Abstract
The design of novel and functional biomimetic foldamers remains a major challenge in creating mimics of native protein structures. Herein, we report the stabilization of a remarkably short β-sheet by incorporating N-(hydroxy)glycine (Hyg) residues into the backbone of peptides. These peptide–peptoid hybrids form unique parallel β-sheet structures by self-assembly upon hydrogenation. Our spectroscopic and crystallographic data suggest that the local conformational perturbations induced by N-(hydroxy)amides are outweighed by a network of strong interstrand hydrogen bonds.
Graphical Abstract

INTRODUCTION
In recent years, much interest has been devoted to the development of peptoids (N-substituted glycines)1 due to their proteolytic stability and the ease to introduce various side chains without incorporating the asymmetric α-stereocenters typically found in peptides.2 A major drawback in controlling these peptoid secondary structures is the lack of hydrogen-bonding motifs and the high backbone flexibility of the central glycine methylene. Interestingly, one subclass of peptoid blocks, namely, N-hydroxy-α-amino acids, can be found in numerous bioactive natural products, such as aurantimycins, polyoxypeptins, dentigerumycin, and penicisulfuranols A–F,3 and other bacterial siderophores.4 Despite an intriguing potential arising from the secondary hydroxamate functional group as a hydrogen-bond donor and acceptor, as well as a metal chelator,5,6 the notorious lack of stability of this motif (pH-dependent decarboxylation and possible dismutation) has hampered most synthetic studies.7 Seemingly, only sparse examples of synthetic and structural studies have been reported on the propensity of N-(hydroxy)glycine (Hyg) to induce secondary interactions like β-turn8 and sheet-like structures (e.g., 1, Figure 1).9 Blackwell’s study demonstrated that Hyg residues positioned at hydrogen-bonded (inward) sites in peptoid sequences like 1 enabled the formation of antiparallel β-sheets. Therefore, we hypothesized that by intercalating α-amino acids in such peptoids, sturdy N-(hydroxy)peptoid–peptide β-sheet hybrids (e.g., 2, Figure 1) could potentially be generated. Given the important acidity of N-(hydroxyl)amides in comparison to typical amide groups,10 we became interested in evaluating their hydrogen-bond donor/acceptor abilities to form β-sheets. N-(hydroxy)9 and N-(alkoxy)peptoids11 are known to preferentially adopt a trans-conformation, which is also appropriate to the design of novel β-sheet tertiary structures. Herein, we report the stabilizing effect of Hyg blocks inside a unique intra/intermolecular hydrogen-bond network for the self-assembly of particularly short β-sheets (Figure 1).12 This work provides a foundation to understand and rationalize the role of N-(hydroxy)amides as hydrogen-bond donor/acceptor motifs in the backbone of peptide–peptoid hybrids.
Figure 1.

Proposed novel β-sheet scaffold built by intercalation of the Hyg blocks (left). Peptoid blocks synthesized in this study (right).
RESULTS AND DISCUSSION
Synthesis of Peptide–Peptoid Hybrids.
The syntheses of N-hydroxy-α-amino acids in both racemic13 and asymmetric14 manners have been previously reported; however, the insertion of these peptoid blocks into peptides has been fairly scarce.15 To develop a robust synthesis of N-(hydroxy)-peptoid–peptide hybrids, a series of N-(benzyloxy)glycine derivatives 3–6 were prepared with typical protecting groups for liquid- or solid-phase peptide synthesis (Figure 1). For the initial study of N-terminal elongation, dipeptide–peptoids 7a–c were synthesized starting from N-(OBn)-Gly-OEt 6a, which was coupled with three different Fmoc-protected amino acids (Scheme 1). As for N-alkyl peptoid blocks, N-alkoxy residues are known to have a low reactivity for amide-bond formation, but a cocktail of HATU/HOAt reagents was found suitable to the synthesis of dipeptide–peptoids 7a–c in 70–87% yields. Compounds 7a–c were then engaged in a sequence of deprotection/coupling toward the tripeptide–peptoids 9a–c. Unfortunately, under the large set of conditions tested,16 N-terminal free dipeptide 8 proved difficult to isolate and the coupling toward 9a–c failed. Instead, large amounts of degradation were observed along with the intramolecular cyclization into diketopiperazines (DKPs) 10a–c. Although Hyg favors trans amide bonds thermodynamically, the low transition-state energy barrier previously reported for the cis–trans isomerization of this motif (~16 kcal/mol)7b can explain the ease of forming DKPs. While N-(benzyloxy)peptoid oligomers are known to be exclusively in a most stable trans amide conformation,11 the fast kinetics of cis–trans isomerization do not allow the cis rotamer to be observed on the NMR time scale. To obtain a direct evidence of the genuine cis–trans isomerization, we thought to force the equilibrium toward the cis rotamer by exploiting the chelating nature of the Hyg residue. Dipeptide–peptoid 7c was therefore hydrogenated to the corresponding Fmoc-Ile-Hyg-OEt dipeptide, which was used as the chelating ligand to form a gallium(III) complex (3:1 ligand/metal ratio).17 In this complex, the strong nuclear Overhauser effect spectroscopy (NOESY) 1H–1H cross-peak signal between the Ile Hα and the Hyg methylene detected by NMR supports the presence of a cis N-(hydroxy)amide rotamer. Taken together, this spectroscopic evidence of a cis rotamer and the facile DKP formation (8 → 10) further support that N-(hydroxy) and N-(benzyloxy)-amides undergo fast cis–trans isomerization.
Scheme 1. N-Terminal Elongation toward Tripeptide–Peptoids 9a–c.

aReaction conditions: (a) Fmoc-AA-OH (4.0 equiv), HATU (4.0 equiv), HOAt (4.4 equiv), N,N-diisopropylethylamine (DIEA; 4.0 equiv) in CH2Cl2/DMF (3:1); (b) piperidine or NHEt2; (c) 3 (4.0 equiv), HATU (4.0 equiv), HOAt (4.4 equiv), DIEA (4.0 equiv) in CH2Cl2.
To circumvent the formation of DKP 10, we turned our attention to a strategy of C-terminal elongation (Scheme 2). Dipeptides 11a–c were synthesized by coupling the acetyl Hyg 4 with three different t-butyl α-amino esters by activation with EDCi or HATU. When similar conditions were tested to couple Fmoc-Hyg 5 with Ile-OtBu, an unexpected cleavage of the benzyl protecting group occurred. The coupling was therefore optimized with N-ethoxycarbonyl-2-ethoxy-1,2-dihydroquinoline (EEDQ) to deliver dipeptide 11d without epimerization.18 A smooth deprotection with trifluoroacetic acid (TFA) leads quantitatively to all of the corresponding free carboxylic acid of dipeptides 11a–d, which were directly coupled with 6a in the presence of HATU/HOAt to deliver tripeptide–peptoids 9a–d. Unfortunately, 9c was highly epimerized at the Ile α-stereocenter (2:3 dr). To improve the pivotal amide-bond formation, the coupling reaction was further evaluated with the unprotected block 6b. In this case, the epimerization of tripeptide–peptoids 9c′ could not be detected by 1H NMR, but the racemization was later confirmed by the presence of a mixture of Ile/allo-Ile residues in the crystal structure of 12c (Figure 2). N-(hydroxy) and N-(benzyloxy)glycine blocks are less reactive than typical N-methylated amino acids, which are already known to be poor nucleophiles in amide-bond formation. Due to the oxygen electronegativity and the steric hindrance carried by the hydroxy- and benzyloxy-side chains during the approach to the activated carbonyl π* orbital, such peptoid ligation is particularly slow and difficult, leading ultimately to potential epimerization. To address this issue, several activating cocktails were evaluated for the difficult coupling of 6a. To our delight, tripeptide–peptoid 9c was prepared with ~17% of epimerization using EEDQ as the activating agent. Given the electron-withdrawing nature of the N-(benzyloxy) side chain, we hypothesized that the N-terminal acetyl of 11c might be unusually basic and participates into the Ile epimerization (see Figure SI-1). Thus, by switching to a Fmoc-protecting group in 11d, the EEDQ coupling was achieved successfully to produce 9d in 21% yield with minimal epimerization (~3%).16,18 To circumvent the low coupling yield, further optimizations were required. A cocktail of pivaloyl chloride with lutidine (hindered base) provided tripeptide–peptoid hybrid 9d in 50% yield with low epimerization (<5%).16 Finally, benzyl groups from all synthetic tripeptide–peptoids 9a–d, 9c′ were removed under typical hydrogenation conditions to afford 12a–d in 79–99% yields.
Scheme 2. C-terminal Elongation toward Tripeptide–Peptoids 12a–d.

aReaction conditions: (a) from 4 (1.0 equiv), DIPEA (1.5 equiv), for 11a,b: EDCi (1.2 equiv), HOBt (1.2 equiv) in CH2Cl2, for 11c: H AA-OtBu (2.0 equiv), HATU (1.5 equiv), HOAt (1.5 equiv) in dimethylformamide (DMF); for 11d: from 5 (1.0 equiv), EEDQ (1.2 equiv) in CH2Cl2; (b) TFA/CH2Cl2 (1:1); (c) for 9a–c′: 6a or 6b(2.0 equiv), HATU (2.0 equiv), HOAt (2.0 equiv), DIEA (2.0 equiv) in DMF, for 9d: PivCl (1.2 equiv), lutidine (1.2 equiv) in CH2Cl2; (d) Pd/C 10 mol %, H2 in EtOH.
Figure 2.

NOESY data of 12c in the region of the N-OH protons (A) in CDCl3, (B) in CD3CN, and (C) in DMSO-d6. (D) X-ray structure of 12c showing the solid-state conformation with interstrand hydrogen bonds represented by the dashed lines in the parallel β-sheet.
Structural Analysis of the Peptide–Peptoid Hybrids by Nuclear Magnetic Resonance (NMR) Spectroscopy.
The synthesis of “turnless” β-sheets inspired by short protein sequences or by peptidomimetics remains a major challenge.19 Indeed, motifs capable of enhancing nonbonded interstrand interactions to stabilize artificial short β-sheets are rare.20 To assess the presence of secondary and potential tertiary structures in the synthetic peptide–peptoid hybrids 9a–d, 9c′, and 12a–d, a careful solution NMR spectral analysis at 18 °C was performed (Figure 2A–C).16,21 A combination of 1H, 1H–1H NOESY, and 1H–13C heteronuclear multiple-quantum correlation (HMBC) NMRs in solvents of various polarity (CDCl3, CD3CN, DMSO-d6) facilitated the conformational and structural analyses of these peptoid–peptide hybrids.16 The cis–trans equilibrium constants for both N-(benzyloxy) and N-(hydroxy)glycine residues in peptides 7a, 11a, and 9a were obtained with Ktrans/cis values > 20 in each case, meaning that independently of the solvent, these amides have an exclusive thermodynamic preference for the trans geometry.16,22 Interestingly, in all constructs 9, 11–12, the methylene protons (Hα/Hα′) of N-(hydroxy) or N-(benzyloxy)glycine blocks are always highly diastereotopic (doublets with a geminal coupling constant of ~17 Hz) with a Δδ(Hα/Hα′) of ~0.5 ppm, which typically characterize well-folded structures. Each hydroxyl proton OH1 and OH2 from Hyg(1) and Hyg(2) residues was easily assigned and differentiated based on distinct NOESY correlations as well as a clear HMBC correlation between OH2 and the vicinal Ile carbonyl. Given the downfield chemical shifts of both OH1 (δ 8.36–8.76 and 9.73–9.91 ppm) and OH2 (δ 8.85–9.44 and 10.19–10.25 ppm) in CDCl3 and DMSO-d6, respectively, tripeptide–peptoids 9c′ and 12a–d, hydroxyl hydrogens were determined to participate in inter- or intramolecular hydrogen bonds. A deshielding of ~0.5 ppm between δ OH1 and OH2 suggests that the OH2 hydroxyl is more strongly hydrogen bonded and likely to participate in the β-sheet stabilization. As shown in Figure 2C, all NOESY cross-peaks obtained for 12c in DMSO-d6 can be attributed to interstrand interactions suggesting that no intermolecular assembly occurs in this solvent. In contrast, in CD3CN and more markedly in CDCl3 (Figure 2A,B), a stronger NOESY correlation between OH1 on one strand and the Ile Hα on the second strand supports the formation of a β-sheet self-assembly. Indeed, in a pair of facing residues i,j on the opposite strands from a parallel β-sheet (e.g., Ile/allo-Ile in 12c,d) residue i is H-bonded to both residues j − 1 and j + 1, which in this case both are Hyg residues.23 As established by Abraham, the difference in chemical shits of exchangeable protons or chemical shift deviation (CSD) due to solvent exposure Δδ = δ(DMSO-d6) – δ(CDCl3) can be used to assess the strength of intra- and intermolecular hydrogen bonds for the OH and NH groups.24 The Δδ(HN) of 0.53 and 0.57 ppm in constructs 12c,d suggest the presence of relatively strong hydrogen bonds likely arising from an intermolecular interaction between β-strands (see Table SI-8). Similarly, Δδ(OH2) of 0.80 and 0.77 ppm in 12c and 12d, respectively, support the existence of hydrogen bonds from the Hyg(2) hydroxyls, while the larger Δδ(OH1) of 1.13 and 1.11 ppm firmly indicate that OH1 hydroxyls are more solvent-exposed and not engaged in hydrogen bonding.16 Collectively, the downfield shifts observed for the exchangeable protons (OH/HN) as well as the small CSD values for both Ile HN and Hyg(2) OH in constructs 12c,d suggest that such hydrogens participate in strong interstrand hydrogen bonding. Furthermore, the self-assembly of 12c was characterized by monitoring the chemical shift deviations of OH/HN protons as a function of concentration in CDCl3 (Figures SI-12 and SI-13).16,21 In the range of concentrations studied, changes in chemical shifts values fitted well to a dimerization isotherm equation representing the assembly of the strands into β-sheets.20a These results also suggest that upon hydrogenation of the N-(benzyloxy)glycine residues, Hyg blocks play an important role in the self-assembly of constructs 12a–d into short β-sheets.
X-ray Analysis.
The structural features of β-sheets revealed by the NMR study were further confirmed in the crystal structure of construct 12c obtained by X-ray analysis (Figure 2D). In this crystal, the parallel β-sheets are packed in a head-to-head manner forming columns through a single and relatively weak hydrogen bond OH1⋯O = C (N-terminal acetyl). Certainly, the low-lying conformation observed in the crystal lattice reflects closely the major conformer of 12c in solution (CDCl3). Both HN and OH2 groups are shown to participate in intermolecular hydrogen bonding to stabilize a parallel β-sheet structure. Both Φ and ψ dihedral angles of Hyg(1/2) and Ile residues collected from the crystal structure are in agreement with an extended β-sheet arrangement (see the Ramachandran plot, Figure SI-14). A notable feature of the parallel β-sheet is the nitrogen pyramidalization in each N-(hydroxy)amide (ω(OH1) 171.5° and ω(OH2) 170.9°). A similar pyramidalization was previously reported by Kirshenbaum11 and others25 for N-alkoxy peptoids. This phenomenon is likely responsible for decreasing the nitrogen lone pair resonance in the N-(hydroxy)amide, which in turn accentuates the basicity of the vicinal carbonyl. This stereoelectronic effect likely explains the unique hydrogen-bond acceptor nature of the N-(hydroxy)amide carbonyl responsible for the unusual parallel β-sheet alignment. In the original report from Blackwell, the central α-amino acid was instead a N-aryl peptoid block (i.e., construct 1 vs 2, Figure 1). In such constructs, the central carbonyls were also good hydrogen-bond acceptors, but these peptoid blocks could not participate in intermolecular H-bond, which could justify the preference for an antiparallel β-sheet direction. Our results on peptoid–peptide hybrids 12a–d complement well Blackwell’s study on peptoids in which Hyg residues were positioned at the H-bonded inward sites of the antiparallel β-sheets.9 Herein, we demonstrated that N-(hydroxy)amide groups can force intermolecular hydrogen bonds with Hyg at both H-bonded and non-H-bonded sites (inward and outward of strands) for stabilizing these novel β-sheets with a different parallel directionality.26
β-Sheet Self-assembly and Thermal Stability.
Secondary and tertiary peptidyl structures have well-established signatures in far-UV circular dichroism (CD), but the assignment of the bands in the peptoids is more challenging due to their important flexibility. N-Alkoxy peptoid oligomers have been proposed by Kirshenbaum to adopt a polyproline II type (PPII) secondary structure that exhibit two maxima at 197 and 215 ± 5 nm.27 The CD spectra obtained for tripeptide–peptoid hybrid 9d present similar bands at 195/216 nm in CH3CN and 200/218 nm in HFIP, suggesting that a similar PPII secondary structure formed (Figure 3). As shown by the superimposition of the CD spectra in Figure 3, the hydrogenation of 9d into 12d triggers a dramatic change in the tripeptide–peptoid 3-dimensional structure. Typically, antiparallel β-sheets are characterized by two bands, a positive maximum at 193–195 nm and a negative exciton at 213–218 nm, profoundly resembling the two bands for parallel β-sheets at 200–203 and 214–221 nm.28 Therefore, the combination of a positive exciton at 201 nm and a large CD band overlay at 213–219 nm obtained for construct 12d conclusively supports the formation of a parallel β-sheet. Furthermore, a positive band at 298 nm, characteristic of a π–π* transition, was also observed in the near-UV CD spectrum of 12d (Figure SI-15C).29 This band arising from an interaction between the proximate fluorenyl π-systems of N-terminal Fmoc groups also supports the existence of a parallel arrangement between β-strands. Overall, the structural elucidation by CD from the Fmoc interaction at 298 nm and the exciton minima at 213/219 nm distinctively confirm a parallel β-sheet self-assembly of 12d in solution.
Figure 3.

CD spectra comparison of 9d and 12d showing the self-assembly process a CD spectra of 9d in acetonitrile (90 mM) and HFIP (103 mM) and of 12d in HFIP (116 mM) at 0 °C. Sample concentrations were accurately determined by UV-absorption of the fluorene (Fmoc) chromophore.
To test the stability of these novel parallel β-sheets, variable temperature NMR spectra (VT-NMR) were recorded between −60 and 110 °C (Figure 4). As expected, the spectra of tripeptide–peptoids 12c,d in DMSO-d6 revealed no evidence of tertiary structures as shown by the small-temperature dependence coefficients (TC, Δδ/T < −4.6 ppb/K) for HN, while both N-(hydroxy)amide hydrogens (OH1/OH2) might exist in a relatively strong intramolecular hydrogen bond (−4.6 < TC < −4.0 ppb/K).30,31 The TC representing the slopes of the best-fitted linear regression of OH1 and OH2 chemical shift drifts as a function of temperature were invariably low for all of the constructs studied in DMSO-d6 (Figure 4A and Table SI-7). It can be concluded that a strong intramolecular sixmembered ring hydrogen bond stabilizes each Hyg residue in DMSO-d6 in single-stranded molecules (Figures SI-6 and SI-7).16 In contrast, the significantly larger TC (OH1) and TC (OH2) values (−17.0 < TC < −12.2 ppb/K) in CDCl3 (Figure 4A) indicate that the hydrogen bonds are weaker in apolar solvents. These counterintuitive results could be explained by the competition between a strong interstrand H-bond and the innate intramolecular H-bond of the Hyg residues (as shown in DMSO-d6). Competing intra- and intermolecular H-bonds result in a net weakening of the hydroxyls’ hydrogen-bond network. The results obtained for β-sheets 12c,d are in line with the interactions reported earlier by Blackwell for a peptoid tetramer β-sheet (e.g., 1, Figure 1). While examining the temperature dependency on the N-(hydroxy)amide chemical shifts in CDCl3, it was found that both hydroxyl signals shifted in a nonlinear fashion (Figure 4B). The sigmoidal dependence of N-(hydroxy)amide proton (OH1/OH2) chemical shifts as a function of temperature is distinctive of a β-sheet denaturation transition.32 The melting curves observed for β-sheets 12c,d indicate a rupture of folding at elevated temperatures. Thus, melting temperatures (Tm) corresponding to the hydrogen-bond network cleavage within the parallel β-sheets of 12c,d were estimated from the best-fitted curves of thermal transition to the Boltzmann equation.16 Tm(OH1) of 23 and 14 ± 5 °C and Tm(OH2) of 46 and 29 ± 5 °C were calculated for the β-sheets of 12c and 12d, respectively (Figure 4B, Table SI-6). In both molecules, the significantly higher melting temperature (>15 °C) associated with the H-bond rupture of hydroxyl OH2 is a direct evidence that the interstrand hydrogen bond from Hyg(2) stabilizes the entire structure.
Figure 4.

(A) Temperature dependence of the HN and OH chemical shifts. (B) Plots of hydroxyl proton (OH1/OH2) chemical shifts as a function of temperature in DMSO-d6 and CDCl3.
CONCLUSIONS
In conclusion, a coupling of the readily available N-(hydroxy)-glycine peptoid blocks (Hyg) has been developed with a low epimerization rate (3–5%) for C- and N-terminal insertion into peptide–peptoid hybrids. The synthetic constructs self-assembled in solution to form short parallel β-sheets, which were characterized by a combination of NMR and CD experiments and confirmed by the crystal structure of 12c. A detailed examination of the major differences in NOESY cross-strand correlations, CSDs, TC, and Tm in various solvents clearly highlights the role of N-(hydroxy)amides from Hyg residues in the self-assembly and the stabilization of these tertiary structures.8a Our results demonstrated that N-(hydroxy)amides not only favor exclusively trans amide conformers via intramolecular H-bonds (Ktrans/cis > 20) but are also amenable to strong interstrand hydrogen bonds in solvents of low polarity. Recent studies from Del Valle have shown that the insertion of N-(amino)33 and N-(hydroxy) α-amino acids15f at specific non-hydrogen-bonded sites (outward positions) can stabilize the β-sheets within hairpins. In contrast, the present study establishes that a combination of N-(hydroxy)amides at hydrogen-bonded (inward) and non-hydrogen-bonded outward sites can be favorable due to a strong intramolecular and interstrand H-bond network. Given the typical difficulty to create stable β-sheets within short peptide constructs,19,20 the ability of Hyg to improve the hydrogen-bond donor/acceptor properties of proteinogenic backbone amides will certainly find numerous applications in foldamer chemistry. Ongoing studies in our laboratory will leverage the unique stereoelectronic properties of Hyg and other N-hydroxy α-amino acids to study the assembly of the parallel and antiparallel β-hairpins in water.34
EXPERIMENTAL SECTION
General Information.
All reagents used in the present paper were acquired from Alfa Aesar, Acros Organics, or Sigma Millipore. All bulk solvents were acquired from Fischer Scientific. Freshly distillated solvents were used in the reactions presented herein. Chloroform was dried over CaCl2 overnight prior to distillation (BP 61 °C) and transferred under an argon atmosphere to a dark glass bottle with 3 Å molecular sieves for storage. Tetrahydrofuran was purified by refluxing with and distilling from sodium with benzophenone and transferred under an argon atmosphere to a dark glass bottle for storage. Dichloromethane was dried over CaCl2 overnight prior to distillation (BP 40 °C) and transferred under an argon atmosphere to a dark glass bottle with 3 Å molecular sieves for storage. Toluene was dried over CaCl2, CaH2, or CaSO4 and molecular sieves and transferred under an argon atmosphere to a dark glass bottle for storage. Full procedures can be found in Purification of Laboratory Chemicals by Armarego, W. L.F., and Chai C. L. L. editor (sixth edition). Reactions were performed in flame-dried glassware under a positive pressure of argon, unless the reaction occurs in water or aqueous solvent. Yields refer to chromatographically and spectroscopically pure compounds, unless otherwise noted. Analytical thin-layer chromatography (TLC) was performed on 0.25 mm glass-backed 60 Å F-254 TLC plates. Flash chromatography was performed using 230–400 mesh silica gels (Silicycle, Inc.). The plates were visualized by exposure to UV light (254 nm) and developed by a solution of phosphomolybdic acid in ethanol, vanillin/sulfuric acid in ethanol, ninhydrin in ethanol, potassium permanganate in water/potassium carbonate/sodium hydroxide, or cerium-ammonium-molybdate in water/sulfuric acid and heat. Melting points were determined using an MPA 160 digital melting point apparatus. The matrix-assisted laser desorption/ionization-time of flight (MALDI-TOF) mass spectra were performed on a Microflex LRF MALDI-TOF. The high-resolution mass spectra (HRMS) were obtained from the University of Florida using an Agilent 6230 TOF instrument, using electrospray ionization (ESI). The infrared spectra were recorded on a Nicolet iS10 FT-IR spectrophotometer (Thermo Scientific) with a SMART iTX ATR accessory. Optical rotation was measured on a JASCO P-2000 polarimeter.
The 1H NMR spectra were recorded on a Varian Mercury400 (400 MHz) spectrometer using Vnmrj 4.2 software, Bruker AV-400 Ultrashield (400 MHz) spectrometer using Topspin 3.5 software, or Varian Mercury500 (500 MHz) spectrometer using Vnmrj 4.2 software and are reported in ppm using solvent as an internal standard (C6D6 at 7.16 ppm, CDCl3 at 7.26 ppm, CD3CN at 1.94 ppm, CD3OD at 3.31 ppm, DMSO-d6 at 2.50 ppm, and 1,4-dioxane-d8 at 3.53 ppm). Data are reported as (br = broad, s = singlet, d = doublet, t = triplet, q = quartet, p = pentet, m = multiplet; coupling constant(s) in hertz, integration). The 13C NMR spectra were recorded on a Bruker (100 MHz) spectrometer. Chemical shifts are reported in ppm, with solvent resonance employed as the internal standard (C6D6 at 128.1 ppm, CDCl3 at 77.2 ppm, CD3CN at 118.3 and 1.3 ppm, CD3OD at 49.0 ppm, and DMSO-d6 at 39.5 ppm). The NOESY spectra were recorded at 291 K, in the solvents mentioned above, at concentrations of 5–10 mM, with a 300 ms mixing time.
Variable temperature NMR were recorded on a Varian Mercury400 (400 MHz) spectrometer. The range of temperature studied was −60–70 °C in CDCl3, −40–80 °C in CD3CN, 0–100 °C in dioxane-d8, and 20–110 °C in DMSO-d6. The spectra were recorded with 16 scans every 10 °C, with a 5 min stabilization at each temperature before data acquisition. A final spectrum was recorded afterward at 20 °C to verify the absence of decomposition or transformation of the sample.
The CD spectra were recorded on a JASCO J-810 spectropolarimeter with a temperature controller module JASCO PFD-425S. The samples were prepared in a concentration range between 30 and 200 μM in CH3CN or HFIP, and the sample concentrations were accurately determined by measuring the sample’s absorbance using a JASCO V-670 spectrophotometer based off the UV-absorption of the fluorenyl (Fmoc) group (ε290 = 6089 M−1 cm−1) at 290 nm. The concentrations were obtained based on the Beer law: c = A/(ε·l), with c the sample concentration, A the measured absorptivity, ε the molar absorptivity, and l the path length of the cuvette. The raw CD data were recorded in mdeg with eight scans from 185 to 270 nm for the far-UV and 270–350 nm for the near-UV, every 0.1 nm at a speed of 100 nm per min. The CD spectra of the blank (pure solvent) were recorded and subtracted, while the baseline was set to 0 mdeg between 260 and 270 nm for the far-UV and 330–340 nm for the near-UV. The spectra were smoothed and baseline-corrected using the Spectragryph 1.2 software,35 then converted into molar elipticity (deg cm2 dmol−1). The products of functionalization 3, 6a, 6b, SI-1, and SI-2 have been reported previously and characterized in refs 36d,c,b,e,a, respectively.
Synthesis of N-(benzyloxy)glycine (3).

O-Benzyl hydroxylamine hydrochloride (5.00 g, 31.1 mmol, 1.0 equiv) and glyoxylic acid monohydrate (4.33 g, 47.0 mmol, 1.5 equiv) were solubilized in AcOH/water (1:1 v/v, 100 mL) and stirred at RT during 30 min (completion monitored by TLC, oxime intermediate, Rf = 0.8, EtOAc/MeOH 1:1 v/v). The mixture was cool down to 0 °C with an ice bath and a solution of sodium cyanoborohydride (7.90 g, 125 mmol, 4.0 equiv) in AcOH/water (1:1 v/v, 50 mL) was added portion wise to avoid byproduct formation. The mixture was stirred for 5 h at RT. The reaction completion was monitored by TLC until the disappearance of the oxime intermediate. Water was then added and the crude reaction mixture was extracted with a mixture of iPrOH/CHCl3 (5 × 20 mL). The combined organic layers were washed with saturated brine (3 × 20 mL), dried over Na2SO4, and filtrated, and the solvent was removed under reduced pressure (T < 35 °C) to afford 3 as a white solid (5.60 g, 30.9 mmol, 99% yield). Rf = 0.6 (100% EtOAc); HRMS-ESI (m/z): [M + H]+ calcd for C9H12NO3, 182.0812; found, 182.0810 (−1.1 ppm). N-(Benzyloxy)-glycine 3 was pure enough to be engaged directly in the second step.
N-Acetyl-N-(benzyloxy)glycine (4).

Compound 3 (3.00 g, 16.6 mmol, 1.0 equiv) and sodium carbonate (1.76 g, 16.6 mmol, 1.0 equiv) were solubilized in water (17 mL). Acetic anhydride (1.86 g, 18.2 mmol, 1.1 equiv) was then added dropwise resulting in a precipitation and a yellow coloration. The mixture was stirred for an additional 30 min at RT. An initial wash using diethyl ether (2 × 20 mL) was achieved directly from the crude reaction mixture, and then, the aqueous layer was acidified with a solution of 2.0 M HCl to pH ~ 3. The aqueous layer was extracted with dichloromethane (3 × 20 mL), the combined organic layers were dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford product 4 in a pure form as a yellowish powder (2.86 g, 12.8 mmol, 77% yield). Rf = 0.7 (EtOAc/MeOH 7:3 v/v); mp 109.0–115.0 (±0.6) °C; IR νmax: 2928, 1706, 1596, 1456, 1379, 1260, 964, 911, 847, 749 cm−1; 1H NMR (400 MHz, CDCl3, δ): 9.97 (br, 1H), 7.37 (s, 5H), 4.87 (s, 2H), 4.31 (s, 2H), 2.15 (s, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ): 175.0, 172.5, 134.2, 129.5 (2C), 129.2, 128.8 (2C), 77.5, 49.4, 20.2; HRMS-ESI (m/z): [M + Na]+ calcd for C11H13NO4Na, 246.0737; found, 246.0728 (−3.7 ppm).
N-(((9H-Fluoren-9-yl)methoxy)carbonyl)-N-(benzyloxy)glycine (5).

Compound 3 (1.00 g, 5.52 mmol, 1.0 equiv) and sodium bicarbonate (928 mg, 11.0 mmol, 2.0 equiv) were solubilized in water (10 mL) by stirring 10 min at RT. The mixture was cooled down to 0 °C and (9H-fluoren-9-yl)methyl chloroformate (1.71 g, 6.63 mmol, 1.2 equiv) in dioxane (10 mL) was added to the mixture and stirred for 2 h at 0 °C. The dioxane was evaporated under reduced pressure, a solution of saturated sodium bicarbonate was added (pH ~ 9), and the aqueous phase was washed with hexanes (20 mL). The resulting aqueous layer was acidified with a solution of HCl (2.0 M) to pH ~ 4 allowing for the product precipitation. The product was extracted with dichloromethane (3 × 20 mL), the combined organic layers were dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure. The crude product was purified by precipitation into petroleum ether at RT and then filtrated to afford product 5 in a pure form as a white powder (1.73 g, 4.29 mmol, 77% yield). Rf = 0.7 (100% EtOAc); mp 132.0–137.0 (±0.6) °C; IR νmax: 3039, 2954, 1757, 1737, 1666, 1427, 1350, 1248, 1184, 1116, 872, 760, 733 cm−1; 1H NMR (400 MHz, CD3OD, δ): 7.77 (d, J = 7.5 Hz, 2H), 7.64 (d, J = 7.4 Hz, 2H), 7.37 (t, J = 7.3 Hz, 2H), 7.30 (m, 5H), 7.23 (m, 2H),4.64 (s, 2H), 4.58 (d, J = 6.0 Hz, 2H), 4.25 (t, J = 6.0 Hz, 1H), 4.02 (s, 2H); 13C{1H} NMR (100 MHz, DMSO-d6, δ): 169.9, 157.3, 144.1 (2C), 141.4 (2C), 135.5, 129.9 (2C), 128.9, 128.7 (2C), 128.2 (2C), 127.6 (2C), 125.5 (2C), 120.6 (2C), 76.5, 67.3, 51.6, 47.1; HRMS-ESI (m/z): [M + H]+ calcd for C24H22NO5, 404.1492; found, 404.1485 (−1.7 ppm).
Substrates 6a and 6b were Prepared by the Following Procedures.

The appropriate alkoxyamine hydrochloride (1.0 mmol, 1.0 equiv), ptoluenesulfonic acid (0.1 mmol, 0.1 equiv), and a solution of ethyl glyoxylate (1.1 mmol, 1.1 equiv) in toluene (50% w/w) were mixed in EtOH (1.0 mL) and stirred at RT during 1 h until complete solubilization. The reaction was quenched with addition of a saturated NaHCO3 solution (10 mL), and the product was extracted using dichloromethane (3 × 40 mL). The combined organic layers were washed with saturated brine (40 mL), dried over Na2SO4, filtrated, and the solvent was removed under reduced pressure to afford the corresponding ethyl glyoxylate oxime, which was pure enough to be engaged directly. Ethyl glyoxylate oxime (1.0 mmol, 1.0 equiv) was dissolved in EtOH (0.10 mL). BH3.pyr (1.5 mmol, 1.5 equiv) was added to the mixture, which was cooled down to 0 °C. An ethanolic solution of HCl (6.0 M, 9.0 equiv, freshly prepared by addition of AcCl into EtOH at 0 °C) was added dropwise over a period of 3 h at 0 °C. After the addition of the acid, the mixture was stirred at RT for 30 min (the reaction completion was monitored by TLC until the disappearance of the ethyl glyoxylate oxime). Isolations of 6a and 6b were achieved as described below. Ethyl (benzyloxy)glycinate 6a: The reaction was quenched with addition of an aqueous solution of NaOH (2.0 M) at 0 °C and the crude mixture was concentrated under reduced pressure to remove EtOH. The resulting oil was taken into dichloromethane and washed with water (3 × 40 mL). Combined organic layers were washed with saturated brine (3 × 20 mL), dried over Na2SO4, and filtrated, and the solvent was removed under reduced pressure to afford product 6a as a yellowish oil (2.32 g, 11.1 mmol, 76% yield). Rf = 0.5 (hexanes/EtOAc 4:1 v/v); IR νmax: 3068, 2923, 2854, 1685, 1451, 1231, 1121, 1002, 956, 758, 738 cm−1; 13C{1H} NMR (100 MHz, CDCl3, δ): 171.1, 137.8, 128.4 (3C), 127.9 (2C), 76.1, 61.1, 53.5, 14.2. Ethyl hydroxyglycinate 6b: The crude mixture was concentrated under reduced pressure. The resulting oil was taken into dichloromethane and solid sodium carbonate was added to the solution at 0 °C. The mixture was then stirred overnight at room temperature. The solvent was filtrated and evaporated under reduced pressure to afford product 6b as a yellowish oil (2.43 g, 20.4 mmol, 72% yield). Rf = 0.4 (hexanes/EtOAc 3:2 v/v); 13C{1H} NMR (100 MHz, CDCl3, δ): 171.1, 61.2, 54.9, 14.2.
General Procedure A, for N-O-benzyl α-Amino Ester Coupling With N-Fmoc α-Amino Acids.
Carboxylic acid 6a (4.0 mmol, 4.0 equiv) and the secondary amine Fmoc-AA-OH (1.0 mmol,1.0 equiv) were solubilized in CH2Cl2 and DMF (3:1; 4.0 mL). HOAt (4.4 mmol, 4.4 equiv) and HATU (4.0 mmol, 4.0 equiv) were added and the solution was stirred 10 min at RT. DIPEA (4.0 mmol, 4.0 equiv) was added to the mixture, which was stirred 4–15 h at RT. The reaction progress was monitored by TLC. CH2Cl2 (5.0 mL) was added and the reaction mixture was washed successively with a citric acid solution (10 mL, 5% w/w), saturated NaHCO3 solution (10 mL), and saturated brine (10 mL). The organic layer was dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford the crude product 7a–c, which was purified by chromatography.
Fmoc-N(Me)-Gly-N(OBn)-Ethyl-Glycinate (7a).

was prepared according to the general procedure A, using ethyl N-(benzyloxy)glycinate 6a (50 mg, 0.24 mmol, 1.0 equiv) and Fmoc-Sar-OH (149 mg, 0.48 mmol, 2.0 equiv). The crude product was purified by chromatography (hexanes/EtOAc 9:1 v/v) to afford 7a in a pure form as a white foam (104 mg, 0.21 mmol, 87% yield). Rf = 0.6 (hexanes/EtOAc 1:1 v/v); mp 42–45 °C; IR νmax: 2944, 1747, 1702, 1450, 1399, 1372, 1204, 1146, 992, 758, 739, 700 cm−1; Product 7a was characterized as a mixture of cis–trans rotamers in a ratio 2:3 by 1H–13C HSQC: 1H NMR (400 MHz, CDCl3, δ) trans rotamer: 7.77 (d, J = 7.5 Hz, 2H), 7.63 (d, J = 7.5 Hz, 2H), 7.40 (m, 5H), 7.34 (m, 4H), 4.93 (s, 2H), 4.39 (d, J = 7.0 Hz, 2H), 4.29 (s, 2H), 4.26 (s, 2H), 4.24 (t, J = 3.3 Hz, 1H), 4.21 (q, J = 7.1 Hz, 2H), 3.02 (s, 3H),1.28 (t, J = 6.7 Hz, 3H); cis rotamer: 7.72 (d, J = 7.5 Hz, 2H), 7.56 (d, J = 7.5 Hz, 2H), 7.40 (m, 5H), 7.34 (m, 4H), 7.16 (m, 2H), 4.73 (s, 2H), 4.43 (d, J = 6.6 Hz, 2H), 4.29 (s, 2H), 4.24 (t, J = 3.3 Hz, 1H), 4.18 (q, J = 7.2 Hz, 2H), 4.07 (s, 2H), 2.91 (s, 2H), 1.25 (t, J =7.1 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ) trans rotamer: 167.7 (2C), 157.0, 144.1 (2C), 141.3 (2C), 134.3, 129.6 (2C),129.2, 128.8 (2C), 127.6 (2C), 127.1 (2C), 125.2 (2C), 120.0 (2C), 77.3, 67.9, 61.7, 50.6, 49.9, 47.2, 35.6, 14.1; cis rotamer: 167.6 (2C), 156.4, 144.2 (2C), 141.3 (2C), 134.1, 129.6 (2C), 129.2, 128.8 (2C), 127.6 (2C), 127.1 (2C), 125.1 (2C), 119.9 (2C), 77.5, 67.3, 61.7, 50.3, 50.1, 47.3, 36.0, 14.1; HRMS-ESI (m/z): [M + H]+ calcd for C29H31N2O6, 503.2177; found, 503.2161 (−3.2 ppm).
Fmoc-Gly-N(OBn)-Ethyl-Glycinate (7b).

was prepared according to the general procedure A, using ethyl N-(benzyloxy)glycinate 6a (70 mg, 0.33 mmol, 1.0 equiv) and Fmoc-Gly-OH (398 mg, 1.34 mmol, 4.0 equiv). The crude product was purified by chromatography (hexanes/EtOAc 70:30 v/v) to afford 7b in a pure form as a white foam (113 mg, 0.23 mmol, 70% yield). Rf = 0.2 (hexanes/EtOAc 70:30 v/v); mp 99–101 °C; IR νmax: 3376, 2953, 1745, 1691, 1532, 1400, 1277, 1226, 992, 741, 730, 694 cm−1; 1H NMR (400 MHz, CDCl3, δ): 7.77 (d, J = 7.5 Hz, 2H), 7.61 (d, J =7.4 Hz, 2H), 7.40 (m, 7H), 7.32 (t, J = 7.3 Hz, 2H), 5.50 (br, 1H),4.91 (s, 2H), 4.39 (d, J = 7.2 Hz, 2H), 4.28 (s, 2H), 4.22 (m, 5H),1.27 (t, J = 7.1 Hz, 3H); 13C{1H} NMR (100 MHz, CD3OD, δ): 173.1, 167.9, 157.7, 143.9 (2C), 141.2 (2C), 134.5, 129.4 (2C),128.8, 128.4 (2C), 127.4 (2C), 126.8 (2C), 124.9 (2C), 119.5 (2C), 76.9, 66.8, 61.3, 49.1, 47.1, 41.9, 13.0; HRMS-ESI (m/z): [M + H]+ calcd for C28H29N2O6, 489.2020; found, 489.2013 (−1.4 ppm).
Fmoc-Ile-N(OBn)-Ethyl-Glycinate (7c).

was prepared according to the general procedure A, using ethyl N-(benzyloxy)glycinate 6a (50 mg, 0.24 mmol, 1.0 equiv) and Fmoc-Ile-OH (298 mg, 0.96 mmol, 4.0 equiv). The crude product was purified by chromatography (hexanes/EtOAc 80:20 v/v) to afford 7c in a pure form as a white foam (94 mg, 0.17 mmol, 72% yield). Rf = 0.4 (hexanes/EtOAc 70:30 v/v); mp 40–44 °C; IR νmax: 3057, 2964, 1716, 1660, 1506, 1449, 1264, 1201, 1021, 732, 700 cm−1; 1H NMR (400 MHz, CDCl3, δ): 7.47 (d, J = 7.5 Hz, 2H), 7.62 (d, J = 6.5 Hz, 2H), 7.41 (m, 7H), 7.32 (t, J = 7.4 Hz, 2H), 5.42 (d, J = 9.8 Hz, 1H),5.05 (dd, J = 32.3, 10.1 Hz, 2H), 4.86 (dd, J = 9.7, 6.0 Hz, 1H), 4.64 (d, J = 17.4 Hz, 1H), 4.40 (m, 2H), 4.25 (t, J = 7.2 Hz, 1H), 4.20 (q, J = 7.1 Hz, 2H), 3.99 (d, J = 17.5 Hz, 1H), 1.95 (m, 1H), 1.60 (m, 1H),1.26 (t, J = 7.1 Hz, 3H), 1.16 (m, 1H), 1.01 (d, J = 6.8 Hz, 3H), 0.91 (t, J = 7.4 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ): 174.0, 167.3, 156.5, 144.0, 143.9, 141.3 (2C), 133.9, 129.6 (2C),129.2, 128.8 (2C), 127.7 (2C), 127.1 (2C), 125.2 (2C), 120.0 (2C), 77.9, 67.1, 61.6, 55.4, 48.6, 47.2, 37.3, 23.9, 15.7, 14.1, 11.2; HRMS-ESI (m/z): [M + H]+ calcd for C32H37N2O6, 545.2646; found, 545.2631 (−2.8 ppm).
General Procedure B, for the Coupling of 4 With t-butyl α-Amino Ester Residues.
Carboxylic acid 4 (1.0 mmol, 1.0 equiv) and the H-AA-OtBu (1.2 mmol, 1.2 equiv) were solubilized in CH2Cl2 (5.0 mL) under an argon atmosphere and cooled down to 0 °C. EDCi (1.2 mmol, 1.2 equiv) and HOBt (1.2 mmol, 1.2 equiv) were added to the mixture. DIPEA (2.0 mmol, 2.0 equiv) was added to the mixture, which was stirred 15 h at RT. The reaction progress was monitored by TLC. CH2Cl2 (5.0 mL) was added and the mixture was washed successively with a citric acid solution (10 mL, 5% w/w), saturated NaHCO3 solution (10 mL), and saturated brine (10 mL). The organic layer was dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford the crude product 11a,b, which was purified by chromatography.
Ac-N(OBn)-Gly-N(Me)-tert-Butyl-Glycinate (11a).

was prepared according to the general procedure B, using N-acetyl-N-(benzyloxy)glycine 4 (100 mg, 0.45 mmol, 1.0 equiv) and H-Sar-OtBu (98 mg, 0.54 mmol, 1.2 equiv). The crude product was purified by chromatography (ether/EtOAc 1:1 v/v) to afford 11a in a pure form as an uncolored oil (106 mg, 0.30 mmol, 67% yield). Rf = 0.2 (hexanes/EtOAc 1:1 v/v); IR (dry film) νmax: 2980, 2937, 2865, 1737, 1658, 1454, 1367, 1227, 1154, 1119, 1055, 1032, 1013, 846, 747, 700 cm−1; Product 11a was characterized as a mixture of cis–trans rotamers in a ratio 2:5 by 1H–13C HSQC: 1H NMR (400 MHz, CDCl3, δ) trans rotamer: 7.39 (m, 5H), 4.93 (s, 2H), 4.41 (s, 2H),4.04 (s, 2H), 2.99 (s, 3H), 2.16 (s, 3H), 1.45 (s, 9H); cis rotamer:7.39 (m, 5H), 4.93 (s, 2H), 4.33 (s, 2H), 3.91 (s, 2H), 2.98 (s, 3H),2.14 (s, 3H), 1.48 (s, 9H); 13C{1H} NMR (100 MHz, CDCl3, δ) trans rotamer: 175.1, 168.0, 167.2, 134.9, 129.5 (2C),128.9, 128.7 (2C), 82.0, 77.3, 50.3, 50.0, 35.6, 28.1, 20.4; cis rotamer: 175.1, 167.7, 167.4, 134.8, 129.5 (2C),128.9, 128.7 (2C), 82.9, 77.2, 52.1, 49.8, 35.4, 28.0 (3C), 20.4; HRMS-ESI (m/z): [M + H]+ calcd for C18H27N2O5, 351.1914; found, 351.1911 (−0.9 ppm).
Ac-N(OBn)-Gly-tert-Butyl-Glycinate (11b).

was prepared according to the general procedure B, using N-acetyl-N-(benzyloxy)glycine 4 (500 mg, 2.24 mmol, 1.0 equiv) and H-Gly-OtBu (376 mg, 2.69 mmol, 1.2 equiv). The crude product was purified by chromatography (hexanes/EtOAc 1:1 v/v) to afford 11b in a pure form as a yellowish oil (604 mg, 1.80 mmol, 67% yield). Rf = 0.3 (hexanes/EtOAc 1:1 v/v); IR (dry film) νmax: 3288, 2980, 2941, 2865, 1736, 1685, 1620, 1545, 1454, 1404, 1364, 1223, 1154, 1055, 1033, 1016, 961, 853, 753, 697, 641 cm−1; 1H NMR (400 MHz, CDCl3, δ): 7.39 (s, 5H), 6.61 (br, 1H), 4.93 (s, 2H), 4.30 (s, 2H),3.93 (d, J = 5.1 Hz, 2H), 2.14 (s, 3H), 1.46 (s, 9H); 13C{1H} NMR (100 MHz, CDCl3, δ): 173.8, 168.5, 167.8, 134.1, 129.5 (2C),129.2, 128.8 (2C), 82.4, 77.4, 51.9, 42.0, 28.0 (3C), 20.3; HRMS-ESI (m/z): [M + H]+ Calcd for C17H25N2O5, 337.1758; found, 337.1754 (−1.2 ppm).

Ac-N(OBn)-Gly-tert-Butyl-Isoleucinate (11c).
N-acetyl-N-(benzyloxy)glycine 4 (700 mg, 3.14 mmol, 1.0 equiv), HATU (1.79 g, 4.71 mmol, 1.5 equiv), and HOAt (641 mg, 4.71 mmol, 1.5 equiv) were solubilized in DMF (8.0 mL) under an argon atmosphere. DIPEA (609 mg, 4.71 mmol, 1.5 equiv) was added to the mixture, which was preactivated 5 min at RT. H-Ile-OtBu (1.40 g, 6.28 mmol, 2.0 equiv) was then added and stirred 15 h at RT. The reaction progress was monitored by TLC. CH2Cl2 (10 mL) was added and the mixture was washed successively with a citric acid solution (15 mL, 5% w/w), saturated NaHCO3 solution (15 mL), and saturated brine (15 mL). The organic layer was dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford the crude product 11c in a pure form as a yellow oily solid (841.5 mg, 2.1 mmol, 68% yield). Rf = 0.4 (hexanes/EtOAc 1:1 v/v); (c 1.00, MeOH); IR (dry film) νmax: 3358, 3276, 2962, 2932, 2873, 1735, 1720, 1689, 1650, 1556, 1536, 1448, 1412, 1391, 1365, 1287, 1257, 1151, 1133, 1079, 1028, 1009, 973, 948, 809, 737, 697 cm−1; 1H NMR (400 MHz, CDCl3, δ): 7.38 (s, 5H), 6.66 (d, J = 8.1 Hz, 1H),4.92 (s, 2H), 4.47 (dd, J = 8.5, 4.5 Hz, 1H), 4.34 (d, J = 16.5 Hz, 1H),4.23 (d, J = 16.5 Hz, 1H), 2.14 (s, 3H), 1.87 (m, 1H), 1.58 (m, 1H),1.45 (s, 9H), 1.15 (m, 1H), 0.92 (t, J = 7.4 Hz, 3H), 0.89 (d, J = 6.9 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ): 173.8, 170.5, 167.4, 134.1, 129.4 (2C),129.1, 128.8 (2C), 82.1, 77.2, 56.9, 52.0, 38.1, 28.1 (3C), 25.3, 20.3, 15.3, 11.7; HRMS-ESI (m/z): [M + Na]+ calcd for C21H32N2O5Na, 415.2203; found, 415.2204 (+0.2 ppm).

Fmoc-N(OBn)-Gly-tert-Butyl-Isoleucinate (11d).
N-Fmoc-N-(benzyloxy)glycine 5 (500 mg, 1.24 mmol, 1.0 equiv) and EEDQ (460 mg, 1.86 mmol, 1.5 equiv) were solubilized in CH2Cl2 (3.0 mL) under an argon atmosphere. The mixture was preactivated 5 min at RT. H-Ile-OtBu (832 mg, 3.72 mmol, 3.0 equiv) was then added and stirred 5 h at RT. The reaction progress was monitored by TLC. CH2Cl2 (3.0 mL) was added and the mixture was washed successively with a citric acid solution (10 mL, 5% w/w), saturated NaHCO3 solution (10 mL), and saturated brine (10 mL). The organic layer was dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford the crude product 11d in a pure form as a yellow oily solid (618 mg, 1.08 mmol, 87% yield). Rf = 0.3 (hexanes/EtOAc 4:1 v/v); (c 1.00, MeOH); IR (dry film) νmax: 3343, 3067, 2965, 2936, 2880, 1725, 1685, 1523, 1451, 1414, 1392, 1368, 1340, 1249, 1216, 1143, 1092, 984, 912, 846, 757, 738, 698, 621 cm−1; 1H NMR (400 MHz, CDCl3, δ): 7.75 (d, J = 7.5 Hz, 1H),7.62 (m, 2H), 7.40 (t, J = 7.4 Hz, 2H), 7.32 (m, 7H), 6.55 (d, J = 8.4 Hz, 1H), 4.85 (s, 2H), 4.59 (dd, J = 6.6, 2.8 Hz, 2H), 4.48 (dd, J =8.5, 4.5 Hz, 1H), 4.28 (t, J = 6.6 Hz, 1H), 4.14 (d, J = 17.2 Hz, 1H),4.03 (d, J = 17.2 Hz, 1H), 1.84 (m, 1H), 1.42 (s, 9H), 1.40 (m, 1H),1.11 (m, 1H), 0.86 (m, 6H); 13C{1H} NMR (100 MHz, CDCl3, δ): 170.5, 167.4, 157.5, 143.5 (2C), 141.4 (2C) 134.9, 129.4 (2C),128.7, 128.5 (2C), 127.8 (2C), 127.2 (2C), 125.1 (2C), 120.0 (2C), 82.1, 77.2, 68.3, 56.8, 54.3, 47.1, 38.2, 28.0 (3C), 25.3, 15.3, 11.7; HRMS-ESI (m/z): [M + Na]+ calcd for C34H40N2O6Na, 595.2779; found, 595.2763 (−2.7 ppm).
General Procedure C, for the Coupling of 6a or 6b With 11a–c.
t-Butyl protected dipeptide 11a–c (1.0 mmol, 1.0 equiv) was solubilized in dichloromethane (5.0 mL) and cooled down to 0 °C. Trifluoroacetic acid (5.0 mL) was added dropwise, the mixture was then stirred at 0 °C for 3–5 h. The reaction progress was monitored by TLC. The solvent and residual byproducts were evaporated under reduced pressure to afford the desired unprotected acid dimer. The resulting product and the ethyl (benzyloxy)glycinate 6a or ethyl hydroxyglycinate 6b (2.0 mmol, 2.0 equiv) were solubilized in CH2Cl2 or DMF (5.0 mL). Then, HOAt (2.0 mmol, 2.0 equiv) and HATU (2.0 mmol, 2.0 equiv) were added, and the solution was stirred 10 min at RT. DIPEA (2.0 mmol, 2.0 equiv) was added dropwise to the mixture, which was stirred 15 h at RT. The reaction progress was monitored by TLC. CH2Cl2 (5.0 mL) was added and the mixture was washed successively with a citric acid solution (10 mL, 5% w/w), saturated NaHCO3 solution (10 mL), and saturated brine (10 mL). The organic layer was dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford crude products 9a–d, which were further purified by chromatography.
Ac-N(OBn)-Gly-N(Me)-Gly-N(OBn)-Ethyl-Glycinate (9a).

was prepared according to the general procedure C, using N-(benzyloxy)glycinate 6a (740 mg, 3.54 mmol, 2.0 equiv) and peptide 11a (520 mg, 1.77 mmol, 1.0 equiv). The crude product was purified by chromatography (hexanes/EtOAc 7:3 v/v) to afford 9a as a white foam (277 mg, 0.57 mmol, 32% yield). Rf = 0.3 (hexanes/EtOAc 3:7 v/v); mp 47–49 °C; IR νmax: 2931, 1743, 1658, 1454, 1373, 1204, 1021, 973, 844, 733, 700 cm−1; Product 9a was characterized as a mixture of cis–trans rotamers in a ratio 1:2 by 1H–13C HSQC: 1H NMR (400 MHz, CDCl3, δ) trans rotamer: 7.38 (m, 5H), 4.92 (s, 2H), 4.91 (s, 2H), 4.42 (s, 2H), 4.38 (s, 2H), 4.23 (s, 2H), 4.18 (m, 2H), 2.97 (s, 3H), 2.16 (s, 3H), 1.26 (t, J = 7.1 Hz, 3H); cis rotamer:7.38 (m, 5H), 4.88 (s, 2H), 4.84 (s, 2H), 4.33 (s, 2H), 4.18 (m, 4H),3.98 (br, 2H), 2.80 (s, 3H), 2.10 (s, 3H), 1.26 (t, J = 7.1 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ) trans rotamer: 175.0, 172.3, 167.7, 167.5, 134.9, 134.2, 129.6 (2C), 129.5 (2C), 129.2, 129.0, 128.8 (2C), 128.7 (2C) 77.6 (2C), 61.6, 50.1 (2C), 49.6, 35.9, 20.4, 14.1; cis rotamer: 175.0, 173.3, 167.8, 167.4, 134.9, 134.1, 130.1 (2C), 129.6 (2C), 128.9, 128.8 (2C), 128.7, 128.5 (2C), 82.9, 77.0, 76.9, 61.9, 51.0, 50.1, 49.2, 35.4, 20.3, 14.1; HRMS-ESI (m/z): [M + H]+ calcd for C25H32N3O7, 486.2235; found 486.2214 (−4.3 ppm).
Ac-N(OBn)-Gly-Gly-N(OBn)-Ethyl-Glycinate (9b).

was prepared according to the general procedure C, using N-(benzyloxy)glycinate 6a (242 mg, 1.16 mmol, 2.0 equiv) and peptide 11b (171 mg, 0.58 mmol, 1.0 equiv). The crude product was purified by precipitation into petroleum ether and filtration to afford 9b as a beige solid (127 mg, 0.27 mmol, 46% yield). Rf = 0.3 (hexanes/EtOAc 3:7 v/v); mp 91–96 °C; IR νmax: 3239, 2979, 2953, 1745, 1695, 1674, 1644 1544, 1455, 1403, 1372, 1240, 1207, 1054, 1033, 1007, 972, 737 cm−1; 1H NMR (400 MHz, CDCl3, δ): 7.38 (s, 10H), 6.76 (br, 1H),4.91 (d, J = 10.7 Hz, 2H), 4.29 (s, 2H), 4.26 (d, J = 4.6 Hz, 2H), 4.24 (s, 2H), 4.18 (q, J = 7.2 Hz, 2H), 2.15 (s, 3H), 1.25 (t, J = 7.2 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ): 174.1, 172.5, 167.8, 167.3, 134.2, 133.9, 129.5 (4C),129.3, 129.1, 128.9 (2C), 128.8 (2C),77.8, 77.5, 61.8, 51.8, 50.0, 41.3, 20.4, 14.1; HRMS-ESI (m/z): [M +H]+ Calcd for C24H30N3O7, 472.2078; found, 472.2061 (−3.6 ppm).
Ac-N(OBn)-Gly-Ile-N(OBn)-Ethyl-Glycinate (9c).

was prepared according to the general procedure C, using N-(benzyloxy)glycinate 6a (363 mg, 1.74 mmol, 2.0 equiv) and peptide 11c (292 mg, 0.87 mmol, 1.0 equiv). The crude product was purified by precipitation into petroleum ether and filtration to afford 9c as an orange powder in a mixture of diastereoisomer (189 mg, 0.44 mmol, 41% yield, d.r. 3:2). Rf = 0.7 (hexanes/EtOAc 3:7 v/v); mp 117–120 °C; IR νmax: 3279, 2964, 2933, 1751, 1637, 1544, 1454, 1390, 1375, 1212, 1019, 972, 840, 750, 699 cm−1; 1H NMR (400 MHz, CDCl3, δ) (S, S) isomer: 7.44 (m, 10H), 6.88 (d, J = 9.3 Hz, 1H), 5.10 (dd, J =9.3, 6.7 Hz, 1H), 5.00 (m, 2H), 4.93 (s, 2H), 4.58 (d, J = 8.3 Hz, 1H),4.33 (d, J = 3.5 Hz, 2H), 4.17 (m, 3H), 2.12 (s, 3H), 1.90 (m, 1H),1.55 (m, 1H), 1.24 (t, J = 7.1 Hz, 3H), 1.13 (m, 1H), 0.94 (d, J = 6.9 Hz, 3H), 0.86 (m, 3H); (R, S) isomer: 7.44 (m, 10H), 6.83 (d, J = 8.9 Hz, 1H), 5.19 (dd, J = 9.3, 4.3 Hz, 1H), 5.01 (s, 2H), 5.00 (m, 2H),4.63 (d, J = 8.4 Hz, 1H), 4.33 (d, J = 3.5 Hz, 2H), 4.17 (m, 3H), 2.12 (s, 3H), 1.97 (m, 1H), 1.39 (m, 1H), 1.24 (t, J = 7.1 Hz, 3H), 1.20 (m, 1H), 0.87 (m, 3H), 0.84 (m, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ) (S, S) isomer: 173.7, 172.9, 167.7, 167.4, 135.0, 134.4, 129.7 (2C),129.6 (2C), 128.8, 128.7 (2C), 128.6 (2C), 77.2, 76.5, 61.3, 53.0, 50.2, 48.1, 36.9, 23.9, 19.8, 15.0, 13.5, 10.4; (R, S) isomer: 173.7, 173.2, 167.7, 167.6, 135.0, 134.3, 129.7 (2C),129.6 (2C), 129.0, 128.7 (2C), 128.6 (2C), 77.1, 76.5, 61.3, 51.9, 50.2, 48.2, 36.3, 26.4, 19.8, 15.0, 13.5, 11.1; HRMS-ESI (m/z): [M + H]+ calcd for C28H38N3O7, 528.2704; found, 528.2681 (−4.4 ppm).
Ac-N(OBn)-Gly-Ile-N(OH)-Ethyl-Glycinate (9c′).

was prepared according to the general procedure C, using N-hydroxyglycinate 6b (283 mg, 2.38 mmol, 2.0 equiv) and peptide 11c (400 mg, 1.19 mmol, 1.0 equiv). The crude product was purified by precipitation into petroleum ether and filtration to afford 9c′ as a light brown solid (231 mg, 0.53 mmol, 53% yield). Rf = 0.4 (hexanes/EtOAc 3:7 v/v); (c 1.00, MeOH); mp 118–120 °C; IR νmax: 3280, 2968, 2934, 1753, 1655, 1637, 1541, 1372, 1245, 1201, 1018, 737, 696, 621 cm−1; 1H NMR (400 MHz, CD3CN, δ): 8.38 (s, 1H), 7.43 (m, 5H), 6.88 (d, J = 8.9 Hz, 1H), 4.97 (dd, J = 9.1, 6.5 Hz, 1H), 4.92 (s, 2H), 4.44 (d, J = 17.5 Hz, 1H), 4.29 (d, J = 2.1 Hz, 2H),4.25 (d, J = 17.5 Hz, 1H), 4.17 (q, J = 7.0 Hz, 2H), 2.11 (s, 3H), 1.97 (m, 1H), 1.52 (m, 1H), 1.24 (t, J = 7.1 Hz, 3H), 1.11 (m, 1H), 0.94 (d, J = 6.9 Hz, 3H), 0.90 (t, J = 7.4 Hz, 3H); 13C{1H} NMR (100 MHz, CD3CN, δ): 173.8, 171.9, 167.9, 167.7, 134.9, 129.6 (2C),128.8, 128.6 (2C), 76.5, 61.2, 53.3, 50.1 (2C), 36.5, 23.8, 19.7, 15.1, 13.5, 10.6; HRMS-ESI (m/z): [M + H]+ calcd for C21H32N3O7, 438.2235; found, 438.2214 (−4.8 ppm).
Fmoc-N(OBn)-Gly-Ile-N(OBn)-Ethyl-Glycinate (9d).

Peptide 11d (190 mg, 0.37 mmol, 1.0 equiv), lutidine (0.03 mL, 0.44 mmol, 1.2 equiv), and 6a (308 mg, 1.47 mmol, 4.0 equiv) were solubilized in CH2Cl2 (1.0 mL) under an argon atmosphere. ClCOtBu (0.05 mL, 0.44 mmol, 1.2 equiv) was then added and stirred 15 h at RT. The reaction progress was monitored by TLC. CH2Cl2 (3.0 mL) was added and the mixture was washed successively with a citric acid solution (3 × 5.0 mL, 5% w/w), saturated NaHCO3 solution (3 × 5.0 mL), and saturated brine (5.0 mL). The organic layer was dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure to afford the crude product 9d in a pure form as a yellowish oily gum (135 mg, 0.19 mmol, 51% yield). Rf = 0.65 (hexanes/EtOAc 6:4 v/v); (c 1.00, MeOH); IR ν(dry film) νmax: 3320, 2961, 2934, 2876, 1743, 1712, 1655, 1526, 1451, 1419, 1389, 1344, 1200, 1094, 1022, 979, 911, 854, 739, 698 cm−1; 1H NMR (400 MHz, CDCl3, δ) 7.75 (d, J = 7.5 Hz, 2H), 7.64 (dd, J =7.3, 2.4 Hz, 2H), 7.37 (m, 14H), 6.59 (d, J = 9.4 Hz, 1H), 5.15 (dd, J = 9.4, 5.9 Hz, 1H), 5.03 (d, J = 10.1 Hz, 1H), 4.93 (d, J = 10.1 Hz, 1H), 4.86 (dd, J = 13.6, 10.1 Hz, 2H), 4.59 (m, 3H), 4.29 (m, 1H),4.16 (m, 3H), 4.06 (d, J = 17.1 Hz, 1H), 3.90 (d, J = 17.5 Hz, 1H),1.92 (m, 1H), 1.52 (m, 1H), 1.23 (t, J = 7.1 Hz, 3H), 1.03 (m, 1H), 0.95 (d, J = 6.8 Hz, 3H), 0.82 (t, J = 7.4 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ) 173.3, 167.8, 167.3, 157.7, 143.6 (2C), 141.4 (2C), 135.0, 133.9, 129.6 (2C), 129.4 (2C), 129.1, 128.8, 128.7, 128.5 (2C), 127.8 (2C), 127.2 (2C), 125.2 (2C), 120.0 (2C), 77.9, 77.5, 68.3, 61.6, 54.3, 53.3, 48.5, 47.1, 37.2, 23.9, 15.7, 14.1, 11.1; HRMS-ESI (m/z): [M + Na]+ calcd for C31H45N3O8Na, 730.3099; found, 730.3087 (−1.6 ppm).
General Procedure D, for the Catalytic Hydrogenation of N-(benzyloxy)-Protecting Groups.
Peptides 9a–d (1.0 mmol, 1.0 equiv) were solubilized in EtOH (10 mL) and Pd/C (0.10 mmol,0.10 equiv, 10% w/w) was added to the solution. For 9d, the mixture was cooled down to 0 °C. The flask was placed under an atmosphere of a balloon of H2,gas and the mixture was stirred for 1–18 h (reaction progress monitored by TLC). The reaction mixture was filtered through a pad of cleaned Celite with silica (1% w/w), and the solvent was evaporated to afford the crude product 12a–d, which were further purified by chromatography.
Ac-N(OH)-Gly-N(Me)-Gly-N(OH)-Ethyl-Glycinate (12a).

was prepared according to the general procedure D, using peptide 9a (63 mg, 0.13 mmol, 1.0 equiv). The crude product was purified by chromatography (gradient of hexanes/EtOAc 1:1 to EtOAc 100% and then EtOAc/MeOH 19:1) to afford 12a in a pure form as an uncolored oil (32 mg, 0.11 mmol, 81% yield). Rf = 0.10 (hexanes/EtOAc/AcOH 5:4:1); mp 118–121 °C; IR νmax: 3433, 3193, 2924, 2852, 1741, 1636, 1460, 1401, 1208, 1124, 1023, 822, 562 cm−1; product 12a was characterized as a mixture of cis–trans rotamers in a ratio 2:3 by 1H–13C HSQC: 1H NMR (400 MHz, DMSO-d6, δ) trans rotamer: 10.18 (s, 1H), 9.76 (s, 1H), 4.42 (s, 2H), 4.29 (s, 2H), 4.28 (s, 2H), 4.13 (q, J = 7.1 Hz, 2H), 2.93 (s, 3H), 2.02 (s, 3H), 1.20 (t,, J = 7.1 Hz, 3H); cis rotamer: 10.23 (s, 1H), 9.72 (s, 1H), 4.34 (s, 2H),4.33 (s, 2H), 4.24 (s, 2H), 4.13 (q, J = 7.1 Hz, 2H), 2.81 (s, 3H), 2.02 (s, 3H), 1.20 (t,, J = 7.1 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ) trans rotamer: 172.8, 169.9, 169.6, 168.5, 61.9, 50.1, 49.5, 49.0, 36.4, 20.0, 14.1; cis rotamer: 172.8, 170.4, 169.4, 168.2, 61.9, 50.6, 50.5, 49.1, 35.4, 19.9, 14.1; HRMS-ESI (m/z): [M + Na]+ calcd for C11H19N3O7Na, 328.1115; found, 328.1101 (−4.2 ppm).
Ac-N(OH)-Gly-Gly-N(OH)-Ethyl-Glycinate (12b).

was prepared according to the general procedure D, using peptide 9b (170 mg, 0.36 mmol, 1.0 equiv). The crude product was purified by chromatography (gradient of hexanes/EtOAc 1:1 to EtOAc 100% and then EtOAc/MeOH 19:1) to afford 12b in a pure form as a white solid (82 mg, 0.28 mmol, 79% yield). Rf = 0.05 (hexanes/EtOAc/AcOH 5:4:1); mp 140–142 °C; IR νmax: 3299, 3097, 2911, 2778, 2680, 1743, 1678, 1646, 1593, 1564, 1464, 1411, 1388, 1376, 1257, 1218, 1189, 1027, 823, 786, 729, 615, 541 cm−1; 1H NMR (400 MHz, DMSO-d6, δ): 10.21 (s, 1H), 9.91 (s, 1H), 8.00 (br, 1H), 4.30 (s, 2H), 4.17 (s, 2H), 4.12 (q, J = 7.1 Hz, 2H), 4.08 (d, J = 5.6 Hz, 2H),2.02 (s, 3H), 1.20 (t, J = 7.1 Hz, 3H); 13C{1H} NMR (100 MHz, DMSO-d6, δ): 171.9, 170.9, 168.3, 168.0, 61.3, 51.4, 51.1, 40.4, 20.6, 14.5; HRMS-ESI (m/z): [M + Na]+ calcd for C10H17N3O7Na, 314.0959; found, 314.0958 (+0.3 ppm).
Ac-N(OH)-Gly-Ile-N(OH)-Ethyl-Glycinate (12c).

was prepared according to the general procedure D, using peptide 9c′ (300 mg, 0.69 mmol, 1.0 equiv). The crude product was purified by chromatography (gradient of hexanes/EtOAc 1:1 to EtOAc 100% and then EtOAc/MeOH 19:1) to afford 12c in a pure form as a yellow solid (239 mg, 0.69 mmol, 99% yield). Rf = 0.54 (hexanes/EtOAc/AcOH 5:4:1); (c 1.00, MeOH); mp 97–99 °C; IR νmax: 3276, 3210, 2965, 2–934, 1743, 1629, 1534, 1456, 1375, 1201, 1022, 634, 557 cm−1; 1H NMR (400 MHz, DMSO-d6, δ): 10.24 (s, 1H),9.89 (s, 1H), 7.82 (d, 1H), 5.02 (dd, J = 9.1, 6.0 Hz, 1H), 4.52 (d, J =17.3 Hz, 1H), 4.18 (s, 2H), 4.12 (q, J = 7.0 Hz, 2H), 4.08 (d, J = 17.4 Hz, 1H), 2.01 (s, 3H), 1.87 (m, 1H), 1.44 (m, 1H), 1.19 (t, J = 7.1 Hz, 3H), 1.05 (m, 1H), 0.87 (d, J = 6.8 Hz, 3H), 0.83 (t, J = 7.3 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ): 173.5, 172.2, 169.8, 168.0, 61.8, 53.6, 51.4, 50.0, 36.0, 24.4, 20.1, 15.4, 14.1, 11.0; HRMS-ESI (m/z): [M + H]+ calcd for C25H31N2O6, 348.1765; found, 348.1766 (+0.3 ppm).
Fmoc-N(OH)-Gly-Ile-N(OH)-Ethyl-Glycinate (12d).

was prepared according to the general procedure D, using peptide 9d (15 mg, 0.02 mmol, 1.0 equiv). The crude product was purified by chromatography (gradient of hexanes/EtOAc 4:1 to 2:6) to afford 12d in a pure form as a white powder (7 mg, 0.01 mmol, 62% yield). Rf = 0.38 (hexanes/EtOAc 2:3 v/v); (c 0.50, MeOH); mp 164–169 °C; IR νmax: 3284, 2961, 2923, 2873, 1745, 1711, 1685, 1610, 1541, 1449, 1401, 1376, 1352, 1213, 1111, 1020, 974, 942, 738, 621 cm−1; 1H NMR (400 MHz, DMSO-d6, δ) 10.26 (s, 1H), 9.69 (s, 1H), 7.89 (d, J = 7.8 Hz, 3H), 7.70 (d, J = 7.5 Hz, 2H), 7.42 (t, J = 7.5 Hz, 2H), 7.32 (t, J = 7.5 Hz, 2H), 5.06 (dd, J = 9.3, 5.9 Hz, 1H), 4.51 (d, J = 17.3 Hz, 1H), 4.26 (m, 3H), 4.13 (m, 5H), 1.88 (m, 1H), 1.45 (m, 1H), 1.18 (t, J = 7.1 Hz, 3H), 1.06 (m, 1H), 0.88 (d, J = 6.8 Hz, 3H), 0.82 (t, J = 7.2 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ) 172.0, 169.9, 168.1, 158.1, 143.5 (2C), 141.3 (2C), 127.8 (2C),127.2 (2C), 125.2 (2C), 120.0 (2C), 68.9, 61.8, 54.4, 53.4, 49.9, 46.9, 35.8, 24.6, 15.4, 14.0, 10.9; HRMS-ESI (m/z): [M + Na]+ calcd for C27H33N3O8Na, 550.2160; found, 550.2163 (+0.5 ppm).
N-(((9H-Fluoren-9-yl)methoxy)carbonyl)-N-benzylglycine (SI-1).

Benzylamine (200 mg, 1.87 mmol, 1.0 equiv) and glyoxylic acid monohydrate (344 mg, 3.73 mmol, 2.0 equiv) were solubilized in CH2Cl2 (9.0 mL) and stirred overnight at RT. The solvent was then evaporated and the N-benzyl-N-formylglycine intermediate was characterized by 1H NMR (400 MHz, DMSO-d6, δ): 8.31 (s, 1H),7.35 (m, 5H), 4.50 (s, 2H), and 3.83 (s, 2H). The crude N-benzyl-N-formylglycine was solubilized in a 1.0 M aqueous solution of HCl (9.0 mL) and stirred at reflux overnight. The water was then evaporated under reduced pressure to afford the desired N-benzylglycine hydrochloride (203 mg, 1.0 mmol, 54% yield) as a brown solid, which was used without any purification for the second step. Rf = 0.7 (CH2Cl2/MeOH/AcOH 16:3:1 v/v); N-benzylglycine hydrochloride (200 mg, 0.99 mmol, 1.0 equiv) was solubilized in dioxane (6.0 mL) and the solution was cooled down to 0 °C. A solution of sodium bicarbonate (200 mg, 2.44 mmol, 2.4 equiv) in water (2.0 mL) was then added and the mixture was stirred for 10 min at 0 °C. A solution of (9H-fluoren-9-yl)methyl chloroformate (471 mg, 1.82 mmol, 1.8 equiv) in dioxane (2.0 mL) was added to the mixture, which was slowly warm to RT and stirred overnight. Water was added and the aqueous layer (pH ~ 8) was washed with ethyl acetate (2 × 15 mL). The aqueous layer was acidified with a 5.0 M aqueous solution of HCl to pH ~ 3. The product was extracted with ethyl acetate (3 × 15 mL), the combined organic layers were dried over Na2SO4, filtrated, and the solvent was evaporated under reduced pressure. The crude product was purified by flash chromatography (isocratic solvent mixture of hexanes/EtOAc, 4:1 v/v) to afford product SI-1 in a pure form as a white powder (107 mg, 0.28 mmol, 28% yield over 3 steps). Rf = 0.7 (100% EtOAc); mp 110.0–113.0 (±0.6) °C; IR νmax: 3068, 2923, 2854, 1685, 1451, 1231, 1121, 1002, 956, 758, 738 cm−1; Product SI-1 was characterized as a mixture of cis–trans rotamers in a 1:1 ratio by 1H–13C HSQC: 1H NMR (400 MHz, CD3OD, δ) trans rotamer: 7.77 (d, J = 7.5 Hz, 2H), 7.50 (d, J = 7.5 Hz, 2H), 7.34 (t, J = 5.9 Hz, 2H), 7.23 (m, 5H), 6.96 (m, 2H), 4.53 (d, J = 5.8 Hz, 2H),4.32 (s, 2H), 4.22 (t, J = 5.8 Hz, 1H), 3.85 (s, 2H); cis rotamer: 7.73 (d, J = 7.5 Hz, 2H), 7.59 (d, J = 7.5 Hz, 2H), 7.37 (t, J = 5.8 Hz, 2H),7.29 (m, 5H), 7.16 (m, 2H), 4.49 (s, 2H), 4.47 (d, J = 6.3 Hz, 2H),4.22 (t, J = 5.8 Hz, 1H), 3.73 (s, 2H).
Ethyl Benzylglycinate (SI-2).

Benzylamine (1.00 g, 9.33 mmol, 1.0 equiv) and a solution of ethyl glyoxylate (1.05 g, 10.3 mmol, 1.1 equiv) in toluene (50% w/w) were solubilized in dry ethanol (19.0 mL) and stirred for 1 h at RT. A solution of sodium cyanoborohydride (1.18 g, 18.7 mmol, 2.0 equiv) in dry ethanol (3.0 mL) and acetic acid (0.05 mL, 0.93 mmol, 0.1 equiv) were added to the mixture, which was stirred overnight at RT. The reaction completion was monitored by TLC (Rf = 0.35, EtOAc/hexanes 1:1 v/v). The solvent was then evaporated and the crude was solubilized in dichloromethane (10 mL). The organic layer was washed with water (3 × 10 mL), dried over Na2SO4, and filtrated, and the solvent was evaporated under reduced pressure. The crude product was purified by chromatography (isocratic solvent mixture of hexanes/EtOAc, 7:3 v/v) to afford SI-2 in a pure form as a yellow oil (542 mg, 2.81 mmol, 30% yield). Rf = 0.3 (EtOAc/hexanes 1:1 v/v); IR νmax: 3029, 2980, 2866, 1724, 1453, 1371, 1188, 1142, 1025, 736, 698 cm−1.
Fmoc-Ile-N(OH)-Ethyl-Glycinate (SI-3).

was prepared according to the general procedure D, using peptide 7c (450 mg, 0.83 mmol, 1.0 equiv). The crude product was purified by chromatography (hexanes/EtOAc 7:3 v/v) to afford SI-3 in a pure form as a white solid (339 mg, 0.75 mmol, 90% yield). Rf = 0.3 (hexanes/EtOAc 7:3 v/v); (c 1.00, MeOH); mp 147–149 °C; IR νmax: 3325, 3144, 2968, 1741, 1692, 1605, 1530, 1449, 1211, 1023, 758, 737, 646, 621 cm−1; 1H NMR (400 MHz, CDCl3, δ): 8.35 (br, 1H), 7.75 (d, J = 7.5 Hz, 2H), 7.56 (dd, J = 7.3, 2.9 Hz, 2H), 7.39 (t, J = 7.5 Hz, 2H), 7.29 (t, J = 7.4 Hz, 2H), 5.53 (d, J = 9.1 Hz, 1H),4.82 (t, J = 8.4 Hz, 1H), 4.45 (s, 2H), 4.37 (m, 2H), 4.20 (m, 3H),1.94 (m, 1H), 1.61 (m, 1H), 1.25 (t, J = 7.1 Hz, 3H), 1.20 (m, 1H),1.00 (d, J = 6.7 Hz, 3H), 0.93 (t, J = 7.3 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3, δ): 172.5, 168.4, 157.2, 143.8, 143.7, 141.3 (2C), 127.7 (2C), 127.1 (2C), 125.1 (2C), 120.0 (2C), 67.5, 61.8, 55.4, 49.4, 47.1, 36.4, 24.4, 15.6, 14.1, 11.1; HRMS-ESI (m/z): [M + H]+ calcd for C25H31N2O6, 455.2177; found, 455.2157 (−4.4 ppm).
Ac-N(OH)-Gly-tert-Butyl-Isoleucinate (SI-4).

was prepared according to the general procedure D, using peptide 11c (450 g, 1.15 mmol, 1.0 equiv). The crude product was purified by chromatography (hexanes/EtOAc 6:4 v/v) to afford SI-4 in a pure form as an uncolored oil (344 g, 1.14 mmol, 99% yield). Rf = 0.2 (hexanes/EtOAc 6:4 v/v); (c 1.00, MeOH); IR (dry film) νmax: 3268, 2967, 2934, 1732, 1659, 1637, 1536, 1458, 1393, 1367, 1248, 1142, 1038, 989, 846, 776, 582 cm−1; 1H NMR (400 MHz, DMSO-d6, δ): 9.87 (s, 1H), 6.02 (d, J = 8.4 Hz, 1H), 4.17 (m, 3H),2.01 (s, 3H), 1.74 (m, 1H), 1.41 (s, 3H), 1.37 (m, 1H), 1.18 (m, 1H),0.86 (m, 6H); 13C{1H} NMR (100 MHz, DMSO-d6, δ): 171.9, 170.9, 167.6, 81.2, 57.1, 51.2, 37.3, 28.1 (3C), 25.3, 20.6, 15.9, 11.8; HRMS-ESI (m/z): [M + Na]+ calcd for C14H26N2O5Na, 325.1734; found, 325.1732 (−0.6 ppm).
Ga(Fmoc-Ile-N(O)-Ethyl-Glycinate)3 (SI-5).

Gallium sulfate (28 mg, 0.07 mmol, 3.5 equiv) was solubilized in DMF (0.30 mL). A solution of NH3 in 1,4-dioxane (0.5 M, 0.20 mL) was added to the mixture. A solution of dipeptide SI-3 (10 mg, 0.02 mmol, 1.0 equiv) in DMF (0.30 mL) was added dropwise to the reaction mixture, which was stirred overnight at RT. The resulting crude product was precipitated into cold water, filtrated, and washed with water (3 × 1.0 mL). The solid was then solubilized in CH2Cl2, dried over Na2SO4, and filtrated, and the solvent was removed under reduced pressure to afford product SI-5 as a white powder (29 mg,0.02 mmol, 100% yield) mp 148–151 °C; IR νmax: 3298, 2964, 2928, 1753, 1709, 1589, 1509, 1450, 1219, 1197, 1024, 758, 736 cm−1; The broadness of the peaks in the 1H NMR of product SI-5 suggested that several conformers or several ligand arrangements around the Ga center might co-exist; therefore, the following chemical shifts are representative of the major cis-conformation of the ligand SI-5: 1H NMR (400 MHz, CD3CN, δ): 7.82 (m, 2H), 7.64 (m, 2H), 7.41 (m, 2H), 7.33 (m, 2H), 6.02 (d, J = 8.1 Hz, 1H), 4.87 (d, J = 17.9 Hz, 1H), 4.21 (m, 7H), 1.79 (m, 1H), 1.52 (m, 1H), 1.21 (m, 3H), 1.08 (m, 1H), 0.88 (m, 6H); 13C{1H} NMR (100 MHz, CDCl3, δ): 170.3, 164.7, 156.3, 143.7, 143.6, 141.3 (2C), 127.8 (2C), 127.1 (2C), 125.1 (2C), 120.0 (2C), 67.2, 61.8, 54.0, 53.1, 47.2, 36.9, 24.7, 15.4, 13.9, 10.9; MALDI-TOF (m/z): [M + Na]+ calcd for C75H87N6O18GaNa, 1451.5225; found, 1453. 2538.
Supplementary Material
ACKNOWLEDGMENTS
We are very grateful for the financial support from the National Institutes of Health (NIGMS Grant: R21GM132754 to S.P.R and A.D.R.). The authors also thank Dr. Kari B. Basso at the Mass Spectrometry Research and Education Center from the Department of Chemistry at the University of Florida for the high-resolution mass spectrometry analysis supported by the NIH (S10 OD021758-01A1). The authors also thank Dr. Maren Pink, director of the Molecular Structure Center at the Indiana University Bloomington for the high-resolution crystal structure analysis (X-ray). We thank the NSF’s ChemMat-CARS Sector 15 supported by the Divisions of Chemistry (CHE) and Materials Research (DMR), National Science Foundation, under grant number NSF/CHE-1834750.
Footnotes
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.0c01441.
Tables of the selected results from the C- and N-terminal coupling optimization and epimerization are reported; complete experimental procedures and characterization data including 1H, 13C, NOESY, and HMBC NMR spectra, as well as HPLC chromatograms for measuring the levels of epimerization are available online (PDF)
CCDC 2009540 crystallographic data (for 12c) (CIF)
Complete contact information is available at: https://pubs.acs.org/10.1021/acs.joc.0c01441
The authors declare no competing financial interest.
Contributor Information
Alexis D. Richaud, Department of Chemistry and Biochemistry, Florida Atlantic University, Boca Raton, Florida 33431, United States
Stéphane P. Roche, Department of Chemistry and Biochemistry and Center for Molecular Biology and Biotechnology, Florida Atlantic University, Boca Raton, Florida 33431, United States.
REFERENCES
- (1).(a) Patch JA; Barron AE Mimicry of bioactive peptides via non-natural, sequence-specific peptidomimetic oligomers. Curr. Opin. Chem. Biol 2002, 6, 872–877. [DOI] [PubMed] [Google Scholar]; (b) Yoo B; Kirshenbaum K Peptoid architectures: elaboration, actuation, and application. Curr. Opin. Chem. Biol 2008, 12, 714–721. [DOI] [PubMed] [Google Scholar]; (c) Fowler SA; Blackwell HE Structure-function relationships in peptoids: Recent advances toward deciphering the structural requirements for biological function. Org. Biomol. Chem 2009, 7, 1508–1524. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Sun J; Zuckermann RN Peptoid Polymers: A Highly Designable Bioinspired Material. ACS Nano 2013, 7, 4715–4732. [DOI] [PubMed] [Google Scholar]
- (2).(a) Zuckermann RN; Kerr JM; Kent SBH; Moos WH Efficient method for the preparation of peptoids [oligo(N-substituted glycines)] by submonomer solid-phase synthesis. J. Am. Chem. Soc 1992, 114, 10646–10647. [Google Scholar]; (b) Simon RJ; Kania RS; Zuckermann RN; Huebner VD; Jewell DA; Banville S; Ng S; Wang L; Rosenberg S; Marlowe CK Peptoids: a modular approach to drug discovery. Proc Natl. Acad. Sci. U.S.A 1992, 89, 9367–9371. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Zuckermann RN; Kodadek T Peptoids as potential therapeutics. Curr. Opin. Mol. Ther 2009, 3, 299–307. [PubMed] [Google Scholar]
- (3).(a) Umezawa K; Nakazawa K; Ikeda Y; Naganawa H; Kondo S Polyoxypeptins A and B Produced by Streptomyces: Apoptosis-Inducing Cyclic Depsipeptides Containing the Novel Amino Acid (2S,3R)-3-Hydroxy-3-methylproline. J. Org. Chem 1999, 64, 3034–3038. [DOI] [PubMed] [Google Scholar]; (b) Young DW; Bender A; Hoyt J; McWhinnie E; Chirn G-W; Tao CY; Tallarico JA; Labow M; Jenkins JL; Mitchison TJ; Feng Y Integrating high-content screening and ligand-target prediction to identify mechanism of action. Nat. Chem. Biol 2008, 4, 59–68. [DOI] [PubMed] [Google Scholar]; (c) Oh D-C; Poulsen M; Currie CR; Clardy J Dentigerumycin: a bacterial mediator of an ant-fungus symbiosis. Nat. Chem. Biol 2009, 5, 391–393. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Song F; He H; Ma R; Xiao X; Wei Q; Wang Q; Ji Z; Dai H; Zhang L; Capon RJ Structure revision of the Penicillium alkaloids haenamindole and citreoindole. Tetrahedron Lett. 2016, 57, 3851–3852. [Google Scholar]; (e) Zhu M; Zhang X; Feng H; Dai J; Li J; Che Q; Gu Q; Zhu T; Li D Penicisulfuranols A–F, Alkaloids from the Mangrove Endophytic Fungus Penicillium janthinellum HDN13–309. J. Nat. Prod 2017, 80, 71–75. [DOI] [PubMed] [Google Scholar]; (f) Dewapriya P; Prasad P; Damodar R; Salim AA; Capon RJ Talarolide A, a Cyclic Heptapeptide Hydroxamate from an Australian Marine Tunicate-Associated Fungus, Talaromyces sp. (CMB-TU011). Org. Lett 2017, 19, 2046–2049. [DOI] [PubMed] [Google Scholar]; (g) Shin D; Byun WS; Moon K; Kwon Y; Bae M; Um S; Lee SK; Oh D-C Coculture of Marine Streptomyces sp. With Bacillus sp. Produces a New Piperazic Acid-Bearing Cyclic Peptide. Front. Chem 2018, 6, 1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).(a) Neilands JB Hydroxamic Acids in Nature. Science 1967, 156, 1443–1447. [DOI] [PubMed] [Google Scholar]; (b) Emery T Biosynthesis and mechanism of action of hydroxamate-type siderochromes Microbial Iron Metabolism; Elsevier, 1974; pp 107–123. [Google Scholar]; (c) Palanché T; Blanc S; Hennard C; Abdallah MA; Albrecht-Gary A-M Bacterial Iron Transport: Coordination Properties of Azotobactin, the Highly Fluorescent Siderophore of Azotobacter vinelandii. Inorg. Chem 2004, 43, 1137–1152. [DOI] [PubMed] [Google Scholar]; (d) Hermenau R; Ishida K; Gama S; Hoffmann B; Pfeifer-Leeg M; Plass W; Mohr JF; Wichard T; Saluz H-P; Hertweck C Gramibactin is a bacterial siderophore with a diazeniumdiolate ligand system. Nat. Chem. Biol 2018, 14, 841–843. [DOI] [PubMed] [Google Scholar]
- (5).(a) Bianco A; Zabel C; Walden P; Jung N N-Hydroxy-Amide Analogues of MHC-Class I Peptide Ligands with Nanomolar Binding Affinities. J. Pept. Sci 1998, 4, 471–478. [DOI] [PubMed] [Google Scholar]; (b) Hin S; Zabel C; Bianco A; Jung G; Walden P Cutting Edge: N-Hydroxy Peptides: A New Class of TCR Antagonists. J. Immunol 1999, 163, 2363. [PubMed] [Google Scholar]; (c) Bianco A; Kaiser D; Jung G N-Hydroxy Peptides as Substrates for α-Chymotrypsin. J. Pept. Res 1999, 54, 544–548. [DOI] [PubMed] [Google Scholar]
- (6).(a) Bauer L; Exner O The Chemistry of Hydroxamic Acids and N-Hydroximides. Angew. Chem., Int. Ed 1974, 13, 376–384. [Google Scholar]; (b) Ye Y; Liu M; Kao JLK; Marshall GR Peptide-Bond Modification for Metal Coordination: Peptides Containing Two Hydroxamate Groups. Biopolymers 2003, 71, 489–515. [DOI] [PubMed] [Google Scholar]; (c) Matsumoto K; Ozawa T; Jitsukawa K; Masuda H Synthesis, Solution Behavior, Thermal Stability, and Biological Activity of an Fe(III) Complex of an Artificial Siderophore with Intramolecular Hydrogen Bonding Networks. Inorg. Chem 2004, 43, 8538–8546. [DOI] [PubMed] [Google Scholar]; (d) Geurink P; Klein T; Leeuwenburgh M; van der Marel G; Kauffman H; Bischoff R; Overkleeft H A peptide hydroxamate library for enrichment of metalloproteinases: towards an affinity-based metal-loproteinase profiling protocol. Org. Biomol. Chem 2008, 6, 1244–1250. [DOI] [PubMed] [Google Scholar]
- (7).(a) Ahmad A The Chemistry of α,N-Hydroxyamino Acids. Bull. Chem. Soc. Jpn 1974, 47, 2583–2587. [Google Scholar]; (b) Ottenheijm HCJ; Herscheid JDM N-Hydroxy-α-amino Acids in Organic Chemistry. Chem. Rev 1986, 86, 697–707. [Google Scholar]; (c) Murahashi S-I; Shiota T Short-step synthesis of amino acids and N-hydroxyamino acids from amines. Tetrahedron Lett. 1987, 28, 6469–6472. [Google Scholar]; (d) Novak M; Bonham GA; Mohler LK; Peet KM Acid-Catalyzed Hydrolysis of N-Hydroxyacetanilides: Amide Hydrolysis versus Nitrogen-Oxygen Bond Heterolysis. J. Org. Chem 1988, 53, 3903–3908. [Google Scholar]
- (8).(a) Dupont V; Lecoq A; Mangeot JP; Aubry A; Boussard G; Marraud M Conformational Perturbations Induced by N-Amination and N-Hydroxylation of Peptides. J. Am. Chem. Soc 1993, 115, 8898–8906. [Google Scholar]; (b) Takeuchi Y; Marshall GR Conformational Analysis of Reverse-Turn Constraints by N-Methylation and N-Hydroxylation of Amide Bonds in Peptides and Non-Peptide Mimetics. J. Am. Chem. Soc 1998, 120, 5363–5372. [Google Scholar]
- (9).Crapster JA; Stringer JR; Guzei IA; Blackwell HE Design and Conformational Analysis of Peptoids Containing N-Hydroxy Amides Reveals a Unique Sheet-like Secondary Structure. Biopolymers 2011, 96, 604–616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).(a) Busetti V; Ottenheijm HCJ; Zeegers HJM; Ajó D; Casarin M Crystal and Molecular Structure of N-Acetyl N-Hydroxy α-Amino Acids. Recl. Trav. Chim. Pays-Bas 1987, 106, 151–156. [Google Scholar]; (b) The Chemistry of Hydroxylamines, Oximes, and Hydroxamic Acids In Patai Series: The Chemistry of Functional Groups; Rappoport Z; Liebman JF, Eds.; Wiley: Chichester, England; Hoboken, NJ, 2009; Vol. 1. [Google Scholar]
- (11).Jordan PA; Paul B; Butterfoss GL; Renfrew PD; Bonneau R; Kirshenbaum K Oligo(N-Alkoxy Glycines): Trans Substantiating Peptoid Conformations. Biopolymers 2011, 96, 617–626. [DOI] [PubMed] [Google Scholar]
- (12).Richaud AD; Roche SP Structure-Property Relationship Study of N-(Hydroxy)Peptides for the Design of Self-Assembled Parallel β-Sheets. ChemRxiv 2020, DOI: 10.26434/chemrxiv.12486236.v1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).(a) Ahmad A Syntheses of α-Hydroxyamino Acids from α-Keto Acids. Bull. Chem. Soc. Jpn 1974, 47, 1819–1820. [Google Scholar]; (b) Kolasa T; Chimiak A; Kitowska A Esters of N-Benzyloxyamino Acids. J. Prakt. Chem 1975, 317, 252–256. [Google Scholar]; (c) Polonski T; Chimiak A Nitrones as Intermediates in the Synthesis of N-Hydroxyamino Acid Esters. J. Org. Chem 1976, 41, 2092–2095. [DOI] [PubMed] [Google Scholar]
- (14).(a) Feenstra RW; Stokkingreef EHM; Nivard RJF; Ottenheijm HCJ An Efficient Synthesis of N-Hydroxy-α-Amino Acid Derivatives of High Optical Purity. Tetrahedron Lett. 1987, 28, 1215–1218. [Google Scholar]; (b) Tokuyama H; Kuboyama T; Amano A; Yamashita T; Fukuyama T A Novel Transformation of Primary Amines to N-Monoalkylhydroxylamines. Synthesis 2000, 2000, 1299–1304. [Google Scholar]; (c) Medina SI; Wu J; Bode JW Nitrone Protecting Groups for Enantiopure N-Hydroxyamino Acids: Synthesis of N-Terminal Peptide Hydroxylamines for Chemoselective Ligations. Org. Biomol. Chem 2010, 8, 3405–3417. [DOI] [PubMed] [Google Scholar]
- (15).(a) Kolasa T; Chimiak A Unambiguous Synthesis of N-Hydroxypeptides. Tetrahedron 1977, 33, 3279–3284. [Google Scholar]; (b) Maire P; Blandin V; Lopez M; Vallée Y Straightforward Synthesis of N-Hydroxy Peptides. Synlett 2003, 671–674. [Google Scholar]; (c) Hara Y; Akiyama M Peptide-Based Trihydroxamates as Models for Desferrioxamines. Iron(III)-Holding Properties of Linear and Cyclic N-Hydroxy Peptides with an l-Alanyl-l-Alanyl-N-Hydroxy-β-Alanyl Sequence. Inorg. Chem 1996, 35, 5173–5180. [Google Scholar]; (d) Lawrence J; Cointeaux L; Maire P; Vallée Y; Blandin V N-Hydroxy and N-Acyloxy Peptides: Synthesis and Chemical Modifications. Org. Biomol. Chem 2006, 4, 3125–3141. [DOI] [PubMed] [Google Scholar]; (e) Ye Y; Liu M; Kao JL-F; Marshall GR Novel Trihydroxamate-Containing Peptides: Design, Synthesis, and Metal Coordination. Biopolymers 2006, 84, 472–489. [DOI] [PubMed] [Google Scholar]; (f) Sarnowski MP; Del Valle JR N-Hydroxy peptides: solid-phase synthesis and β-sheet propensity. Org. Biomol. Chem 2020, 18, 3690–3696. See also reference [5a]. [DOI] [PubMed] [Google Scholar]
- (16).See Supporting Information for full experimental details.
- (17).(a) Birnara C; Kessler VG; Papaefstathiou GS Mononuclear Gallium(III) Complexes Based on Salicylaldoximes: Synthesis, Structure and Spectroscopic Characterization. Polyhedron 2009, 28, 3291–3297. [Google Scholar]; (b) Deshmukh RD Preparation of Semiconductor Films. U.S. Patent WO2015/113733A1, 2015. See also ref 4c.
- (18).(a) Valeur E; Bradley M Amide bond formation: beyond the myth of coupling reagents. Chem. Soc. Rev 2009, 38, 606–631. [DOI] [PubMed] [Google Scholar]; (b) Dunetz JR; Magano J; Weisenburger GA Large-Scale Applications of Amide Coupling Reagents for the Synthesis of Pharmaceuticals. Org. Process Res. Dev 2016, 20, 140–177. [Google Scholar]
- (19).Asakura T; Okonogi M; Nakazawa Y; Yamauchi K Structural Analysis of Alanine Tripeptide with Antiparallel and Parallel β-Sheet Structures in Relation to the Analysis of Mixed β-Sheet Structures in Samia cynthia ricini Silk Protein Fiber Using Solid-State NMR Spectroscopy. J. Am. Chem. Soc 2006, 128, 6231–6238. [DOI] [PubMed] [Google Scholar]
- (20).(a) Nowick JS; Chung DM; Maitra K; Maitra S; Stigers KD; Sun Y An Unnatural Amino Acid that Mimics a Tripeptide β-Strand and Forms β-Sheetlike Hydrogen-Bonded Dimers. J. Am. Chem. Soc 2000, 122, 7654–7661. [Google Scholar]; (b) Phillips ST; Rezac M; Abel U; Kossenjans M; Bartlett PA “@-Tides”: The 1,2-Dihydro-3(6H)-pyridinone Unit as a β-Strand Mimic. J. Am. Chem. Soc 2002, 124, 58–66. [DOI] [PubMed] [Google Scholar]; (c) Laungani AC; Slattery JM; Krossing I; Breit B Supramolecular Bidentate Ligands by Metal-Directed in situ Formation of Antiparallel β-Sheet Structures and Application in Asymmetric Catalysis. Chem. - Eur. J 2008, 14, 4488–4502. [DOI] [PubMed] [Google Scholar]; (d) Ong ZY; Gao SJ; Yang YY Short Synthetic β-Sheet Forming Peptide Amphiphiles as Broad Spectrum Antimicrobials with Antibiofilm and Endotoxin Neutralizing Capabilities. Adv. Funct. Mater 2013, 23, 3682–3692. [Google Scholar]
- (21).For all peptide–peptoids studied, NMR data were recorded in a range of concentrations between 5 to 10 mM. The chemical shift deviations of amides and N-hydroxy amides were insignificant above 30 mM, which suggest that no aggregation occurs within the range of concentration studied.
- (22).(a) Sui Q; Borchardt D; Rabenstein DL Kinetics and Equilibria of Cis/Trans Isomerization of Backbone Amide Bonds in Peptoids. J. Am. Chem. Soc 2007, 129, 12042–12048. [DOI] [PubMed] [Google Scholar]; (b) Laursen JS; Engel-Andreasen J; Fristrup P; Harris P; Olsen CA Cis–Trans Amide Bond Rotamers in β-Peptoids and Peptoids: Evaluation of Stereoelectronic Effects in Backbone and Side Chains. J. Am. Chem. Soc 2013, 135, 2835–2844. [DOI] [PubMed] [Google Scholar]; (c) Stringer JR; Crapster JA; Guzei IA; Blackwell HE Construction of Peptoids with All Trans-Amide Backbones and Peptoid Reverse Turns via the Tactical Incorporation of N-Aryl Side Chains Capable of Hydrogen Bonding. J. Org. Chem 2010, 75, 6068–6078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (23).Wüthrich K NMR of Proteins and Nucleic Acids; Wiley: New York, 1986. [Google Scholar]
- (24).(a) Abraham MH; Abraham RJ; Byrne J; Griffiths L NMR Method for the Determination of Solute Hydrogen Bond Acidity. J. Org. Chem 2006, 71, 3389–3394. [DOI] [PubMed] [Google Scholar]; (b) Abraham MH; Abraham RJ; Acree WE; Aliev AE; Leo AJ; Whaley WL An NMR Method for the Quantitative Assessment of Intramolecular Hydrogen Bonding; Application to Physicochemical, Environmental, and Biochemical Properties. J. Org. Chem 2014, 79, 11075–11083. [DOI] [PubMed] [Google Scholar]
- (25).(a) Olivato PR; Domingues NLC; Mondino MG; Lima FS; Zukerman-Schpector J; Rittner R; Colle MD Stereochemical and electronic interaction studies of some N-methoxy-N-methyl-2-[(4′-substituted)phenylsulfinyl]propanamides. J. Mol. Struct 2008, 892, 360–372. [Google Scholar]; (b) Olivato PR; Domingues NLC; Reis AKCA; Vinhato E; Mondino MG; Zukerman-Schpector J; Rittner R; Colle MD Spectroscopic and theoretical studies of some N-methoxy-N-methyl-2-[(4′-substituted) phenylsulfonyl]propanamides. J. Mol. Struct 2009, 935, 60–68. [Google Scholar]
- (26).(a) Kaur D; Kohli R Hydrogen bond cooperativity in dimers of hydroxamic acids. Int. J. Quantum Chem 2011, 111, 2931–2943. [Google Scholar]; (b) Gupta SP QSAR Studies on Hydroxamic Acids: A Fascinating Family of Chemicals with a Wide Spectrum of Activities. Chem. Rev 2015, 115, 6427–6490. [DOI] [PubMed] [Google Scholar]
- (27).Shin SBY; Kirshenbaum K Conformational Rearrangements by Water-Soluble Peptoid Foldamers. Org. Lett 2007, 9, 5003–5006. see also reference 10. [DOI] [PubMed] [Google Scholar]
- (28).(a) Madison V; Schellman J Optical Activity of Polypeptides and Proteins. Biopolymers 1972, 11, 1041–1076. [DOI] [PubMed] [Google Scholar]; (b) Sreerama N; Woody RW Computation and Analysis of Protein Circular Dichroism Spectra Methods Enzymol.; Elsevier, 2004; Vol. 383, pp 318–351. [DOI] [PubMed] [Google Scholar]; (c) Kung VM; Cornilescu G; Gellman SH Impact of Strand Number on Parallel β-Sheet Stability. Angew. Chem., Int. Ed 2015, 54, 14336–14339. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Guo Y; Wang S; Du H; Chen X; Fei H Silver Ion-Histidine Interplay Switches Peptide Hydrogel from Antiparallel to Parallel β-Assembly and Enables Controlled Antibacterial Activity. Biomacromolecules 2019, 20, 558–565. [DOI] [PubMed] [Google Scholar]
- (29).(a) Smith AM; Williams RJ; Tang C; Coppo P; Collins RF; Turner ML; Saiani A; Ulijn RV Fmoc-Diphenylalanine Self Assembles to a Hydrogel via a Novel Architecture Based on π–π Interlocked β-Sheets. Adv. Mater 2008, 20, 37–41. [Google Scholar]; (b) Ryan DM; Doran TM; Anderson SB; Nilsson BL Effect of C -Terminal Modification on the Self-Assembly and Hydrogelation of Fluorinated Fmoc-Phe Derivatives. Langmuir 2011, 27, 4029–4039. [DOI] [PubMed] [Google Scholar]; (c) Wang Y; Qi W; Wang J; Li Q; Yang X; Zhang J; Liu X; Huang R; Wang M; Su R; He Z Columnar Liquid Crystals Self-Assembled by Minimalistic Peptides for Chiral Sensing and Synthesis of Ordered Mesoporous Silica. Chem. Mater 2018, 30, 7902–7911. [Google Scholar]
- (30).(a) Stevens ES; Sugawara N; Bonora GM; Toniolo C Conformational Analysis of Linear Peptides. 3. Temperature Dependence of NH Chemical Shifts in Chloroform. J. Am. Chem. Soc 1980, 102, 7048–7050. [Google Scholar]; (b) Wang CK; Northfield SE; Colless B; Chaousis S; Hamernig I; Lohman R-J; Nielsen DS; Schroeder CI; Liras S; Price DA; Fairlie DP; Craik DJ Rational Design and Synthesis of an Orally Bioavailable Peptide Guided by NMR Amide Temperature Coefficients. Proc. Natl. Acad. Sci. U.S.A 2014, 111, 17504–17509. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Farley KA; Che Y; Navarro-Vázquez A; Limberakis C; Anderson D; Yan J; Shapiro M; Shanmugasundaram V; Gil RR Cyclic Peptide Design Guided by Residual Dipolar Couplings, J-Couplings, and Intramolecular Hydrogen Bond Analysis. J. Org. Chem 2019, 84, 4803–4813. [DOI] [PubMed] [Google Scholar]
- (31).Seemingly, only one example of temperature dependence coefficients for N-(hydroxyl)peptides has been ″reported, see reference [5c]
- (32).(a) Andersen NH; Neidigh JW; Harris SM; Lee GM; Liu Z; Tong H Extracting Information from the Temperature Gradients of Polypeptide NH Chemical Shifts. 1. The Importance of Conformational Averaging. J. Am. Chem. Soc 1997, 119, 8547–8561. [Google Scholar]; (b) Mahalakshmi R; Raghothama S; Balaram P NMR Analysis of Aromatic Interactions in Designed Peptide β-Hairpins. J. Am. Chem. Soc 2006, 128, 1125–1138. [DOI] [PubMed] [Google Scholar]; (c) Körling M; Geyer A Beyond Natural Limitations: Long-Range Influence of Non-Natural Flexible and Rigid β-Turn Mimetics in a Native β-Hairpin Motif: Non-Natural β-Turn Mimetics in a Native β-Hairpin Motif. Eur. J. Org. Chem 2015, 2015, 6448–6457. [Google Scholar]; (d) Trainor K; Palumbo JA; MacKenzie DWS; Meiering EM Temperature Dependence of NMR Chemical Shifts: Tracking and Statistical Analysis. Protein Sci. 2020, 29, 306–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).(a) Kang CW; Sarnowski MP; Ranatunga S; Wojtas L; Metcalf RS; Guida WC; Del Valle JR β-Strand mimics based on tetrahydropyridazinedione (tpd) peptide stitching. Chem. Commun 2015, 51, 16259–16262. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Sarnowski MP; Kang CW; Elbatrawi YM; Wojtas L; Del Valle JR Peptide N-Amination Supports β-Sheet Conformations. Angew. Chem., Int. Ed 2017, 56, 2083–2086. [DOI] [PubMed] [Google Scholar]
- (34).(a) Langenhan JM; Guzei IA; Gellman SH Parallel sheet secondary structure in β-peptides. Angew. Chem., Int. Ed 2003, 42, 2402–2405. [DOI] [PubMed] [Google Scholar]; (b) Freire F; Fisk JD; Peoples AJ; Ivancic M; Guzei IA; Gellman SH Diacid Linkers That Promote Parallel β-Sheet Secondary Structure in Water. J. Am. Chem. Soc 2008, 130, 7839–7841. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Freire F; Gellman SH Macrocyclic design strategies for small, stable parallel β-sheet scaffolds. J. Am. Chem. Soc 2009, 131, 7970–7972. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Almeida AM; Li R; Gellman SH Parallel β-Sheet Secondary Structure Is Stabilized and Terminated by Interstrand Disulfide Cross-Linking. J. Am. Chem. Soc 2012, 134, 75–78. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Calvelo M; Guerra A; Amorin M; Garcia-Fandino R; Lamas A; Granja JR Parallel versus antiparallel β-sheet structure in cyclic peptide hybrids containing γ- or δ-cyclic amino acids. Chem. -Eur. J 2020, 26, 5846–5858. [DOI] [PubMed] [Google Scholar]
- (35).Menges F “Spectragryph - optical spectroscopy software″, Version 1.2.13, 2019, http://www.effemm2.de/spectragryph/ (accessed May 11, 2020). [Google Scholar]
- (36).(a) Kruijtzer JAW; Hofmeyer LJF; Heerma W; Versluis C; Liskamp RMJ Solid-Phase Syntheses of Peptoids Using Fmoc-Protected N-Substituted Glycines: The Synthesis of (Retro)Peptoids of Leu-Enkephalin and Substance P. Chem. - Eur. J 1998, 4, 1570–1580. [Google Scholar]; (b) Tijhuis MW; Herscheid JDM; Ottenheijm HCJ A Practical Synthesis of N-Hydroxy-α-Amino Acid Derivatives. Synthesis 2002, 1980, 890–893. [Google Scholar]; (c) Wolfe S; Akuche C; Ro S; Wilson M-C; Kim C-K; Shi Z 5-Hydroxy[1,2]Oxazinan-3-Ones as Potential Carbapenem and D-Ala-D-Ala Surrogates. Can. J. Chem 2003, 81, 915–936. [Google Scholar]; (d) Nuti E; Orlandini E; Nencetti S; Rossello A; Innocenti A; Scozzafava A; Supuran CT Carbonic Anhydrase and Matrix Metalloproteinase Inhibitors. Inhibition of Human Tumor-Associated Isozymes IX and Cytosolic Isozyme I and II with Sulfonylated Hydroxamates. Bioorg. Med. Chem 2007, 15, 2298–2311. [DOI] [PubMed] [Google Scholar]; (e) Kaminker R; Anastasaki A; Gutekunst WR; Luo Y; Lee S-H; Hawker CJ Tuning of Protease Resistance in Oligopeptides through N-Alkylation. Chem. Commun 2018, 54, 9631–9634. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
