, DMF, MeOH, and Et2O. The loading was confirmed by quantification of the Fmoc group [18].

    A manual synthesis of Fmoc-SPPS was conducted as follows: (coupling) Fmoc-protected amino acid, OxymaPure, and DIPCI (4 equiv each) were reacted in DMF (0.3 mol/L), 1 h, rt; (Fmoc removal) was achieved with 20% (v/v) piperidine in DMF for 10 min at rt. After the elongation, the resulting resin was treated with TFA-TIS-H2O [95:2.5:2.5 (v/v), 50 μL/1 mg resin) at rt. After 2 h, Et2O was added to the mixture to give a precipitate which was collected by centrifugation and thoroughly washed with Et2O. The LCMS data of the crude materials are shown in Figures S118.

    Oxidative conversion of hydrazides

    The elongated resin containing 1 (5 mg, approximately 1.5 μmol) was deprotected through the protocol described above and the resulting crude material was dissolved in 50% CH3CN-H2O (200 μL) containing an oxidant (30 µmmol, 20 equiv of CuSO4∙5H2O; 7.5 µmmol, 5.0 equiv for Oxone, IBX, and NBS), and the mixture was allowed to react at rt. The reaction progress was monitored by LCMS (Figures S1937).

    BME-mediated conversion of hydrazides

    The resin of elongated 1 (5 mg, approximately 1.5 µmol) was deprotected through the protocol described above. The resulting crude material was dissolved in 300 µL of a low pH buffer (6 mol/L Gn∙HCl, 0.2 mol/L Na phosphate, pH 3), then 30 µL of 0.5 mol/L NaNO2 in a low pH buffer was added to the solution at –10℃ [19]. After 30 min reaction at –10℃, 1 mol/L BME in a high pH buffer (6 mol/L Gn∙HCl, 0.2 mol/L Na phosphate, pH 8, 300 µL) was added to the mixture. The pH was adjusted to pH 8.9 and the reaction was allowed to stand for 30 min at rt. The reaction progress was analyzed by LCMS (Figures S3854).

    Ammonolysis of peptidyl azides derived from hydrazides

    The resin of elongated 1 (5 mg, approximately 1.5 µmol) was deprotected through the protocol described above. The resulting crude material was dissolved in 40 µL of a low pH buffer (6 mol/L Gn∙HCl, 0.2 mol/L Na phosphate, pH 3), then 10 µL of 1 mol/L NaNO2 aq was added to the mixture at –10℃. After 30 min reaction at –10℃, 8 mol/L NH4OAc (pH 8.3, 150 µL) was added to the mixture. The reaction was allowed to stand for 30 min at rt and the reaction progress was analyzed by LCMS (Figures S5572).

    Amidation of peptidyl azides by the Staudinger reaction

    The reaction via azidation in an acidic aq was performed as follows. The resin of elongated 1 (5 mg, approximately 1.5 µmol) was deprotected using the method described above, and the resulting crude material was dissolved in a low pH buffer (6 mol/L Gn∙HCl, 0.1 mol/L phosphate, pH 3, 100 µL). To the solution was added 10% (w/w) NaNO2 aq (5 µL) at –10℃. After 30 min at –10℃, 100 µL of phosphate buffer (0.1 mol/L, 6 mol/L Gn∙HCl, 0.2 mol/L TCEP∙HCl, pH 3) was added. The mixture was allowed to react for 30 min at rt, the reaction progress being analyzed by LCMS (Figures S7390).

    The azidation in TFA solution was performed as follows (Figure S91 and Table S1) [20]. The resin bearing elongated 1 (5 mg, approximately 1.5 µmol) was treated with TFA-TIS-H2O-m-cresol-thioanisole [80:2.5:2.5:5:10 (v/v), 50 µL/1 mg resin] at rt. After a 2-h reaction, 10% (w/w) NaNO2 aq (5 µL) was added to the mixture at –10℃. The solution was allowed to react for 30 min at –10℃ and then cold Et2O was added to give a precipitate. This crude product was dissolved in 50% CH3CN-H2O (200 µL) containing TCEP·HCl (2 equiv) and the mixture was allowed to stand at rt for 30 min. The reaction progress was analyzed by LCMS (Figures S92109).

    Evaluation of epimerization

    To monitor the epimerization, the amino acids were coupled by DIPCI and DMAP for carboxylic acids or Rink amide ChemMatrix resin for amides. Then, epimers with C-terminal D-amino acids were synthesized by standard manual Fmoc SPPS. The resulting epimers were used to determine the HPLC conditions for the separation (Figures S110 and 111).

    Synthesis of modelin-5 derivatives

    Hydrazide-incorporated resin (0.31 mmol/g) was used for the preparation of the hydrazide (9). On this resin, the peptide was elongated using an automated Fmoc SPPS procedure with a Biotage Initiator+ Alstra peptide synthesizer with a 10 mL open-type vial, coupling with Fmoc-protected amino acids, OxymaPure, and DIPCI (4 equiv each) in DMF (0.5 mol/L), for 5 min at 75℃ under microwave irradiation; Fmoc removal was accomplished with 20% (v/v) piperidine in DMF for 3 min at 50℃ under microwave irradiation. The resulting resin was treated with TFA-TIS-H2O [95:2.5:2.5 (v/v); 50 μL/1 mg resin] at rt for 2 h. The resin in the reaction mixture was removed by filtration and then the filtrate was concentrated in a stream of air. The crude product was dissolved in 0.1% TFA aq (3 mL) and the aqueous phase was washed with Et2O. The crude peptide was purified by preparative HPLC with a linear gradient of solvent B in solvent A, 22–32% over 60 min to give the peptide hydrazide (9; 20 mg from 103 mg of protected resin, 34% isolated yield, Figure S112).

    For the preparation of the peptide acid (10), the crude 9, prepared as described above (from 102 mg of protected resin, 23 µmol) was dissolved in 3 mL of 50% CH3CN-H2O with Oxone (116 µmol, 5 equiv), after which the solution was incubated at 37℃ for 30 min. The crude peptide was purified by preparative HPLC with a linear gradient of solvent B in solvent A, 22–32% over 60 min to give the purified peptide (10; 18 mg, 32% isolated yield, Figure S113).

    For the preparation of the peptide amide (11), the crude 9, prepared from 99 mg (23 µmmol) of protected resin, as described above was dissolved in 1.5 mL of 6 mol/L Gn∙HCl, 0.1 mol/L phosphate (pH 3). After the addition of 0.1 mL of 10% (w/w) NaNO2 aq at –10℃, the mixture was allowed to react for 30 min at the same temperature. To the solution was added 1.5 mL of a phosphate buffer (0.1 mol/L, 6 mol/L Gn∙HCl, 0.2 mol/L TCEP·HCl, pH 3), and then the reaction mixture was incubated at rt for 30 min. The crude peptide was purified by preparative HPLC with a linear gradient of solvent B in solvent A, 22–32% over 60 min to give the peptide amide (11; 21 mg, 37% isolated yield, Figure S114).

    Evaluation of antimicrobial activity

    Escherichia coli (E. coli) W3110 [21] was grown on LB [1% (w/v) tryptone, 0.5% (w/v) yeast extract, and 1% (w/v) NaCl] containing 1.5% agar, and colonies were inoculated in 5 mL LB medium at 200 rpm at 37℃ for approximately 4 h (OD600 = 1.6). The bacterial cells were collected by centrifugation (6,000 × g, 10 min, 4℃) and washed with 1 mL of a 25% strength Ringer’s solution three times through centrifugation (6,000 × g, 5 min, 4℃). The bacterial cells were suspended with Ringer’s solution and diluted to adjust the OD600 to 0.5. A 100 µL sample of the resulting suspensions was mixed with 100 µL of synthesized modelin-5 derivatives (9–11) in Ringer’s solution to give a final peptide concentration of 500 µmol/L. This mixture was incubated in an orbital shaker (100 rpm, 37℃) for 1 h. The bacterial suspension was inoculated onto LB agar plates and incubated at 37℃. After 14 h incubation, the colony forming unit (CFU) was calculated.

    Results

    To examine the feasibility of converting peptide hydrazides to carboxylic acids, the model peptide (H-LYRAG-NHNH2, 1a) was treated with several oxidants (Figure 2 and Table 1). Although copper (II) sulfate (CuSO4) was effective in converting an N-alkyl hydrazide peptide to carboxylic acid [6, 7], it produced the corresponding carboxylic acid (2a) only in a low yield (entry 1). Oxone oxidation proceeded smoothly within 30 min (entry 2). Treatment with IBX or NBS consumed the starting material but failed to generate the desired product (entries 3 and 4). The plausible mechanism of Oxone-mediated transformation of hydrazides to acids is given in Figure 3 [22].

    Direct oxidation of peptide hydrazides

    Optimization of oxidant

    EntryOxidantEquivpHHPLC purity of 2a (%)a
    1CuSO4∙5H2O203.811
    2Oxone52.193
    3IBX51.8< 1
    4NBS54.4< 1
    Display full size

    a detected at 220 nm

    Plausible mechanism of Oxone-mediated transformation of hydrazides to acids

    Formation of azides from hydrazides followed by a reaction with BME is another approach for the conversion of hydrazides to carboxylic acids. Unprotected hydrazides can be transformed into peptidyl azides with the aid of NaNO2 under acidic conditions. In the presence of thiols [19], azides readily afford the corresponding thioesters. Thioesters of BME are formally hydrolyzed through the intramolecular S-to-O acyl transfer followed by thiirane formation, releasing the appropriate carboxylate (Figure 4) [11]. This reaction sequence has been used for the epimerization-free synthesis of peptide acids with a C-terminal Cys [12], but has not been systematically and comprehensively evaluated.

    BME-mediated formal hydrolysis of peptidyl azides preparable from unprotected hydrazides

    To evaluate the scope of these two methods to generate carboxylic acids, peptide hydrazides bearing amino acids at the C-terminus were prepared. The exceptions were aspartic acid (Asp), asparagine (Asn), and glutamine (Gln) because of the synthetic difficulty due to intramolecular cyclization [19]. Using oxidation by Oxone, the desired carboxylic acids were obtained with 63–87% purity (Figure 5, Table 2). Oxidation-prone amino acids, namely Cys, Met, and Trp, were however incompatible with this condition. A separate protocol using BME after azide formation succeeded in the transformation of hydrazides to the corresponding carboxylic acids in 60–95% yield. Ile and Pro required a longer reaction time to completely convert the intermediate thioester to the acid (entries 4 and 7). Unlike the reaction conditions of the Oxone reaction, the BME-mediated reaction was applicable to the peptides with C-terminal Cys, Met, or Trp (entries 11, 13, and 18). The epimerization of C-terminal amino acids (Ala) was suppressed during the reactions (epimer ratio > 99:1, Figure S110).

    Conversion of hydrazides to carboxylic acids

    Scope of the C-terminal amino acids in conversion to carboxylic acids

    EntryXaaProductHPLC purity of 2 (%)a
    OxoneBME
    1Gly2a8482
    2Ala2b8282
    3Val2c7776
    4Ile2d6361 (74)b
    5Leu2e8079
    6Phe2f8789
    7Pro2g7765 (87)b
    8Ser2h8281
    9Thr2i8191
    10Glu2j8195
    11Cys2k< 166
    12Cys(Acm)2k’< 1-
    13Met2l< 180
    14Tyr2m8187
    15His2n8185
    16Lys2o7675
    17Arg2p8460
    18Trp2q< 186
    Display full size

    a detected at 220 nm; b after the addition of BME, a 20 h reaction was performed. -: not examined

    To convert a peptide hydrazide to the peptide amide, two methods were used: 1) Azidation followed by ammonolysis, and 2) azidation followed by the Staudinger reaction (Figure 6). After preliminary experiments, peptide amides were found to be formed by ammonolysis of peptidyl azides using 6 mol/L NH4OAc in a slightly basic buffer. Similarly, the Staudinger reaction of peptidyl azides using 0.1 mol/L TCEP·HCl as a water-soluble phosphine gave the desired amide within 30 min.

    Synthesis of peptide amides from hydrazides

    With the optimal conditions in hand, the scope of these conditions with respect to various C-terminal amino acids was examined (Figure 7 and Table 3). In the NH4OAc-mediated ammonolysis, the desired amides (4) were obtained together with the hydrolyzed products (2) and compounds with molecular weight 17 less than these, probably because of the intramolecularly cyclized peptides (8). In the case of C-terminal Glu and His, significant amounts of the hydrolyzed peptide (2) were obtained (entries 10 and 15). The peptide with a C-terminal Cys residue afforded a low yield because of the formation of a disulfide-related byproduct, which was suppressed by using Acm protection for the thiol (entries 11 and 12).

    Conversion of hydrazides to amides

    Scope of the C-terminal amino acids in conversion to amides

    EntryXaaProductHPLC purity (%)a
    AmmonolysisStaudinger reaction
    42842
    1Gly2a8656946
    2Ala2b8049847
    3Val2c8228911
    4Ile2d91< 1< 192< 1
    5Leu2e74513944
    6Phe2f876< 1916
    7Pro2g9014862
    8Ser2h7939934
    9Thr2i71313913
    10Glu2j< 177< 14 (74)b87
    11Cys2k39c< 1< 1267
    12Cys(Acm)2k’7591093< 1
    13Met2l8069874
    14Tyr2m8382855
    15His2n4539468 (83)b28
    16Lys2o8656876
    17Arg2p916< 1915
    18Trp2q828< 1808
    Display full size

    a detected at 220 nm; b azidation was performed in TFA solution instead of acidic aqueous condition; c disulfide dimer of 2k was observed

    The procedure using the Staudinger reaction also proceeded smoothly, yielding the peptide amides (4) in acceptable yields (80–94%) except for sequences containing Glu-, His-, or Cys-. Glu- or His-containing peptides afforded a low yield of the desired products, but 87% and 28% of the hydrolyzed product (2) were observed (entries 10 and 15). When performing the azide formation in a TFA cocktail [20] instead of under aqueous conditions, the hydrolysis was suppressed, giving the desired amides in 74% for Glu and 83% for His. The thiol group of Cys-containing peptide was oxidized to the sulfinic acid while the side reaction was suppressed using Cys(Acm) as with the ammonolysis condition (entries 11 and 12). Azidation in TFA followed by the Staudinger reaction was also effective in generating the corresponding amides from hydrazides, but in some cases, amines, oxazolidinones, or thiazolidinones were observed as a main byproduct formed through Curtius rearrangement followed by hydrolysis or cyclization of the resulting isocyanate (Table S1). The peptides having Ala at the C-terminus showed no epimerization in the transformation to amides (epimer ratio > 99:1, Figure S111).

    The applicability of the developed late-stage diversification strategies was validated by synthesizing modelin-5 derivatives (H-KLAKKLAKLAKLAKAL-X; X = OH, NH2, or NHNH2), an artificial antimicrobial peptide (Figure 8A). Modelin-5 is known to have varied antimicrobial potencies depending on its C-terminal structure; in particular, modelin-5-CONH2 is more potent than modelin-5-COOH [23]. The modelin-5 sequence was elongated on a hydrazine-incorporated resin through the conventional Fmoc SPPS. The standard cleavage of the elongated peptide afforded the modelin-5-CONHNH2 (9) in 34% isolated yield (Figure 8B). The global deprotection of the peptidyl resin followed by oxidation by Oxone smoothly yielded the modeline-5-COOH (10) in 32% isolated yield from the resin loading (Figure 8C). The modelin-5-CONH2 (11) was synthesized in 37% yield through the azidation of the corresponding hydrazide followed by the Staudinger reaction (Figure 8D).

    Synthesis and evaluation of modelin-5 derivatives. (A) Synthetic scheme of modelin-5 derivatives from hydrazide; crude HPLC charts of (B) modelin-5-CONHNH2; (C) modeline-5-COOH; or (D) modelin-5-CONH2; (E) antimicrobial activity of modelin-5 derivatives against E. coli. Error bars represent the standard deviation (n = 3). Retention time (tR): 14.5 min for peptide hydrazide (9), 15.1 min for peptide acid (10), and 15.4 min for peptide amide (11). HPLC conditions: Cosmosil 5C18-AR-II analytical column with the linear gradient of MeCN/0.1% aqueous TFA (10:90–60:40 over 30 min) at flow rate 1.0 mL/min, detected at 220 nm.

    Finally, the antimicrobial potency against E. coli of the synthesized modelin-5 derivatives was evaluated. The cells were incubated with 500 μmol/L of each peptide derivative for 1 h. The treated cells were plated and then incubated at 37℃ for 14 h, after which the number of colonies was counted (Figure 8E). As has been reported [23], the potency of the amide derivative (11) was higher than that of the carboxylic acid (10). Surprisingly, the hydrazide-type derivative (9) showed the most potent antimicrobial activity against E. coli but the mechanism that underlies such potency differences was unclear.

    Discussion

    Derivatization of peptides is an important method to control and improve their properties. Modification of the C-terminus of bioactive peptides has important implications for their potency. Late-stage diversification is a practical concept to create libraries with diversity, although it requires a chemoselective toolbox to apply it to peptides because peptides have various types of functional groups. In this study, simple and practical protocols were developed for the late-stage transformation of peptide C-terminus residues from hydrazides to the corresponding peptide acids and amides. Oxone oxidation yielded carboxylic acids directly from hydrazides without any additional transformations although some oxidation-prone amino acids showed low compatibility with this reaction. Chemoselective azidation of hydrazides followed by the addition of BME also succeeded in yielding the carboxylic acids via a thioester intermediate. This method is compatible with oxidation-prone amino acids. Different kinds of reaction sequences have been attempted for some specific examples. The same procedure, hydrazide activation by acetylacetone followed by BME addition [13], azidation followed by hydrolysis in basic aqueous conditions have been used [12, 24], but to the best of our knowledge, such a systematic and comprehensive evaluation has not been completed to date.

    Chemoselective azidation of hydrazides and subsequent ammonolysis or a Staudinger reaction also succeeded in the late-stage amide synthesis although a C-terminal Cys must be protected. C-terminal Glu hampers the efficient amide synthesis probably due to the intramolecular anhydride formation followed by hydrolysis. This undesired reaction was suppressed by using a protocol other than a low pH aqueous buffer, to generate a peptide azide in TFA, but the versatility of the TFA condition was not so strong and side reactions involving Curtius rearrangement occurred more frequently. These results suggest that it is important to choose appropriate reaction conditions according to the components and sequence of peptides.

    A general strategy to obtain both acid and amide peptides requires double SPPS using a resin with different linkers. In contrast, our protocols allow the performance of only a single SPPS to obtain both peptides. For example, the synthesis of modelin-5-COOH and -CONH2 using different linkers requires a total of 64 steps (16 couplings and 16 removals of Nα protection for each SPPS). In comparison, late-stage transformation reduces the number of synthetic steps to 34 (32 steps for peptide elongation and two steps for the transformation) and additionally produces a hydrazide as another component to enrich the library. This benefit was confirmed by evaluating the antimicrobial activity of the modelin-5 derivatives, leading to the unexpected discovery that modelin-5-CONHNH2 showed the most potent activity. Considering the antimicrobial peptide with C-terminal hydrazide, Li et al. [24] reported that the antimicrobial spectrum of Chex1-Arg20, an antimicrobial peptide, was broadened by incorporation of the hydrazide. They hypothesized that the broadened spectrum of activity may be due to the stability against bacterial enzyme attacks being improved in the hydrazide. Since modelin-5-CONH2 has a potent ability to kill Bacillus subtilis via membranolytic modes of action [25], the hydrazide might also improve the stability against bacterial enzymes. Efforts to exploit the mechanism and influence of C-terminal hydrazide modification into the antimicrobial activity are in progress in our laboratory.

    Abbreviations

    Acm:

    acetamidomethyl

    Ala:

    alanine

    aq:

    aqueous solution

    Arg:

    arginine

    BME:

    β-mercaptoethanol

    Boc:

    tert-butoxycarbonyl

    CH3CN:

    acetonitrile

    CuSO4∙5H2O:

    copper (II) sulfate, 5-hydrate

    Cys:

    cysteine

    DCM:

    dichloromethane

    DIPCI:

    N,N’-diisopropylcarbodiimide

    DMF:

    N,N-dimethylformamide

    E. coli:

    Escherichia coli

    Et2O:

    diethyl ether

    Fmoc:

    9-fluorenylmethyloxycarbonyl

    Glu:

    glutamic acid

    Gly:

    glycine

    Gn:

    guanidine

    HCl:

    hydrochloride

    His:

    histidine

    HPLC:

    high-performance liquid chromatography

    IBX:

    2-iodoxybenzoic acid

    Ile:

    isoleucine

    LB:

    lysogeny broth

    LCMS:

    liquid chromatography-mass spectrometry

    Leu:

    leucine

    Lys:

    lysine

    MeOH:

    methanol

    Met:

    methionine

    NaNO2:

    sodium nitrite

    NBS:

    N-bromosuccinimide

    NH4OAc:

    ammonium acetate

    Oxone:

    potassium peroxymonosulfate

    OxymaPure:

    ethyl cyanohydroxyiminoacetate

    Phe:

    phenylalanine

    Pro:

    proline

    rt:

    room temperature

    Ser:

    serine

    SPPS:

    solid-phase peptide synthesis

    TCEP:

    tris(2-carboxyethyl)phosphine

    TFA:

    trifluoroacetic acid

    Thr:

    threonine

    TIS:

    triisopropylsilane

    Trp:

    tryptophan

    Trt:

    trityl

    Tyr:

    tyrosine

    Val:

    valine

    Supplementary materials

    The supplementary material for this article is available at: https://www.explorationpub.com/uploads/Article/file/100823_sup_1.pdf.

    Declarations

    Acknowledgments

    ST is grateful for a scholarship from the Amano Institute of Technology.

    Author contributions

    ST: Investigation, Writing—original draft. MK: Investigation. YT: Methodology. TN and NM: Writing—review & editing, Supervision. KS: Conceptualization, Funding acquisition, Writing—original draft, Writing—review & editing. All authors read and approved the submitted version.

    Conflicts of interest

    The authors declare that they have no conflicts of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of data and materials

    Not applicable.

    Funding

    This research was supported in part by JSPS KAKENHI Grant number [JP22K05349] and Research Institute of Green Science and Technology Fund for Research Project Support [2023-RIGST-22201] National University Corporation Shizuoka University. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

    Copyright

    © The Author(s) 2023.

    References

    Wang L, Wang N, Zhang W, Cheng X, Yan Z, Shao G, et al. Therapeutic peptides: current applications and future directions. Sig Transduct Target Ther. 2022;7:48. [DOI] [PubMed] [PMC]
    Muttenthaler M, King GF, Adams DJ, Alewood PF. Trends in peptide drug discovery. Nat Rev Drug Discov. 2021;20:30925. [DOI] [PubMed]
    Arbour CA, Mendoza LG, Stockdill JL. Recent advances in the synthesis of C-terminally modified peptides. Org Biomol Chem. 2020;18:725372. [DOI] [PubMed] [PMC]
    Alsina J, Albericio F. Solid-phase synthesis of C-terminal modified peptides. Pept Sci. 2003;71:45477. [DOI] [PubMed]
    Moss JA. Guide for resin and linker selection in solid-phase peptide synthesis. Curr Protoc Protein Sci. 2005;40:18.7.119. [DOI] [PubMed]
    Sato K, Tanaka S, Wang J, Ishikawa K, Tsuda S, Narumi T, et al. Late-stage solubilization of poorly soluble peptides using hydrazide chemistry. Org Lett. 2021;23:16538. [DOI] [PubMed]
    Fabre B, Pícha J, Vaněk V, Buděšínský M, Jiráček J. A CuAAC–hydrazone–CuAAC trifunctional scaffold for the solid-phase synthesis of trimodal compounds: possibilities and limitations. Molecules. 2015;20:1931029. [DOI] [PubMed] [PMC]
    Srinivas R, Reddy BVS, Yadav JS, Ramalingam T. An efficient and selective conversion of hydrazides into esters and acids. J Chem Res. 2000;2000:3767. [DOI]
    Takale BS, Telvekar VN. Selective oxidation of hydrazides using o-iodoxybenzoic acid to carboxylic acids, esters, and aldehydes. Chem Lett. 2010;39:5467. [DOI]
    Cheung HT, Blout ER. The hydrazide as a carboxylic-protecting group in peptide synthesis. J Org Chem. 1965;30:3156. [DOI] [PubMed]
    Gates ZP, Stephan JR, Lee DJ, Kent SBH. Rapid formal hydrolysis of peptide-αthioesters. Chem Commun. 2013;49:7868. [DOI] [PubMed]
    Zuo C, Yan BJ, Zhu HY, Shi WW, Xi TK, Shi J, et al. Robust synthesis of C-terminal cysteine-containing peptide acids through a peptide hydrazide-based strategy. Org Biomol Chem. 2019;17:5698702. [DOI] [PubMed]
    Liang LJ, Chu GC, Qu Q, Zuo C, Mao J, Zheng Q, et al. Chemical synthesis of activity-based E2-ubiquitin probes for the structural analysis of E3 ligase-catalyzed transthiolation. Angew Chem Int Ed. 2021;60:171717. [DOI] [PubMed]
    Tsuda S, Masuda S, Yoshiya T. Solubilizing trityl-type tag to synthesize Asx/Glx-containing peptides. Chembiochem. 2019;20:20639. [DOI] [PubMed]
    Staudinger H, Meyer J. Über neue organische phosphorverbindungen III. Phosphinmethylenderivate und phosphinimine. Helv Chim Acta. 1919;2:63546. German. [DOI]
    Bessalle R, Gorea A, Shalit I, Metzger JW, Dass C, Desiderio DM, et al. Structure-function studies of amphiphilic antibacterial peptides. J Med Chem. 1993;36:12039. [DOI] [PubMed]
    Huang YC, Chen CC, Li SJ, Gao S, Shi J, Li YM. Facile synthesis of C-terminal peptide hydrazide and thioester of NY-ESO-1 (A39-A68) from an Fmoc-hydrazine 2-chlorotrityl chloride resin. Tetrahedron. 2014;70:29515. [DOI]
    Eissler S, Kley M, Bächle D, Loidl G, Meier T, Samson D. Substitution determination of Fmoc-substituted resins at different wavelengths. J Pept Sci. 2017;23:75762. [DOI] [PubMed] [PMC]
    Fang GM, Li YM, Shen F, Huang YC, Li JB, Lin Y, et al. Protein chemical synthesis by ligation of peptide hydrazides. Angew Chem Int Ed. 2011;50:76459. [DOI] [PubMed]
    Sato K, Tanaka S, Yamamoto K, Tashiro Y, Narumi T, Mase N. Direct synthesis of N-terminal thiazolidine-containing peptide thioesters from peptide hydrazides. Chem Commun. 2018;54:912730. [DOI] [PubMed]
    Hayashi K, Morooka N, Yamamoto Y, Fujita K, Isono K, Choi S, et al. Highly accurate genome sequences of Escherichia coli K-12 strains MG1655 and W3110. Mol Syst Biol. 2006;2:2006.0007. [DOI] [PubMed] [PMC]
    Kulkarni PP, Kadam AJ, Desai UV, Mane RB, Wadgaonkar PP. A simple and efficient oxidation of hydrazides to N,N’-diacylhydrazines using Oxone® in an aqueous medium. J Chem Res. 2000;2000:1845. [DOI]
    Dennison SR, Phoenix DA. Influence of C-terminal amidation on the efficacy of modelin-5. Biochemistry. 2011;50:151423. [DOI] [PubMed]
    Li W, Tailhades J, Hossain MA, O’Brien-Simpson NM, Reynolds EC, Otvos L, et al. C-Terminal modifications broaden activity of the proline-rich antimicrobial peptide, Chex1-Arg20. Aust J Chem. 2015;68:13738. [DOI]
    Dennison SR, Hauß T, Badiani K, Harris F, Phoenix DA. Biophysical investigation into the antibacterial action of modelin-5-NH2. Soft Matter. 2019;15:421526. [DOI]