Abstract
This research aimed to develop new tumor targeted theranostic agents taking advantage of the similarities in coordination chemistry between technetium and rhenium. A γ-emitting radioactive isotope of technetium is commonly used in diagnostic imaging, and there are two β– emitting radioactive isotopes of rhenium that have the potential to be of use in radiotherapy. Variants of the 6-hydrazinonicotinamide (HYNIC) bifunctional ligands have been prepared by appending thioamide functional groups to 6-hydrazinonicotinamide to form pyridylthiosemicarbazide ligands (SHYNIC). The new bidentate ligands were conjugated to the tumor targeting peptides Tyr3-octreotate and cyclic-RGD. The new ligands and conjugates were used to prepare well-defined {M=O}3+ complexes (where M = 99mTc or natRe or 188Re) that feature two targeting peptides attached to the single metal ion. These new SHYNIC ligands are capable of forming well-defined rhenium and technetium complexes and offer the possibility of using the 99mTc imaging and 188/186Re therapeutic matched pairs.
Short abstract
New pyridylthiosemicarbazide ligands conjugated to the tumor targeting peptides octreotate and cyclic-RGD peptides were used to form {M=O}3+ complexes (where M = Tc or Re) that feature two targeting peptides attached to the single metal ion. These modified “HYNIC” ligands are capable of forming well-defined rhenium and technetium complexes offering the possibility of using radionuclides of technetium and rhenium as imaging and therapeutic matched pairs.
Introduction
The technetium-99m isotope has excellent properties for detection with single photon emission computed tomography (SPECT) due to its low energy and nonparticulate gamma-ray emission (t1/2 = 6.01 h, Emax = 141 keV γ-ray emission, λ < 10 pm). Despite recent concerns over production related shortages of technetium-99m and the advent of positron emission tomography technetium-99m retains its importance to nuclear medicine to the extent that the isotope is used in over 80% of nuclear imaging procedures worldwide.1 The heavier third row Group VII congener, rhenium, has an ionic radius similar to technetium due to the lanthanide contraction. Technetium and rhenium display similar coordination chemistry often resulting in essentially isostructural technetium and rhenium complexes. It is common for technetium and rhenium complexes to be essentially isostructural. There are two isotopes of rhenium that are of potential use in targeted radiotherapeutics, rhenium-186 (t1/2 = 89.3 h, Emax = 1.07 MeV β– particle emission, 137 keV γ-ray emission) and rhenium-188 ((t1/2 = 16.9 h, Emax = 2.12 MeV β– particle emission, 155 keV γ-ray emission). The similar coordination chemistry of technetium and rhenium offers the possibility of using their radioisotopes as an imaging (99mTc) and therapeutic (186/188Re) matched pair using a single targeted ligand to form essentially isostructural complexes.2−6
One approach to targeted imaging and therapy is to incorporate appropriate metal radionuclides into coordination complexes that are attached to biological targeting vectors such as tumor targeting peptides, antibodies or antibody fragments. Peptides that feature the −RGD- (arginine-glycine-aspartic acid) fibronectin fragment such as the cyclic-RGDfK pentapeptide (cRGDfK) bind to αvβ3 integrin receptors that are overexpressed in certain invasive tumors including osteosarcomas, glioblastoma, melanomas, and breast cancer, and can be used to selectively target tumor cells.7−16 Metabolically stabilized somatostatin analogues such as octreotide and octreotate bind to somatostatin subtype 2 receptors (sstr2) that are overexpressed in many types of neuroendocrine tumors compared to relatively low levels of expression in other tissues and organs.17−21
Tumor targeting technetium based imaging agents can be prepared using 6-hydrazinonicotinamide (HYNIC) derivatives conjugated to targeting molecules as ligands to form technetium(III) diazenido complexes.22−31 The HYNIC ligand forms remarkably stable technetium complexes, and manipulation of the carboxylate functional group to attach a variety of targeting molecules is generally straightforward. HYNIC binds to technetium through the terminal hydrazine nitrogen but probably forms bidentate complexes through coordination to the pyridyl nitrogen.29,32 In all the crystallographically characterized technetium and rhenium complexes with one or more HYNIC-like ligands, such as 2-hydrazinopyridine, the pyridyl nitrogen is also coordinated to the metal center.33 A variety of coligands such as tricine, nicotinic acid, EDDA, and phosphines (TPPTS, TPPDS, TPPMS = tri/bi/sodium triphenylphosphine tri/di/monosulfonate) are required to complete the coordination sphere and stabilize the metal oxidation state, and this leads to a high degree of uncertainty in the exact nature of the primary coordination sphere as well as challenges in ensuring structural homogeneity in the formulated product. Variation of the coligand can modify the in vivo metabolism and excretion.34 A well-established “ternary ligand system” involves combining the HYNIC ligand with tetradentate tricine and monodentate trisodium 3,3′,3″-phosphanetriyltris(benzenesulfonate) (TPPTS) coligands, but the possibility of forming multiple isomers adds complications.33,35−39
Despite the superficial similarities in coordination chemistry between technetium and rhenium extrapolation of the HYNIC strategy to radioactive rhenium isotopes is challenging presumably due to their differences in kinetic lability and redox chemistry.33 Some of the difficulty in isolating pure Re-HYNIC-peptide conjugates can be understood by considering the reaction of [ReO4]− with 2-hydrazinopyridine (used as model for HYNIC).33,40 This reaction results in relatively complex coordination chemistry due to the ability of the pyridylhydrazine derived ligands to coordinate as either monodentate or bidentate ligands and the existence of protic equilibria as well as the formation of complexes where two pyridylhydrazine derived units are coordinated to the rhenium (Figure 1).26,33,40−43
Figure 1.
(a) 6-Hydrazinonicotinic acid. (b) Metal complex (M = Tc or Re) with 6-hydrazinonicotinimide (HYNIC) acting as a monodentate ligand. It is necessary to complete the coordination sphere with coligands (L). (c) Metal complex (M = Tc or Re) with 6-hydrazinonicotinimide (HYNIC) acting as a bidentate ligand. (d) Pyridylphenylthiocarbazide (SHYNIC) ligand. (e) ReV-oxo complex featuring two SHYNIC ligands.44
Modification of the terminal hydrazinic nitrogen of hydrazinopyridine to incorporate an additional thiourea functional group results in a ligand system that is capable of forming well-defined, very stable complexes with {ReVO}3+ cores while retaining the bioconjugation possibilities well established for HYNIC.44,45 A preliminary communication reported the structural characterization of a ReV-oxo complex featuring two pyridylphenylthiocarbazide (SHYNIC) ligands (Figure 1).44 In this manuscript we extend this concept by synthesizing a family of different substituted pyridylthiosemicarbazide ligands with carboxylate or ester functional groups that were used to tether octreotate and cyclic-RGD peptides to the ligands. The new ligands were used to prepare {MO}3+ complexes (where M = Tc or Re) that feature two targeting peptides attached to the single metal ion. These modified HYNIC ligands are capable of forming well-defined rhenium and technetium complexes and offer the possibility of using the two radionuclides as imaging and therapeutic matched pairs.
Results and Discussion
Synthesis of H2L1–4, and Their Ester Derivatives, H2L1–4(OMe) and {ReO}3+ Complexes
Synthesis of 6-hydrazinonictonic acid (HYNIC), 2, required treatment of 6-chloronicotinic acid (1) with aqueous hydrazine.47 Ligands H2L1 to H2L4 were prepared by reaction of either the ethyl, tert-butyl, phenyl, or nitrophenyl isothiocyanate with 6-hydrazinonictonic acid (1) in anhydrous N,N-dimethylacetamide (DMA) (Scheme 1).
Scheme 1. Synthesis of Ligands H2L1–H2L4 and Their Methyl Ester Derivatives H2L1(OMe)–H2L4(OMe).
The rhenium complexes of the methyl ester derivatives of H2L1–3 complexes, [ReO(HL1–3(OMe))2]+, can be prepared by reaction of the either trans-[ReOCl3(PPh3)2] or [tBu4N][ReOCl4] with the two equivalents of ligand (Scheme 2). The IR spectra for the three complexes, [ReO(H2L1–3(OMe))2]+, display medium intensity bands at ν̅ 960–963 cm–1 characteristic of Re=O stretches.48 Bands, which occur between ν̅ 1553 and 1557 cm–1 due to carbonyl stretching of the ester functional group, shift approximately 150 cm–1 lower in energy when compared to the metal-free ligands.
Scheme 2. Synthesis of Rhenium Complexes [ReO(HL1-3)(OMe))2]+.
Analysis of the complexes by 1H NMR data reveals that the two coordinated ligands are magnetically equivalent, with three resonances at δ 8.63, 8.25, and 7.86 ppm corresponding to the six pyridinyl CH protons for [ReO(HL1(OMe))2]+, and similar resonances for the phenyl and tert-butyl derivatives. The pyridine proton which is closest to the rhenium ion shifts from δ 6.55 ppm in free ligand to δ 7.86 ppm in the complex. The methyl ester functional group gives rise to singlets at δ 3.89 (DMSO-d6), 3.86 (CHCl3-d), and 3.92 (DMSO-d6) for complexes [ReO((HL1,2,3)(OMe))2]+ respectively. The [ReO(HL1(OMe))2]+ complex was stable to cysteine and histidine challenge experiments with very little decomposition evident (<5%), as detected by analysis by HPLC and UV/vis spectroscopy, when incubated at 37 °C for 24 h in the presence of a 100-fold excess of cysteine and histidine.
Red crystals of [ReO(HL1(OMe))2]CF3CO2 suitable for X-ray crystallographic analysis were obtained by evaporation of a solution of the compound that had been purified by semipreparative HPLC using an aqueous/CH3CN mobile phase with 0.1% trifluoracetic acid (Figure 2a). The compound crystallizes in the triclinic space group, P1̅, and the rhenium ion is in a distorted square pyramidal environment with the oxo group in the apical position relative to the pseudo basal plane of two five-membered chelate rings. Each thiocarbahydrazide functional group is doubly deprotonated and serves as a dianionic ligand fragment and a N,N/S,S trans configuration about the Re-oxo. The selective formation of the N,N/S,Strans geometric isomer presumably reflects a strong “trans effect”, although steric requirements may also play some role.49−51 Protonation of the pyridyl nitrogen atom in each ligand results in each ligand having a single negative charge and resulting an overall monocationic complex. The two pyridinium protons and the hydrogen atoms of the ethylamino functional group are involved in hydrogen bonds interactions leading to a hydrogen bonded centrosymmetric dimer (Figure 2b) with the two remaining H-bond donors capped by water molecules.
Figure 2.
(a) ORTEP representation of [ReO(HL1(OMe))2]+ (50% probability ellipsoids). The trifluoroacetate counterion and hydrogen atoms (except those bound to nitrogen) are omitted. (b) Representation of hydrogen bonded centrosymmetric dimer with the two remaining H-bond donors capped by water molecules.
The Re–O1 bond distance (1.679(3) Å) is typical for five-coordinate, rhenium(V)-monooxo complexes and consistent with IR spectroscopy (Re=O, ν̅ 963 cm–1).52−55 The N2–C6 and N6–C14 bond lengths (1.287(6) Å) are significantly shorter than the other bonds within the chelate ring suggesting significant sp2 hybridized C=N bond character. The Re–N bond distances average 2.05 Å and are marginally shorter than typical Re–N bonds (ca. 2.15–2.18 Å) suggesting some degree of multiple bond character.51,56 The Re–N1–N2 and Re–N5–N6 bond angles average 124°, suggesting an approximately sp2 hybridized nitrogen. The Re–S bond distances average 2.29 Å and are similar to the Re–S bond distances in rhenium(V) complexes with amino thiolate ligands and Re–S bond distances in rhenium complexes with thiosemicarbazonato ligands.51,57,58
Table 1. Selected Bond Lengths (Å) and Angles (deg) for the Rhenium Complex for [ReO(HL1(OMe))2]CF3CO2a.
| Bond Lengths | |||||
| Re–O(1) | 1.679(3) | Re–S(1) | 2.2820(11) | Re–S(2) | 2.2917(11) |
| N(1)–C(1) | 1.350(6) | Re–N(1) | 2.049(4) | Re–N(5) | 2.050(4) |
| C(6)–N(3) | 1.343(6) | N(1)–N(2) | 1.422(5) | N(5)–N(6) | 1.418(5) |
| C(14)–N(7) | 1.350(6) | N(2)–C(6) | 1.287(6) | N(6)–C(14) | 1.287(6) |
| N(5)–C(9) | 1.356(6) | C(6)–S(1) | 1.784(5) | C(14)–S(2) | 1.769(4) |
| Bond Angles | |||||
| O(1)–Re–S(1) | 115.44(11) | N(1)–Re–S(1) | 80.74(11) | ||
| O(1)–Re–S(2) | 116.75(11) | N(5)–Re–S(2) | 80.23(11) | ||
| O(1)–Re–N(1) | 101.73(15) | N(5)–Re–S(1) | 89.71(11) | ||
| O(1)–Re–N(5) | 102.61(15) | N(1)–Re–S(2) | 87.98(11) | ||
| N(1)–Re–N(5) | 155.64(15) | C(1)–N(1)–Re | 125.0(3) | ||
| S(1)–Re–S(2) | 127.80(4) | C(9)–N(5)–Re | 126.2(3) | ||
| Torsion Angles | |||||
| N(4)–C(1)–N(1)–Re | 169.5(3) | N(6)–C(14)–N(7)–C(15) | 13.8(7) | ||
| N(8)–C(9)–N(5)–Re | 173.3(3) | N(2)–C(6)–N(3)–C(7) | 5.1(7) | ||
| N(4)–C(1)–N(1)–N(2) | –10.5(6) | N(2)–C(6)–S(1)–Re | 3.7(4) | ||
| N(8)–C(9)–N(5)–N(6) | –7.8(6) | N(6)–C(14)–S(2)–Re | 4.1(4) | ||
| C(2)–C(1)–N(1)–N(2) | 167.1(4) | N(3)–C(6)–S(1)–Re | –174.7(3) | ||
| C(10)–C(9)–N(5)–N(6) | 168.4(4) | N(7)–C(14)–S(2)–Re | –175.1(3) |
Trifluoroacetate counterion bond lengths and angles are not provided.
Table 2. Summary of Crystal Data and Structure Refinement for [ReO(HL1(OMe))2]CF3CO2a.
| data collection | compound details | data collection | compound details |
|---|---|---|---|
| empirical formula | C20H26N8O5ReS2.H2O.CF3CO2 | V (Å3) | 1551.05(13) |
| formula weight | 839.84 | Z | 2 |
| crystal size (mm3) | 0.26 × 0.10 × 0.022 | Dcalc. (Mg m–3) | 1.797 |
| crystal system | triclinic | μ (mm–1) | 9.599 |
| space group | P1̅ | F(000) | 828 |
| T (K) | 130.00(10) | reflections measured | 10363 |
| λ (Å) | 1.54184 | independent reflections | 5568 [Rint = 0.0449] |
| a (Å) | 10.9154(4) | final R indices [I > 2σ(I)] | R1 = 0.0329 |
| b (Å) | 11.1201(6) | wR(F2) = 0.0794 | |
| c (Å) | 13.2651(7) | final R indices (all data) | R1 = 0.0398 |
| α (deg) | 86.355(4) | wR(F2) = 0.0831 | |
| β (deg) | 74.965(4) | goodness-of fit on F2 | 1.027 |
| γ (deg) | 87.273(4) |
Crystals were grown from a concentrated solution of the complex in methanol.
The potential of the substituted pyridylthiosemicarbazide (SHYNIC) ligands H2L1–4 to be modified with amino acids using standard solid phase peptide synthesis techniques was first accomplished by attaching l-lysine to H2L1 to give H2L1(Lys). The doubly N-protected lysine derivative, Nα-tBoc-Nε-Fmoc-l-Lys, was immobilized on chlorotrityl resin, and the Nε-Fmoc group was removed by treatment with piperidine. The ligand, H2L1, was added to the resin in a mixture of DMF followed by the coupling agent HATU (HATU = 1-[bis(dimethylamino)methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium-3-oxid hexafluorophosphate) in the presence of N,N-diisopropylethylamine (DIPEA). The product was cleaved from the resin using trifluoroacetic acid that also resulted in the deprotection of the Nα-t-butoxycarbonyl group (Scheme 3). This lysine amino acid conjugate provides an amino acid with an appended chelator, which with appropriate protecting groups, could be incorporated into biological targeting molecules with total site specificity via solid-phase peptide synthesis.59−63
Scheme 3. “On-Resin” Formation of H2L1(Lys) and the Corresponding Rhenium-oxo Complex, [ReO(HL1(Lys))2]+.
The rhenium complex, [ReO(HL1(Lys))2]+, was prepared by adding either trans-[ReOCl3(PPh3)2] or [tBuN][ReOCl4] suspended in DMF to the reaction mixture, while the ligand remained immobilized on the resin. Performing the complexation while the ligand remained immobilized on the resin and with the amino group still protected ensured that the functional groups of the lysine did not complicate the coordination chemistry. When green trans-[ReOCl3(PPh3)2] is used as the starting material the green colored suspension gradually changes to colorless, and the resin beads turn dark red indicative of the formation of [ReO(HL1(Lys))2]+. After being stirred at room temperature, the resin was washed with dimethylformamide and dichloromethane and cleaved off the resin with a 10% trifluoroacetic acid/dichloromethane mixture (Scheme 3). Analysis of [ReO(HL1(Lys))2]+ complex by electrospray ionization mass spectrometry (ESI-MS) reveals the expected peaks. Analysis by 1H NMR shows the singlet attributed to the aromatic proton on the pyridine ring (pyH2) shifts upon coordination to the metal center from δ 8.48 in H2L1(Lys) to 8.58 ppm in [ReO(HL1(Lys))2]+. The four downfield 13C{1H} NMR signals in [ReO(HL1(Lys))2]+, 173.0 (CO2H), 169.0 (N=CS), 165.4 (CONH), and 163.6 (pyC6), were assigned using HSQC and HMBC techniques.
Synthesis of Peptide-Conjugated Ligands, H2L1–3(cRGDfK) and H2L1–3(Tyr3-Octreotate), and Corresponding Rhenium Complexes [ReO((HL1–3)(cRGDfK))2] and [ReO((HL1–3)(Tyr3-Octreotate))2]
The cyclic pentapeptide, cRGDfK, was prepared using standard solid phase peptide synthesis techniques with Fmoc (Fmoc = 9-fluorenylmethyloxycarbonyl) protected amino acids, using HATU/DIPEA coupling methodology on chlorotrityl resin. The Fmoc protecting groups were removed with 20% piperidine in DMF on the solid phase, followed by cleavage from the resin and cyclization using a small modification of published procedures.64 The ligands, H2L1–3, were conjugated to cRGDfK using standard peptide coupling conditions (HATU, DIPEA) to give H2L1–3(cRGDfK) (Scheme 4). The new conjugates were purified by semipreparative RP-HPLC and characterized by electrospray mass spectrometry and analytical HPLC. The reaction for the ethyl-SHYNIC derivative, H2L1(cRGDfK), resulted in a higher isolated yield (80%) than H2L2(cRGDfK) (24%) and H2L3(cRGDfK) (55%). Analysis of aqueous solutions of H2L1–3(cRGDfK) by RP-HPLC revealed no degradation over a 24 h.
Scheme 4. Synthesis of H2L1-3(cRGDfK) and [ReO(HL1-3(cRGDfK))2]+.
The {ReO}3+ complexes of H2L1–3(cRGDfK) were prepared by adding [ReOCl4]− in DMF at ambient temperature followed by isolation and purification using semipreparative RP-HPLC (Scheme 4). Analysis of the complexes by ESI-MS revealed the 2+ molecular ion peaks at m/z 926.343, 954.372, and 974.371 for [ReO(HL1(cRGDfK))2]+, [ReO(HL2(cRGDfK))2]+, and [ReO(HL3(cRGDfK))2]+ respectively.
An intramolecular disulfide bridge between the second and seventh cysteine residues in Tyr3-octreotate improves the metabolic stability of the peptide, and this disulfide is often introduced by oxidation of the linear octapeptide with 2,2′-dithiodipyridine. Unfortunately the bioconjugation of ligands H2L1–3 to Tyr3-octreotate was complicated by degradation of the pyridylthiocarbazide (SHYNIC) ligands (H2L1–3) in the presence of the two cysteine thiol containing residues in the linear peptide, leading to loss of H2S identified in the ESI-MS by a loss of 34 atomic mass units. This loss of H2S from the ligand was most prominent during reactions attempting intramolecular oxidation of thiol groups in the two cysteine residues. The loss of H2S results in the formation of a carbodiimide form of the SHYNIC ligands. The formation of carbodiimides from thioureas is well-known.65 This degradation and loss of sulfur were not observed for RGD-based conjugates, suggesting that in the case of the octreotate conjugates, thiocarbazide-thiol-disulfide interchange/scrambling promotes the loss of H2S from the ligands (Scheme 5).
Scheme 5. Suggested Mechanism for the Formation of a Carbodiimide from H2L1–3(Tyr3-Oct) Resulting in Loss of H2S.
As conventional off-resin oxidative cyclization methodologies were inadequate for this synthesis, H2L1–3(Tyr3-Oct) was prepared entirely on solid support, where intramolecular oxidation/cyclization preceded bioconjugation of H2L1–3. The eight-residue peptide was synthesized by sequential addition of the amino acid residues via solid-phase peptide synthesis, using acetamidomethyl (Acm) protected cysteine residues followed by in situ Acm removal and simultaneous disulfide bond formation using thallium(III) trifluoroacetate.66 Following cyclization, the preactivated SHYNIC derivative (H2L1–3) is reacted at the deprotected d-phenylalanine N-terminus. Cleavage and deprotection of remaining protecting groups are achieved by treatment with trifluoracetic acid (Scheme 6).
Scheme 6. Reaction Scheme for the Formation of Octreotate-Derived Ligands H2L1-3(Tyr3-Oct) and Corresponding Complexes, [ReO(HL1-3(Tyr3-Oct))2]+.
The rhenium complexes of H2L1–3(Tyr3-Oct) could be prepared on-resin or in-solution by treatment with [tBu4N][ReOCl4] in methanol (Scheme 6). The on-resin approach is potentially of interest in producing radioactive complexes in high specific activity as unreacted [ReO4]− and other impurities such as colloidal rhenium could be readily removed by filtration of the resin. The pure complex can be cleaved from the resin with 50% trifluoracetic acid and is stable to this relatively high concentration of acid. Analysis by HPLC and ESI-MS confirmed the identity of the complexes, with the [ReO(HL1–3(Tyr3-Oct))2]+ complexes showing signals in the ESI-MS that could be attributed to the 3+ molecular ion with expected rhenium isotope peak patterns (Figure 3).
Figure 3.
RP-HPLC chromatogram (UV absorbance, a.u., λ 254 nm) of Tyr3-octreotate rhenium complexes, [ReO(HL1–3(Tyr3-Oct))2]+. Inset: Positive ion ESI-MS data of major isotope peaks corresponding to each complex as the tripositive cation, [ReO(HL1–3(Tyr3-Oct))2]3+.
Preparation of [188Re(HL1 (Tyr3-Oct))2]+
Preliminary radiolabeling of H2L1(Tyr3-Oct) with radioactive 188Re was performed using generator-produced [188ReO4]−. A solution of [188ReO4]−, in an aqueous mixture of sodium chloride (0.9% w/v concentration) and sodium tartrate, was reduced with stannous chloride. This mixture was then reacted with H2L1(Tyr3-Oct) at 100 °C, leading to the formation of [188ReO(HL1(Tyr3-Oct))2]+ in ca. 67% radiochemical yield. The compound was characterized by analytical reversed phased HPLC (Figure 4), where [188ReO(HL1(Tyr3-Oct))2]+ (retention time = 11.6 min, detected using a NaI(Tl)) elutes with a similar retention time to the nonradioactive analogue [natReO(HL1(Tyr3-Oct))2]+ (retention time = 11.3 min, detected at λ220), where natRe refers to naturally abundant Re isotopes. The small difference in retention times is due to the different configurations of the radioactivity and UV detectors. This elution profile of [ReO(HL1(Tyr3-Oct))2]+ is distinct from that of the free ligand, H2L1(Tyr3-Oct) that elutes at 9.7 min under the same conditions. Unreacted 188Re species, presumably [188ReO4]−, elute with the solvent front at 2.1 min (Figure 4).
Figure 4.

RP-HPLC chromatogram (UV absorbance, λ 224 nm and radiation detection) of [188/natReO(HL1(Tyr3-Oct))2]+ and free ligand H2L1(Tyr3-Oct). The signal at 2.1 min in the chromatogram of [188ReO(HL1(Tyr3-Oct))2]+ corresponds to unreacted 188Re species.
Preparation of 99mTc-Labeled Complexes, [99mTcO(HL1–3(cRGDfK))2]+ and [99mTcO(HL1–3(Tyr3-Oct))2]+
The technetium-99m complexes [99mTcO(HL1–3(cRGDfK))2]+ and [99mTcO(HL1–3(Tyr3-Oct))2]+ were prepared in ca. 60–80% radiochemical yield using mild conditions and relatively simple procedures (Supporting Information, Figures S1 and S2). Preparation of [99mTcO(HL1–3(cRGDfK))2]+ and [99mTcO(HL1–3(Tyr3-Oct))2]+ involved adding an excess of the appropriate ligand dissolved in aqueous sodium chloride (0.5 mg mL–1, 0.9% NaCl, pH 7.4) to a mixture of [99mTcO4]− that has been reduced by tin(II) chloride in 0.1 M HCl in the presence of tartrate (pH 1–4) at room temperature. The radiolabeled 99mTc complexes were characterized by analysis by HPLC equipped with a radioactivity detector, and the elution profiles were compared to the analogous nonradioactive rhenium compounds (detected by UV absorbance). The close correlation between the retention times of the rhenium and technetium complexes strongly suggests that the complexes are isostructural (Table 3). The small difference in the retention times between the traces for 99mTc and Re complexes is, in part, due to the detector configurations but could also reflect the difference in polarity between the oxorhenium(V) and oxotechnetium(V) cores.
Table 3. RP-HPLC Retention Times (min) for Ligands and [MO(HL)2]+ Complexes (M = Re, 99 mTc)·L1 (ethyl), L2 (t-butyl), and L3 (phenyl) Compoundsa.
| compound | H2Lx | rhenium complex | technetium complex |
|---|---|---|---|
| L1-cRGDfK | 10.0 | 11.4 | 11.5 |
| L2-cRGDfK | 9.3 | 10.3 | 10.9 |
| L3-cRGDfK | 10.9 | 12.0 | 12.7 |
| L1-(Tyr3-Oct) | 10.1 | 11.4 | 11.9 |
| L2-(Tyr3-Oct) | 10.9 | 11.7 | 12.1 |
| L3-(Tyr3-Oct) | 11.0 | 12.9 | 13.0 |
Linear gradient from 0 to 90% Buffer B to A. Buffer A: 0.1% TFA in Milli-Q water. Buffer B: 0.1% TFA in CH3CN.
The stability of [99mTcO(HL3(Tyr3-Oct))2]+ was assessed by incubation in human plasma at 37 °C. The complex was stable for at least 2 h with only small amounts of degradation products (<5%) evident that, based on their retention times in analytical HPLC, are most likely due to degradation of the peptide.
Concluding Remarks
The new pyridylthiocarbazide ligands (SHYNIC, H2L1–3) described here offer a useful alternative to the standard HYNIC system. While HYNIC has proved a very successful and versatile bifunctional ligand for 99mTc coligands are required to complete the coordination sphere of the metal ion and extrapolation to radioactive rhenium isotopes has been challenging.3,63 This family of bidentate ligands form stable complexes with the {ReO}3+ core with two ligands coordinated to a single metal ion. A rhenium complex with a methyl ester functional group has been characterized by X-ray crystallography and features the rhenium ion in a distorted square pyramidal environment with the oxo group in the apical position relative to the pseudo basal plane of two five-membered chelate rings with a N,N/S,S trans configuration about the Re-oxo core. The basic ligands have been decorated with the tumor targeting peptides cyclic-RGD and Tyr3-octreotate, and these conjugates form complexes with rhenium to give well-defined single species, [ReO((HL1–3)(cRGDfK))2] and [ReO((HL1–3)(Tyr3-octreotate))2], without having to add coligands resulting in the formation of a single structural and geometrical isomer. It is possible to form the rhenium complexes using either standard solution chemistry or “on-resin”, and the latter approach may prove useful in isolating radioactive 188/186Re analogues in high specific activity. These complexes feature two targeting peptides separated by 14 chemical bonds, and there is evidence that molecules containing more than one targeting peptide, sometimes referred to as bivalent, can display enhanced receptor binding due to simultaneous binding to more than one receptor on the surface on any given cell.14,67−72 It is possible to prepare [188ReO(HL1(Tyr3-Oct))2]+ in ∼67% yield from generator produced [188ReO4]−, and improved yields should be possible by optimizing the reaction conditions. The analogous technetium complexes, [TcO((HL1–3)(cRGDfK))2] and [TcO((HL1–3)(Tyr3-Octreotate))2], were prepared directly from [99mTcO4]− with tin chloride acting as a reducing agent. Comparison of HPLC profiles suggests the rhenium and technetium complexes are isostructural. The complexes described in this manuscript have two ligands coordinated to a single metal ion, whereas conventional HYNIC systems involve one HYNIC ligand binding to one metal ion. It is likely the two different systems will exhibit quite different biodistribution in vivo. These new systems warrant further investigation as potential theranostic agents employing an imaging (99mTc) and therapeutic (186/188Re) matched pair for a single targeted agent.
General Experimental
All reagents were purchased from standard commercial sources. Nuclear magnetic resonance (NMR) spectra were acquired on either an Agilent 400-MR (1H NMR at 400 MHz and 13C{1H} NMR at 101 MHz) or a Varian FT-NMR 500 spectrometer (1H NMR at 500 MHz and 13C{1H} NMR at 126 MHz) at 298 K. Chemical shifts were referenced to residual solvent peaks and are quoted in ppm relative to TMS.
Fmoc-l-amino acids, Fmoc-d-amino acids, Nvoc-Cl, HATU, DIC, Wang resin, 2-chlorotrityl, Fmoc-Lys(ivDde)–OH, and Fmoc-Cys(Acm)–OH were purchased from standard commercial sources.
Linear protected RGDfK peptide (Arg(Pbf)-Gly(tBoc)-Asp(OtBu)-dPhe-Lys(tBoc)) was synthesized manually using standard Fmoc solid phase peptide synthesis (SPPS) procedures on the 2-chlorotrityl chloride resin. The linear pentapeptide was cleaved from the resin (with retention of protecting groups) using 1% TFA in dichloromethane and shaking for 40 min. The mixture was filtered and the filtrate was reduced in volume to afford crude linear product. Cyclization involved reacting the crude material in a mixture of dichloromethane (1 mg mL–1), HATU (0.9 equiv), and DIPEA (6 equiv) at RT for 2 h, then evaporation to dryness in vacuo. A solution of TFA (97.5%) and Milli-Q water (2.5%) was added to the crude material to deprotect the peptide, followed by removal of the trifluoroacetic acid by sparging with a stream of N2. The peptide was precipitated with diethyl ether, isolated by centrifugation (3 min, 3600 rpm) and dissolved in Milli-Q water (5 mL), and finally purified by semiprep RP-HPLC (Column 1); Gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 40% B to A, over 40 min (1.0% min–1), UV detection at λ 220 nm with a flow rate of 5 mL min–1.
Linear protected Tyr3-octreotate peptide, dPhe-Cys(Trt)-Tyr(tBu)-dTrp(tBoc)-Lys(tBoc)-Thr(tBu)-Cys(Trt)-Thr(tBu)–OH,was synthesized using standard automated Fmoc SPPS procedures on a 2-chorotrityl chloride or Wang resin unless otherwise specified.73
Analytical reversed phase high performance liquid chromatography (RP-HPLC) was undertaken using an Agilent 1100 Series HPLC system at a flow rate of 1 mL min–1 with either Column 1: A Zorbax Eclipse XDB-C18 column (150 mm × 4.6 mm, 5.0 μm) or Column 2: A Phenomenex Aeris Peptide XB-C18 column (250 mm × 4.6 mm, 3.6 μm). Solvent gradients for analytical analysis were either using System A: Gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 100% B over 25 min and UV detection at λ 214, 254, and 280 nm, System B: Gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 60% B over 30 min and UV detection at λ 214, 220, 254, 280, and 350 nm or System E: Gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 20 to 80% B, over 30 min and UV detection at λ 220, 254, and 350 nm with a flow rate of 1 mL min–1.
Semipreparative reversed phase high performance liquid chromatography (semiprep RP-HPLC) was performed using an Agilent 1200 series preparative HPLC unit with variable wavelength detector. An automated Agilent 1200 fraction collector collected 0.5–4 mL fractions. Peak separation was achieved using either Column 3: Kinetex C18 100 Å, AXIA column (150 mm × 21.2 mm, 5 μm), Column 4: Phenomenex Synergi Hydro-RP 80 Å (50 mm × 21.2 mm, 4 μm), Column 5: Varian Pursuit XRs C18 100 Å (150 × 21.2 mm, 5 μm) or Column 6: SGE ProteCol C18 120 Å (250 mm × 10 mm, 5 μm). Gradient elution, flow rate, and wavelength detection are compound specific and are detailed under the Experimental Section of a particular compound. Each fraction collected above 400 mAU was analyzed using ESI-MS and analytical HPLC.
Analytical HPLC traces of radiolabeled 188Re compounds were acquired using an Agilent 1200 LC system with in-line UV and gamma detection (Flow-Count, LabLogic). Peak separation was achieved using an Agilent Eclipse XDB-C18 column (4.6 × 150 mm, 5 μm), with column 1 and system F: Gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 100% B over 20 min and UV detection at λ 220 nm.
Analytical HPLC traces of radiolabeled 99mTc compounds were acquired using a Shimadzu 10 AVP UV–visible spectrophotometer (Shimadzu, Kyoto, Japan) and a sodium iodide scintillation detector with two LC-10ATVP solvent delivery systems for solvents A and B. Peak separation was achieved using Column 7: Nacalai Tesque Cosomosil 5C18-AR Waters column (4.6 × 150 mm, 5 μm) (Kyoto, Japan) at a flow rate of 1 mL min–1. Gradient elution followed System C: Gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 100% B over 20 min and UV detection at λ 254 nm.
X-ray structure determination and refinement was obtained for [ReO(HL1(OMe))2]TFA on an Oxford Diffraction SuperNova CCD diffractometer using Cu–Kα radiation, and the temperature during data collection was maintained at 130.0(1) using an Oxford Cryosystems cooling device. The structure was solved by direct methods using SHELXT and refined using least-squares methods using SHELXL.74,75 Thermal ellipsoid plots were generated using ORTEP-3 integrated within the WINGX suite of programs.76 The trifluoroacetate counterion, although recognizable from the difference electron density maps, was badly disordered and could not be modeled satisfactorily. Application of the Squeeze procedure gave a void volume of 272 Å3 containing 127 electrons, consistent with the presence of two trifluoroacetate anions per unit cell.77 The charge on the complex is unambiguously (+1) given the presence of the two pyridinium protons which are involved in intramolecular hydrogen bonds and the ethylamino protons which are also involved in hydrogen bonds. The crystallographic data has been deposited in the Cambridge Structural Database (CCDC 1543360).
Ligand Synthesis
Note: The designation of H2L1–4refers to the structure with a carboxylic acid-substituted, pyridylhydrazine with the different thiocarbahydrazide functional groups (Et, tBu, and Ph respectively). Further derivatization of the carboxylate functional group is represented by placing the substituted group at the carboxylic carbon in replacement of the OH group (e.g., H2L1–3(OMe), denotes the substitution of a methoxy at the carbonyl carbon to give a methyl ester.
6-[(2-Ethylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, H2L1
To a suspension of 6-hydrazino-3-pyridinecarboxylic acid22 (0.45 g, 2.9 mmol) in anhydrous ethanol (5 mL) was added ethyl isothiocyanate (0.51 mL, 5.8 mmol) under an atmosphere of nitrogen. The suspension was heated to reflux for 16 h then cooled to RT, and the precipitate that formed was collected by filtration, washed with cold ethanol (20 mL), and diethyl ether (20 mL) to give H2L1 as a colorless powder (0.67 g, 95%). 1H NMR [DMSO-d6, 500 MHz]: δ (ppm) 9.39 (1H, s, NH), 9.01 (1H, s, NH), 8.63 (1H, d, J = 1.7 Hz, pyH2), 8.19 (1H, t, J = 4.8 Hz, NH), 8.04 (1H, dd, J = 8.8, 2.2 Hz, pyH4), 6.53 (1H, d, J = 8.8 Hz, pyH5), 3.45 (2H, dt, J = 13.3, 6.8 Hz, CH2), 1.03 (3H, t, J = 7.1 Hz, CH3); 13C{1H} NMR [DMSO-d6, 125.7 MHz]: δ (ppm) 181.5 (C, NCS), 166.5 (C, CO2H), 162.0 (C, pyC6), 150.4 (C, pyC2), 138.8 (C, pyC4), 117.8 (C, pyC3), 105.6 (C, pyC5), 38.4 (C, CH2), 14.6 (C, CH3); IR: ν̅max (cm–1) 2981 (s/sh, N–H), 1605 (s/sh, C=O), 1539 (s/sh), 1319 (m/sh), 1282 (s/sh), 1245 (s/sh), 782 (s/sh); ESI-MS (+): m/z calc’d for C9H13N4O2S 241.0759, found 241.0817 {[M + H]+, 100%}; RP-HPLC (Column 1, System A): RT (min) 7.2.
6-[2-(tert-Butylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, H2L2
To a suspension of 6-hydrazino-3-pyridinecarboxylic acid (0.45 g, 2.9 mmol) in anhydrous DMA (5 mL) was added tert-butyl isothiocyanate (0.56 mL, 4.4 mmol) under an atmosphere of nitrogen. The suspension was heated to 85 °C and after 30 min became a yellow mixture, which was heated at 85 °C for a further 3 h. The mixture was concentrated by evaporation under reduced pressure to a volume of approximately 1 mL and cold diethyl ether (15 mL) was added. The suspension was stirred vigorously overnight at RT. The precipitate was collected via filtration and washed with copious amounts of cold diethyl ether to afford an off-white solid (0.68 g, 86%). 1H NMR [MeOH-d4, 500 MHz]: δ (ppm) 8.73 (1H, d, J = 2.2 Hz, pyH2), 8.18 (1H, dd, J = 8.7, 2.2 Hz, pyH4), 6.76 (1H, dd, J = 8.7, 0.7 Hz, pyH5), 1.51 (9H, s); 13C{1H} NMR [MeOH-d4, 125.7 MHz]: δ (ppm) 181.0 (C, NCS), 166.1 (C, CO2H), 162.0 (C, pyC6), 150.4 (C, pyC2), 138.8 (C, pyC4), 117.8 (C, pyC3), 105.6 (C, pyC5), 54.7 (C, -C(CH3)3), 29.0 (3C, -C(CH3)3); IR: ν̅max (cm–1) 1604 (s/sh, C=O), 1533 (s/sh), 1280 (s/sh), 1252 (s/sh), 1133 (m/sh), 1002 (m/sh), 779 (s/sh); HRMS (ESI+): m/z calc’d for C11H17N4O2S 269.1072, found 269.1110 {[M + H]+, 100%}; RP-HPLC (Column 1, System A): RT (min) 10.6.
6-[2-(Phenylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, H2L3
To a suspension of 6-hydrazino-3-pyridinecarboxylic acid (0.45 g, 2.9 mmol) in anhydrous dimethylacetamide (DMA) (5 mL) was added phenyl isothiocyanate (0.59 mL, 4.4 mmol) under an atmosphere of nitrogen. The suspension was heated to 65 °C and after 5 min became a yellow solution, which was heated at 65 °C for a further 2 h. The mixture was concentrated by evaporation under reduced pressure to a volume of approximately 1 mL, and cold diethyl ether (15 mL) was added A precipitate was collected by filtration and washed with copious amounts of cold diethyl ether to afford an off-white powder (0.81 g, 97%). 1H NMR [DMSO-d6, 400 MHz]: δ (ppm) 9.88 (1H, br s, NH), 9.84 (1H, br s, NH), 9.22 (1H, br s, NH), 8.67 (1H, d, J = 1.9 Hz, pyH2), 8.07 (1H, d, J = 8.4 Hz, pyH4), 7.47 (2H, d, J = 7.2 Hz, ArH), 7.30 (2H, t, J = 7.2 Hz, ArH), 7.13 (1H, t, J = 7.2 Hz, ArH), 6.66 (1H, d, J = 8.8 Hz, pyH5); 13C{1H} NMR [MeOH-d4, 125.7 MHz]: δ (ppm) 181.3 (C, NCS), 166.4 (C, CO2H), 161.7 (C, pyC6), 150.3 (C, pyC2), 139.2 (C, pyC4), 138.7 (C, ArC1), 127.9 (2C, ArC3,5), 125.6 (C, ArC4), 124.9 (2C, ArC2,6), 117.8 (C, pyC3), 106.1 (C, pyC5); IR: ν̅max (cm–1) 1607 (s/sh, C=O), 1596 (s/sh), 1537 (s/sh), 1280 (s/sh), 1254 (s/sh), 1140 (m/sh), 1019 (m/sh), 782 (m/sh); HRMS (ESI+): m/z calc’d for C13H13N4O2S 289.0759, found 289.0749 {[M + H]+, 100%}; RP-HPLC (Column 1, System A): RT (min) 9.9.
Methyl 6-Chloropyridine-3-carboxylate, 3
A suspension of 6-chloronicotinic acid (5.0 g, 32 mmol) in methanol (100 mL) was cooled to 0 °C, and H2SO4 (0.5 mL) was added dropwise. The mixture was heated at 60 °C for 8 h. The crude reaction mixture was evaporated to dryness under reduced pressure. The residue was dissolved in ethyl acetate and washed with sat. NaHCO3. The organic layer was dried over MgSO4, filtered, and evaporated to dryness in vacuo. The pale yellow solid was recrystallized from methanol to form clear, plate-like crystals (5.0 g, 91%). 1H NMR [CHCl3-d, 400 MHz]: δ (ppm) 9.00 (1H, dd, J = 2.4, 0.7 Hz, pyH2), 8.25 (1H, dd, J = 8.3, 2.4 Hz, pyH4), 7.42 (1H, dd, J = 8.3, 0.8 Hz, pyH5), 3.96 (3H, s, CH3); 13C{1H} NMR [CHCl3-d, 125.7 MHz]: δ (ppm) 165.0 (C, CO2Me), 155.8 (C, pyC6), 151.3 (C, pyC2), 139.8 (C, pyC4), 125.2 (C, pyC3), 124.4 (C, pyC5), 52.8 (C, OCH3); ESI-MS (+): m/z calc’d for C7H7ClNO2 172.0165, found 171.9965 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 8.7.
Methyl 6-Hydrazinyl-3-pyridinecarboxylate, 4
Methyl 6-chloropyridine-3-carboxylate, 3, (0.10 g, 0.58 mmol), and hydrazine hydrate (50–60% N2H4, d = 1.029 g cm–3, 0.10 mL, approximately 3 equiv) were added to a 5 mL microwave vial, which was evacuated and purged with N2. Degassed methanol (5 mL) was added to the vial, and the suspension was subjected to microwave irradiation for 15 min at 105 °C. The mixture was filtered and concentrated in vacuo. The residue obtained was dissolved in ethyl acetate and washed with sat. NaHCO3 and sat. NaCl. The organic layer was collected and dried over Na2SO4, filtered, and concentrated to afford a chalky, yellow powder (0.08 g, 81%). 1H NMR [CHCl3-d, 600 MHz]: δ (ppm) 8.95 (1H, d, J = 2.4 Hz, pyH2), 8.21 (1H, dd, J = 8.3, 2.4 Hz, pyH4), 7.40 (1H, d, J = 8.3 Hz, pyH5), 3.92 (3H, s, OCH3); 13C{1H} NMR [CHCl3-d, 151 MHz]: δ (ppm) 165.0 (C, CO2Me), 155.7 (C, pyC6), 151.1 (C, pyC2), 139.7 (C, pyC4), 125.1 (C, pyC3), 124.3 (C, pyC5), 52.7 (C, OCH3); HRMS (ESI+): m/z calc’d for C7H10N3O2 168.0773, found 168.0777 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 7.6.
Methyl 6-[(2-Ethylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylate, H2L1(OMe)
Crude methyl 6-hydrazinyl-3-pyridinecarboxylate, 4, (approximately 0.06 g, 0.67 mmol), was added to anhydrous ethanol (8 mL) and heated until entirely dissolved. Ethyl isothiocyanate (60 μL, 0.67 mmol) was added to the solution and stirred at reflux for 3 h. The solution was cooled to RT. A stream of N2 was blown over the solution, and the resulting precipitate was collected by filtration and washed with copious amounts of diethyl ether to afford a colorless crystalline solid (0.05 g, 51%). 1H NMR [DMSO-d6, 400 MHz]: δ (ppm) 9.41 (1H, s, -NHNHCS), 9.10 (1H, s, -NHNHCS), 8.66 (1H, d, J = 1.8 Hz, pyH2), 8.21 (1H, br t, J = 6.5 Hz, -NHCH2CH3), 8.06 (1H, dd, J = 8.8, 2.1 Hz, pyH4), 6.55 (1H, d, J = 8.8 Hz, pyH5), 3.80 (3H, s, OCH3), 3.45 (2H, quin., J = 6.6 Hz, CH2CH3), 1.03 (3H, t, J = 7.1 Hz, CH2CH3); 13C{1H} NMR [DMSO-d6, 151 MHz]: δ (ppm) 181.3 (C, NCS), 165.4 (C, CO2CH3), 162.2 (C, pyC6), 150.2 (C, pyC2), 138.5 (C, pyC4), 116.7 (C, pyC3), 105.9 (C, pyC5), 51.9 (C, CO2CH3), 38.4 (C, CH2), 14.6 (C, CH3); IR: ν̅max (cm–1) 3143 (m/br, N–H), 2977 (m/br), 1710 (s/sh, C=O), 1533 (m/br), 1295 (s/sh), 1254 (s/sh), 1240 (m/sh), 1119 (w/sh), 781 (m/sh); HRMS (ESI+): m/z calc’d for C10H15N4O2 255.0916, found 255.0938 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 9.2.
Methyl 6-[(2-tert-Butyllcarbamothioyl)hydrazinyl]-3-pyridinecarboxylate, H2L2(OMe)
The compound 6-[2-(tert-butylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, (H2L2) (0.2 g, 0.74 mmol), was added to deoxygenated and dried CH3OH (20 mL). Sulfuric acid (∼0.5 mL) was added dropwise, and the reaction was heated to 50–60 °C for 3 h. Addition of aqueous NaOH (0.1 M) caused a precipitate to form (pH 7–8). The precipitate was collected by filtration as a gray powder (0.13 g, 63%). 1H NMR [CHCl3-d, 500 MHz]: δ (ppm) 8.74 (1H, d, J = 1.7 Hz, pyH2), 8.16 (1H, dd, J = 8.7, 2.2 Hz, pyH4), 6.76 (1H, dd, J = 8.7, 0.3 Hz, pyH5), 3.86 (3H, s, CO2CH3), 1.47 (9H, s, (CH3)3); 13C{1H} NMR [CH3OH-d4, 125.7 MHz]: δ (ppm) 180.9 (C, NCS), 165.6 (C, CO2CH3), 161.4 (C, pyC6), 150.6 (C, pyC2), 140.1 (C, pyC4), 119.8 (C, pyC3), 106.1 (C, pyC5), 53.8 (C, -OCH3), 52.1 (C, -C(CH3)3), 28.7 (3C, -C(CH3)3); IR: ν̅max (cm–1) 1721 (s/sh, C=O), 1523 (m/sh), 1285 (s/sh), 1250 (s/sh), 1133 (m/sh), 775 (s/sh); ESI-MS (+): m/z calc’d for C12H19N4O2S 283.1229, found 283.0450 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 12.9.
Methyl 6-[(2-Phenylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylate, H2L3(OMe)
The same procedure and reagents were used as for the synthesis of H2L1(OMe), except using phenyl isothiocyanate (45 μL, 0.38 mmol) and methyl 6-hydrazinyl-3-pyridinecarboxylate (63 mg, 0.38 mmol) to afford a white, microcrystalline solid (0.08 g, 71%). 1H NMR [CH3OH-d4, 400 MHz]: δ (ppm) 8.77 (1H, s, pyH2), 8.19 (1H, d, J = 8.8 Hz, pyH4), 7.51 (2H, dd, J = 8.1 Hz, ArH2,6), 7.35 (2H, t, J = 7.6 Hz, ArH3,5), 7.21 (1H, t, ArH4), 6.85 (1H, d, J = 8.8 Hz, pyH5), 3.89 (3H, s, CO2CH3); 13C{1H} NMR [DMSO-d6, 125.7 MHz]: δ (ppm) NCS resonance not detected, 165.0 (C, CO2Me), 161.6 (C, pyC6), 150.4 (C, pyC2), 138.8 (C, ArC1), 137.7 (C, pyC4), 128.0 (2C, ArC3,5), 127.0 (C, ArC4), 124.6 (2C, ArC2,6), 116.6 (C, pyC3), 105.3 (C, pyC5), 51.8 (C, CO2CH3); IR: ν̅max (cm–1) 3262 (m/sh, N–H), 1707 (s/sh, C=O), 1618 (m/sh), 1537 (s/sh), 1274 (s/sh), 1133 (m/sh), 742 (m/sh); ESI-MS (+): m/z calc’d for C14H15N4O2S 303.0916, found 303.1274 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 13.1.
H2L1(Lys)
Dichloromethane (10 mL) was added to 2-chlorotrityl resin (1.00 mmol g–1 loading) (3.1 g, 3.1 mmol) swelling the resin. A mixture of Nα-tBoc-Nε-Fmoc-l-Lys (1.7 g, 3.6 mmol) and DIPEA (2.1 mL, 12 mmol) in dichloromethane (15 mL) was added to the resin, which was then shaken at ambient temperature for 2 h. The resin was filtered and the filtrate was discarded. The resulting resin was washed with dichloromethane (3 mL), DMF (3 mL), dichloromethane (3 mL), and finally diethyl ether (3 mL). After several hours of suction drying the resin was weighed and loading determined (0.77 mmol g–1). To the dried resin-bound tBoc-Lys(Fmoc)–OH (0.51 g, 0.39 mmol) was added a DMF/piperidine (80:20 v/v) solution (20 mL), which was manually stirred. After 20 min, the supernatant was removed by filtration through a sintered frit, and the remaining resin was washed with DMF (3 mL). A positive TNBSA assay (2,4,6-trinitrobenzenesulfonic acid in methanol (5% w/v)) was used to indicate deprotection of the Fmoc protected epsilon (ε) amine. A mixture of HATU (285 mg, 0.75 mmol), H2L1 (176 mg, 0.75 mmol), and DIPEA (261 μL, 1.5 mmol) in DMF (4 mL) was added to the resin and reacted for 4 h at ambient temperature. The liquid was drained and the resin was washed successively with dichloromethane (3 mL), then DMF (3 mL), and again with dichloromethane (3 mL). A TFA/H2O/dichloromethane (18:2:80 v/v) solution was added to half the resin (approximately 0.19 mmol), and the mixture shaken for 1 h and then filtered. The filtrate was concentrated in vacuo and diethyl ether (40 mL) was added to form a suspension, which was centrifuged (3 min, 3600 rpm) and the supernatant was discarded. The crude material was dissolved in H2O/CH3CN (90:10 v/v), filtered (Millipore 0.45 μm porosity syringe filter), and then purified by semipreparative RP-HPLC (Column 5). The HPLC system involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 20 to 40% B to A, over 20 min (1.0% min–1) then 40–60% over 40 min (0.5% min–1) and UV detection at 220 nm with a flow rate of 5 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford an off-white, fluffy compound (10 mg, 14%). ESI-MS (+): m/z calc’d for C15H25N6O3S 369.1709, found 369.1798 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 6.00, (Column 2, System B) 24.0.
H2L2(NαOAc-Lys-OMe)
A solution of HATU (0.23 g, 0.56 mmol) and DIPEA (0.29 mL, 1.1 mmol) in DMSO (1 mL) was added to H2L2 (0.15 g, 0.56 mmol) and shaken for 2 min. Nα-Acetyl-l-lysine methyl ester hydrochloride (0.10 g, 0.42 mmol) in DMSO (1 mL) was added and shaken for 2 h. Diethyl ether (40 mL) was added and the mixture was centrifuged. The supernatant was discarded and the remaining yellow oil was purified by semipreparative RP-HPLC (Column 5). The HPLC system involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 10 to 60% B to A, over 50 min (1.0% min–1) and UV detection at λ 220 and 254 nm with a flow rate of 5 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a colorless fluffy solid (0.19 g, 27%). 1H NMR [MeOH-d4, 600 MHz]: δ (ppm) 8.48 (1H, d, J = 2.1 Hz, pyH2), 8.29 (1H, dd, J = 9.1, 2.2 Hz, pyH4), 7.05 (1H, d, J = 9.1 Hz, pyH5), 4.38 (1H, dd, J = 8.9, 5.2 Hz, αCH), 3.70 (3H, s, CO2CH3), 3.38 (2H, t, J = 7.2 Hz, εCH2), 1.97 (3H, s, COCH3), 1.87–1.84 (1H, m, βCH2), 1.75–1.69 (1H, m, βCH2), 1.65–1.61 (2H, m, δCH2), 1.53 (9H, s, (CH3)3), 1.49–1.41 (2H, m, γCH2); 13C{1H} NMR [MeOH-d4, 150.9 MHz]: δ (ppm) 184.0 (C, NCS), 174.2 (C, -COCH3), 173.4 (C, -CO2CH3), 165.9 (C, CONH), 159.1 (C, pyC6), 149.4 (C, pyC2), 141.4 (C, pyC4), 123.7 (C, pyC3), 110.6 (C, pyC5), 54.9 (C, -C(CH3)3), 53.7 (C, αCH2), 52.7 (C, −CO2CH3), 40.7 (C, εCH2), 32.1 (C, βCH2), 29.8 (C, δCH2), 28.9 (3C, -C(CH3)3), 24.2 (C, −COCH3), 22.3 (C, γCH2); HRMS (ESI+): m/z calc’d for C20H33N6O4S 453.2284, found 453.2288 {[M + H]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 12.5.
H2L1(cRGDfK)
To a solution of 6-[(2-ethylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, (H2L1) (16 mg, 66 μmol) in DMF (0.5 mL) was added HATU (25 mg, 66 μmol) and DIPEA (23 μL, 0.13 mmol). A solution of cRGDfK (10 mg, 17 μmol) in DMF (0.3 mL) was added to the initial mixture and then shaken for 2 h at ambient temperature. Diethyl ether (40 mL) was added and the subsequent suspension centrifuged (3 min, 3600 rpm). The supernatant was discarded and the process repeated. The remaining solid was dissolved in a H2O/CH3CN (90:10 v/v) solution, filtered (Millipore 0.45 μm porosity syringe filter), and then purified by semipreparative RP-HPLC (Column 4). The HPLC system involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 50% B to A, over 65 min (1.3% min–1) and UV detection at λ 220, 254, 275, and 350 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford a colorless solid (10 mg, 80%). ESI-MS (+) m/z calc’d for [C36H52N13O8S]+ 826.378, found 826.376, calc’d for [C36H53N13O8S]2+ 413.693, found 413.692; RP-HPLC (Column 2, System A): RT (min) 11.3.
H2L2(cRGDfK)
The same procedure was used as for H2L1(cRGDfK), except H2L1 was replaced with 6-[(2-tert-butylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, (H2L2) (5.3 mg, 19 μmol), and other reagent quantities were adapted accordingly; HATU (6.5 mg, 17 μmol), DIPEA (45 μL, 0.13 mmol), and cRGDfK (10 mg, 17 μmol). Semipreparative RP-HPLC purification (Column 4) involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 40% B to A, over 80 min (0.5% min–1) and UV detection at λ 220, 254, 275, and 350 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford a light yellow solid (3.4 mg, 24%). ESI-MS (+) m/z calc’d for [C38H56N13O8S]+ 854.4096, found 854.4097, calc’d for [C38H57N13O8S]2+ 427.7087, found 427.7083; RP-HPLC (Column 2, System A): RT (min) 12.9.
H2L3(cRGDfK)
The same procedure was used as for H2L1(cRGDfK), except H2L1 was replaced with 6-[(2-phenylcarbamothioyl)hydrazinyl]-3-pyridinecarboxylic acid, (H2L3) (5.2 mg, 19 μmol) and other reagent quantities were adapted accordingly; HATU (6.8 mg, 18 μmol), DIPEA (45 μL, 0.13 mmol), and cRGDfK (10 mg, 17 μmol). Semipreparative RP-HPLC purification (Column 4) involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 40% B to A, over 80 min (0.5% min–1) and UV detection at λ 214, 220, and 254 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a colorless solid (8.6 mg, 55%). ESI-MS (+) m/z calc’d for [C40H52N13O8S]+ 874.3783, found 874.3380, calc’d for [C40H53N13O8S]2+ 437.6930, found 437.6992; RP-HPLC (Column 2, System A): RT (min) 14.2.
H2L1(Tyr3-Oct)
Linear protected Tyr3-octreotate peptide (dPhe-Cys(Acm)-Tyr(tBu)-dTrp(tBoc)-Lys(tBoc)-Thr(tBu)-Cys(Acm)-Thr(tBu)-OH) was prepared manually by standard Fmoc SPPS on Wang resin preloaded with H-Thr(tBu)–OH (100 mg, 0.5 mmol g–1). Amino acid coupling: Coupling of each amino acid was achieved by addition to the resin of a premade solution of Fmoc-amino acid (2 mmol, 4 equiv), HATU (1.9 mmol, 3.8 equiv), and DIPEA (4 mmol, 8 equiv) in DMF (5 mL). The resin mixture was shaken in a reaction vessel equipped with a sintered bottom for 1 h then filtered and washed with DMF (3 mL), dichloromethane (3 mL), and again with DMF (3 mL). Successful coupling was confirmed by a negative TNBSA test. Fmoc deprotection: Cleavage of the Fmoc groups was achieved after each successive coupling reaction. A DMF/piperidine (80:20 v/v) solution was added to the protected resin, which was contained in a reaction vessel with a sintered frit. The mixture was stirred for 5 min, and then the resin was filtered and washed with cleavage solution, then DMF (2 mL), dichloromethane (2 mL) and again DMF (3 mL). Complete deprotection was confirmed with a positive TNBSA test. Acm-group deprotection and on-resin cyclization: After the final Fmoc deprotection of resin-bound Fmoc-dPhe-OH, thallium(III) trifluoroacetate (Tl(CF3CO2)3) (0.54 g, 1.0 mmol, 2 equiv) in DMF (4 mL) was added to the resin, and the mixture was stirred for 2 h. The liquid was decanted from the resin and the resin was washed with DMF (15 mL), dichloromethane (3 mL), and diethyl ether (2 mL). SHYNIC coupling: Half the resin-bound peptide (approximately 0.25 mmol) was used for the following reaction. H2L1 (0.18 g, 0.75 mmol, 3 equiv) was dissolved in DMF (3 mL) and HATU (0.29 g, 0.75 mmol, 3 equiv) added to preactivate the ligand. DIPEA (0.26 mL, 1.5 mmol, 6 equiv) was added and the mixture shaken for 2 min until full dissolution had occurred. The yellow solution was added to the dry resin in a reaction vessel equipped with a sintered frit and shaken for 2 h. The reaction solution was drained and the resin was washed with DMF (15 mL), dichloromethane (3 mL) and again DMF (3 mL). Resin and acid-sensitive protecting group cleavage: The petide-loaded resin was treated with a TFA/TIPS/H2O (96:1.5:2.5 v/v/v) solution (20 mL) and shaken for 2 h. The mixture was filtered and sparged with a steady stream of N2 to reduce the volume by 90%. Cold diethyl ether (40 mL) was added to precipitate the peptide that was isolated by centrifugation (3 min, 3600 rpm). The crude material was dissolved in CH3CN/H2O (50:50 v/v) and lyophilized. The solid was dissolved in CH3CN/H2O (40:60 v/v) and filtered (Millipore 0.45 μm porosity syringe filter). The compound was purified by semipreparative RP-HPLC (Column 5) with a linear gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 20 to 80% B to A, over 120 min (0.5% min–1) and UV detection at λ 214, 220, 254, 280, and 350 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford a pale yellow solid (56 mg, 18%). HRMS (+) m/z calc’d for [C58H75N14O13S3]+ 1271.4800, found 1271.4760, calc’d for [C58H75N14O13S3]2+ 636.2439, found 636.2416; RP-HPLC (Column 2, System A): RT (min) 13.0.
H2L2(Tyr3-Oct)
The same procedure was used as for H2L1(Tyr3-Oct), except H2L1 was replaced with H2L2 (0.20 g, 0.75 mmol, 3 equiv), and all other reagents were used according to their reported equivalencies. The precipitate after lyophilization was dissolved in CH3CN/H2O (8 mL, 50:50 v/v) and filtered (Millipore 0.45 μm porosity syringe filter). The compound was purified by semipreparative RP-HPLC (Column 5) with an isocratic step gradient system of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN). The elution method involved 28% B for 10 min then 32% B for 40 min (desired peak at 34 min) and UV detection at λ 214, 220, 254, 280, and 350 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a fluffy, colorless compound (18 mg, 5.5%). HRMS (+) m/z calc’d for [C60H79N14O13S3]+ 1299.5113, found 1299.5110, calc’d for [C60H80N14O13S3]2+ 650.2596, found 650.2599; RP-HPLC (Column 2, System A): RT (min) 14.4.
H2L3(Tyr3-Oct)
The same procedure was used as for H2L1(Tyr3-Oct), except H2L1 was replaced with H2L3 (0.22 g, 0.75 mmol, 3 equiv). The precipitate after lyophilization was dissolved in CH3CN/H2O (10 mL, 50:50 v/v) and filtered (Millipore 0.45 μm porosity syringe filter). The compound was purified by semipreparative RP-HPLC (Column 5) with a linear gradient system of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 20% to 50% B to A over 60 min (0.5% min–1) (or isocratic elution at 29% Buffer B) and UV detection at λ 254 nm with a flow rate of 7 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford a fluffy white solid (14 mg, 4.2%). HRMS (+) m/z calc’d for [C62H75N14O13S3]+ 1319.4800, found 1319.4849, calc’d for [C62H75N14O13S3]2+ 650.2596, found 636.2425; RP-HPLC (Column 2, System A): RT (min) 14.1.
Synthesis of Rhenium Complexes
Note: Unless specified Re represents ‘rhenium with natural isotope abundance’.
[ReO(HL1(OMe))2]Cl
To H2L1(OMe) (90 mg, 0.35 mmol) and (tBu4N)[ReOCl4] (103 mg, 0.18 mmol) was added anhydrous MeOH (13 mL). The mixture immediately turned deep red and was stirred at room temperature for 4 h. The reaction mixture was filtered, diethyl ether (ca. 15 mL) was added, and the resulting precipitate was collected by filtration to give [ReO(HL1(OMe))2]Cl as a dark-red, microcrystalline solid (0.11 g, 84%). 1H NMR [DMSO-d6, 400 MHz]: δ (ppm) 8.63 (2H, s, pyH2), 8.26 (2H, dd, J = 9.3, 1.8 Hz, pyH5), 7.87 (2H, d, J = 9.4 Hz, pyH4), 7.43 (2H, br m, NH), 3.89 (6H, s, OCH3), 3.47 (4H, qd, J = 13.6, 7.0 Hz, CH2CH3), 1.16 (6H, t, J = 7.5 Hz, CH2CH3); 13C{1H} NMR [CH3CN-d3, 150.9 MHz]: δ (ppm) 173.7 (2C, CO2CH3), 165.1 (2C, N=CS), 162.5 (2C, pyC6), 140.5 (2C, pyC5), 139.9 (2C, pyC2), 123.9 (2C, pyC4), 122.0 (2C, pyC3), 52.3 (2C, CO2CH3), 42.6 (2C, CH2CH3), 14.5 (2C, CH2CH3); IR: ν̅max (cm–1) 1553 (s/sh, C=O), 1287 (s/sh), 963 (m/sh, Re=O), 764 (m/sh); HRMS (ESI+): m/z calc’d for ReC20H26N8O5S2 709.0916, found 709.0898 {[M]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 14.8.
Histidine and Cysteine Challenge Experiments
A 50-fold excess of histidine or cysteine was added to a solution of [ReO(HL1(OMe))2]+ (1 mg/L) in PBS Buffer (10 mM, pH 7.4, 4 mL) and the mixture was heated to 37 °C. At 2, 4, and 24 h after initiation of the experiment, 10 μL aliquots of the reaction mixture were diluted with 90 μL of Milli-Q water and analyzed using analytical HPLC methods. The HPLC traces showed little or no decomposition (<5%) of the rhenium complex at all the time-points.
[ReO(HL2(OMe))2]BPh4
To a stirred suspension of trans-[ReOCl3(PPh3)2] (0.15 g, 0.17 mmol) in MeOH (10 mL) was added H2L2(OMe) (0.10 g, 0.35 mmol) and 2 drops of Et3N. The mixture was heated at reflux for 16 h, then filtered, and diethyl ether was added to the filtrate. The resulting precipitate was collected by filtration, washed with cold diethyl ether, and then dissolved in MeOH. Addition of excess NaBPh4 resulted in the precipitation of a red-brown precipitate that was collected, and washed with hot hexane (10 mL) to afford [ReO(HL2(OMe))2]BPh4 (72 mg, 39%). 1H NMR [CHCl3-d, 600 MHz]: δ (ppm) 8.10 (2H, d, J = 9.5 Hz, pyH2), 7.75 (2H, d, J = 9.4 Hz, pyH5), 7.62 (8H, m, BArH), 7.33 (2H, m, pyH4), 6.93 (8H, t, J = 7.1 Hz, BArH), 6.72 (4H, t, J = 7.1 Hz, BArH), 3.86 (6H, s, CO2CH3), 1.39 (18H, s, (CH3)3); 13C{1H} NMR [CHCl3-d, 150.9 MHz]: δ (ppm) 164.5, 142.5, 135.8–135.8 (8C, BArC), 133.9–133.8, 128.9, 128.6, 128.5, 127.5, 126.2–126.1 (8C, BArC), 122.0 (4C, BArC), 54.9 (2C, -OCH3), 52.6 (2C, -C(CH3)3), 29.0 (6C, -C(CH3)3); IR: ν̅max (cm–1) 1555 (s/sh, C=O), 1284 (s/sh), 960 (m/sh, Re=O), 756 (m/sh); HRMS (ESI+): m/z calc’d for ReC24H34N8O5S2 765.1651, found 765.1559 {[M]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 15.0.
[ReO(HL3(OMe))2]Cl
The same procedure was used as for [ReO(HL1(OMe))2]+, except H2L1(OMe) was replaced with H2L3(OMe) (0.10 g, 0.33 mmol). All other reagents quantities were used accordingly; (tBu4N)[ReOCl4] (95 mg, 0.16 mmol) and MeOH (20 mL). The reaction afforded a red microcrystalline solid (69 mg, 51%). 1H NMR [DMSO-d6, 400 MHz]: δ (ppm) 9.84 (2H, s, NH), 8.79 (2H, s, pyH2), 8.42 (2H, dd, J = 9.1, 1.7 Hz, pyH5), 8.03 (2H, d, J = 9.2 Hz, pyH4), 7.73 (4H, d, J = 7.8 Hz, ArH2,6), 7.32 (4H, t, J = 7.9 Hz, ArH3,5), 6.95 (2H, t, J = 7.3 Hz, ArH4), 3.92 (6H, s, CO2CH3); 13C{1H} NMR [DMSO-d6, 150.9 MHz]: δ (ppm) CO2Me signal not detected, 164.2 (2C), 145.5 (2C), 142.2 (2C, pyC5), 139.5 (2C, pyC2), 130.0 (2C, ArC3,5), 129.3 (4C, ArC4), 128.9 (2C, ArC1), 122.3 (2C, pyC4), 121.4 (2C, pyC3), 117.6 (4C, ArC2,6), 52.6 (2C, CO2CH3)3); IR: ν̅max (cm–1) 1557 (s/sh, C=O), 1282 (s/sh), 960 (m/sh, Re=O), 752 (m/sh); HRMS (ESI+): m/z calc’d for ReC28H26N8O5S2, 805.1014, found 805.1004 {[M]+, 100%}; RP-HPLC (Column 2, System A): RT (min) 15.4.
[ReO(HL2(Nα-Ac-Lys-OMe))2]+
To H2L2(Nα-Ac-Lys-OMe) (15 mg, 33 μmol) in MeOH (1 mL) was added a solution of (tBu4N)[ReOCl4] (0.9 mg, 15 μmol) in MeOH (0.3 mL). The mixture was stirred at ambient temperature for 2 h, then diluted with Milli-Q water and purified by semipreparative RP-HPLC (Column 4). The HPLC system involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 10 to 20% B to A, over 10 min (1.0% min–1), then 20 to 40% B to A over 40 min (0.5% min–1) and UV detection at λ 214, 220, 254, and 280 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford a red solid (5 mg, 27%, assuming [ReO(HL2(Nα-Ac-Lys-OMe))2]CF3CO2). 1H NMR [MeOH-d4, 600 MHz]: δ (ppm) 8.68 (2H, s, pyH2), 8.36 (2H, d, J = 9.6 Hz, pyH5), 8.03 (2H, d, J = 9.5 Hz, pyH4), 4.43–4.40 (2H, m, αCH), 3.72 (6H, s, CO2CH3), 3.43 (4H, q, J = 6.6 Hz, εCH2), 1.99 (6H, s, COCH3), 1.92–1.86 (4H, m, βCH2), 1.78–1.72 (4H, m, δCH2), 1.71–1.66 (4H, m, γCH2), 1.50 (18H, s, (CH3)3); 13C{1H} NMR [MeOH-d4, 150.9 MHz]: δ (ppm) 174.2 (2C, COCH3), 173.4 (2C, CO2CH3), 167.2 (2C, N = CS), 165.1 (2C, CONH), 163.3 (2C, pyC6), 140.7 (2C, d, pyC5), 139.1 (2C, d, pyC2), 124.5 (2C, pyC4), 122.6 (2C, pyC3), 55.3 (2C, -C(CH3)3), 53.7 (C, αCH2), 52.7 (C, −CO2CH3), 40.8 (2C, εCH2), 32.1 (2C, βCH2), 29.9 (2C, δCH2), 28.9 (6C, -C(CH3)3), 24.3 (2C, COCH3), 22.3 (2C, γCH2); HRMS (ESI+) m/z calc’d for [ReC40H62N12O9S2]+ 1105.3762, found 1105.4079, calc’d for [ReC40H62N12O9S]2+ 553.1920, found 553.2094; RP-HPLC (Column 2, System A): RT (min) 16.9.
[ReO(HL1(Lys))2]+
To resin-bound H2L1(Lys) (approximately 0.19 mmol) was added trans-[ReOCl3(PPh3)2] (78 mg, 0.09 mmol) suspended in DMF (5 mL). The slurry was stirred for 2 h. The green suspension turned colorless and the resin turned dark red. The red resin was isolated by filtration and washed with DMF (3 mL), dichloromethane (3 mL), and again with DMF (3 mL). The rhenium complex was cleaved from the resin by treatment with TFA/dichloromethane (50:50 v/v) solution (30 mL) for 3 h. The mixture was sparged with N2 to reduce the volume to 10% of the original volume. Cold diethyl ether (40 mL) was added, the mixture was centrifuged (3 min, 3600 rpm), and the supernatant was discarded. This process was repeated twice. The precipitate was dissolved in CH3CN (1.0 mL) and then Milli-Q water (2.0 mL) was added. The complex was purified by semipreparative RP-HPLC purification (Column 5) involved a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 10 to 70% B to A, over 120 min (0.5% min–1) and UV detection at λ 214, 220, and 254 nm, with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a red, fluffy solid (22 mg, 23%, assuming [ReO(HL1(Lys))2]CF3CO2). 1H NMR [MeOH-d4, 600 MHz]: δ (ppm) 8.58 (2H, s, pyH2), 8.34 (2H, dd, J = 9.5, 1.8 Hz, pyH5), 7.97 (2H, d, J = 9.5 Hz, pyH4), 3.80 (2H, t, J = 6.1 Hz, αCH), 3.60 (4H, dtd, J = 20.2, 13.2, 7.0 Hz, −CH2CH3), 3.46 (4H, dd, J = 8.9, 5.1 Hz, εCH2), 1.99–1.92 (4H, m, βCH2), 1.72 (4H, dt, J = 14.5, 7.3 Hz, δCH2), 1.60–1.50 (4H, m, γCH2), 1.27 (6H, t, −CH2CH3); 13C{1H} NMR [MeOH-d4, 150.9 MHz]: δ (ppm) 172.8 (2C, CO2H), 169.0 (2C, N=CS), 165.4 (2C, CONH), 163.6 (2C, pyC6), 140.5 (2C, pyC5), 138.9 (2C, pyC2), 123.9 (2C, pyC4), 122.5 (2C, pyC3), 54.9 (2C, αCH), 42.5 (2C, CH2CH3), 40.6 (2C, εCH2), 31.5 (2C, βCH2), 29.9 (2C, δCH2), 23.4 (2C, γCH2), 14.9 (2C, CH2CH3); ESI-MS (+) m/z calc’d for [ReC30H46N12O7S2]+ 937.2611, found 937.2574, calc’d for [ReC30H47N12O7S2]2+ 469.1345, found 469.1355, calc’d for [ReC30H48N12O7S2]3+ 313.0256, found 313.0939; RP-HPLC (Column 2, System A): RT (min) 10.1.
[ReO(HL1(cRGDfK))2]+
A solution of H2L1(cRGDfK) (5 mg, 6.1 μmol) in MeOH (250 μL) was added to a mixture of [tBu4N][ReOCl4] (1.8 mg, 3.0 μmol) in MeOH (250 μL). The mixture was shaken for 2.5 h, then diluted with Milli-Q water (5 mL), filtered (Millipore 0.45 μm porosity syringe filter), and purified by semiprep RP-HPLC (Column 6). The HPLC system involved isocratic elution of 77% Buffer A (0.1% TFA in H2O) and 23% Buffer B (0.1% TFA in CH3CN) (0.5% min–1) and UV detection at λ 214, 220, 254, 280, and 350 nm with a flow rate of 5 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a white solid (3.1 mg, 53%, assuming [ReO(HL1(cRGDfK))2]CF3CO2). HRMS (ESI+) m/z calc’d for [ReC72H101N26O17S2]2+ 926.3413, 926.3433, calc’d for [ReC72H102N26O17S2]3+ 617.8968, found 617.9939; RP-HPLC (Column 2, System A): RT (min) 11.0.
[ReO(HL2(cRGDfK))2]+
The same procedure was used as for [ReO(HL1(cRGDfK))2]+, except H2L1(cRGDfK) was replaced with H2L2(cRGDfK) (8.0 mg, 9.4 μmol) in MeOH (250 μL); (tBu4N)[ReOCl4] (2.7 mg, 4.7 μmol) in MeOH (500 μL). The compound was purified by semipreparative RP-HPLC purification (Column 4) involved gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 20% over 20 min (1% min–1), then 20 to 60% B to A over 40 min (0.5% min–1) and UV detection at 214, 220, 254, 280, and 350 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated and lyophilized to afford a white solid (5.1 mg, ∼54%, assuming [ReO(HL2(cRGDfK))2]CF3CO2). HRMS (ESI+) m/z calc’d for [ReC76H109N26O17S2]2+ 954.3732, found 954.3715, calc’d for [ReC76H110N26O17S2]3+ 636.5847, found 636.5837; RP-HPLC (Column 2, System A): RT (min) 12.5.
[ReO(HL3(cRGDfK))2]+
The same procedure was used as for [ReO(HL1(cRGDfK))2]+, except H2L1(cRGDfK) was replaced with H2L3(cRGDfK) (4.0 mg, 4.6 μmol) in MeOH (200 μL), and other reagent quantities were adapted accordingly, [tBu4N][ReOCl4] (1.3 mg, 2.3 μmol) in MeOH (200 μL). The HPLC system (Column 6) involved a linear gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 0 to 20% over 20 min (1% min–1), then 20 to 60% B to A over 40 min (0.5% min–1) and UV detection at 214, 220, 254, 280, and 350 nm with a flow rate of 8 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a white solid (2.0 mg, ∼42%, assuming [ReO(HL3(cRGDfK))2]CF3CO2). HRMS (ESI+) m/z calc’d for [ReC80H101N26O17S2]2+ 974.3419, found 974.3711, calc’d for [ReC80H102N26O17S2]3+ 649.8972, found 650.0023; RP-HPLC (Column 2, System A): RT (min) 12.9.
[ReO(HL1(Tyr3-Oct))2]+
A solution of H2L1(Tyr3-Oct) (9.0 mg, 7.1 μmol) in MeOH (50 μL) was added to a mixture of (tBu4N)[ReOCl4] (2.1 mg, 3.5 μmol) and MeOH (100 μL). The reaction mixture was shaken vigorously for 2 h. Milli-Q water was added (4 mL) and the reaction was filtered (Millipore 0.45 μm porosity syringe filter). The compound was purified by semipreparative RP-HPLC purification (Column 6) with a gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 10 to 80% B to A, over 70 min (1.0% min–1) and UV detection at λ 214, 220, and 254 nm, with a flow rate of 5 mL min–1. Fractions containing the desired compound were identified by HPLC-MS, consolidated, and lyophilized to afford a red, fluffy material (6.8 mg, ∼ 68%, assuming [ReO(HL1(Tyr3-Oct))2]CF3CO2). ESI-MS (+) m/z calc’d for [ReC116H148N28O27S6]2+ 1371.9470, found 1372.0024, calc’d for [ReC116H149N28O27S6]3+ 914.9673, found 915.0145; RP-HPLC (Column 2, System A): RT (min) 15.4.
[ReO(HL2(Tyr3-Oct))2]+
The same procedure was used as for [ReO(HL1(Tyr3-Oct))2]+, except H2L1(Tyr3-Oct) was replaced with H2L2(Tyr3-Oct) (8.0 mg, 6.2 μmol) in MeOH (200 μL) and other reagent quantities were adapted accordingly: (tBu4N)[ReOCl4] (1.8 mg, 3.1 μmol) in MeOH (100 μL). The compound was purified by semipreparative RP-HPLC (Column 6) with isocratic elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) at 38% Buffer B, over 60 min (1.0% min–1) and UV detection at λ 214, 220, and 254 nm, with a flow rate of 4 mL min–1. Fractions containing the desired compound were consolidated and lyophilized to afford a red, fluffy material (5.5 mg, ∼61%, assuming [ReO(HL2(Tyr3-Oct))2]CF3CO2). ESI-MS (+) m/z calc’d for [ReC120H154N28O27S6]2+ 1399.9784, found 1399.4720, calc’d for [ReC120H156N28O27S6]3+ 933.6548, found 933.6460; RP-HPLC (Column 2, System A): RT (min) 16.2.
[ReO(HL3(Tyr3-Oct))2]+
Preloaded, resin-bound (Wang) ligand, H2L3(Tyr3-Oct) (approximately 6.0 μmol), was swollen in CH2Cl2 and drained twice. Anhydrous DMF (2 × 10 mL) was added to the resin the mixture was stirred and drained. DMF (1 mL) was added to the washed resin and (tBu4N)[ReOCl4] (2.1 mg, 2.9 μmol) in DMF (50 μL) was added to the resin. The resin mixture was reacted for 2 h at RT. The solution was drained and the resin was washed with copious amounts of DMF (5 × 10 mL), CH2Cl2 (3 × 10 mL), and diethyl ether (10 mL). The crude material was cleaved off the resin by addition of TFA/CH2Cl2/TIPS/H2O (50:46:1:3 v/v) (20 mL). The mixture was shaken for 1.5 h and filtered and the filtrate sparged with a stream of N2 until the red mixture had reduced to approximately 0.5 mL. Diethyl ether (2 × 10 mL) was added to the crude material, and the resulting precipitate was collected via centrifugation (3 min, 3600 rpm). The compound was purified by semipreparative RP-HPLC purification (Column 6) by gradient elution of Buffer A (0.1% TFA in H2O) and Buffer B (0.1% TFA in CH3CN) from 10 to 80% B to A, over 70 min (1.0% min–1) and UV detection at λ 214, 220, and 254 nm, with a flow rate of 4 mL min–1. Fractions containing the desired compound were consolidated and lyophilized to afford a red, fluffy material (5.0 mg, ∼58%, assuming [ReO(HL3(Tyr3-Oct))2]CF3CO2). ESI-MS (+) m/z calc’d for [ReC124H148N28O27S6]2+ 1419.9470, found 1419.9281, calc’d for [ReC124H149N28O27S6]3+ 946.9673, found 946.1960; RP-HPLC (Column 2, System A): RT (min) 16.2.
Synthesis of [188ReO(HL1(Tyr3-Oct))2]+
Rhenium-188 was produced from an ITG 188W/188Re generator (ITG Isotope Technologies, Garching, Germany). Generator-produced 188ReO4– (86 MBq) in aqueous sodium chloride solution (200 μL, 0.9% w/v) was added to a sealed, N2-purged vial containing sodium tartrate (0.2 mg) and stannous chloride (0.1 mg) in water (400 μL), followed by heating at 80 °C for 30 min. An aliquot of this solution (50 μL) was added to a separate sealed, N2-purged vial containing H2L1(Tyr3-Oct) (100 μg) dissolved in water (100 μL). The reaction was left for 30 min at ambient temperature, after which an aliquot was removed for HPLC analysis, revealing that there was no reaction between 188Re and peptide. The reaction vial was then heated at 100 °C for 1 h, after which an aliquot was analyzed by HPLC (column 1, system F). Radiochemical yield: 67%, retention time of [188ReO(HL1(Tyr3-Oct))2]+ = 11.6 min, compared to retention time of [natReO(HL1(Tyr3-Oct))2]+ = 11.3 min and H2L1(Tyr3-Oct) = 9.7 min.
Synthesis of Technetium-99m Complexes
Sodium pertechnetate, Na[99mTcO4] (1000 MBq), was eluted from a Gentech 99 Mo/99mTc sterile generator (Austin Health: Nuclear Medicine and Centre for PET, Australia, via ANSTO Health) as a 1.0 mL saline solution (0.9% v/v). A solution of SnCl2 in 0.1 M HCl (0.5 mg mL–1) was prepared and purged with nitrogen. Disodium tartrate dihydrate was dissolved in water (1 mg mL–1) in a separate evacuated vial, and a 0.5 mL aliquot was taken from both solutions and mixed together. To the solution was added [99mTcO4]− (0.1 mL in 0.9% saline, 108 MBq). The conjugated peptide, H2L1–4(cRGDfK) or H2L1–3(Tyr3-Oct), was dissolved in degassed Milli-Q water (1 mg mL–1), and 100 μL of this mixture was added to the technetium solution. The sample was neutralized with NaHCO3 (pH 6.5, approximately 55 μL) then filtered or allowed to react without neutralizing at ambient temperature for 30 to 120 min. The samples were filtered (Supelco, Iso-disc Filter, 4 mm x 0.45 μm). Radiochemical yields were evaluated by reverse-phase high-performance liquid chromatography (Column 7, System C).
Stability Studies in Human Serum
Human blood samples were centrifuged with a Heraeus Labofuge 6000 centrifuge at 3000g for 10 min (Heraeus, Hanau, Germany). Radioactivity readings for serum stability studies were taken with a Capintec CRC-35R dose calibrator (Capintec, New Jersey, USA) and were measured in MBq. Centrifugation of radioactive compounds was undertaken using an Eppendorf 5415 D centrifuge (Eppendorf, Hamburg, Germany). Partition coefficient data were collected with a PerkinElmer, Wizard 1470 (PerkinElmer, Massachusetts, USA) automatic γ counter, which measured the radioactive decay of each sample in counts per minute (cpm).
For serum stability studies, blood from a healthy male (20 mL) was centrifuged (10 min, 3000 rpm) to separate blood plasma and red blood cells. The plasma was transferred to a separate vial and the red blood cells were discarded. An aliquot of plasma (0.6 mL) was added to labeled compound, [99mTc(HL3(Tyr3-Oct))2]+ (0.15 mL), the radioactivity was monitored and the mixture was then incubated at 37 °C. Aliquots (0.1 mL) of the mixture were removed from heating at 10 min and after 2 h. Acetonitrile (0.1 mL) was added to the serum/tracer mix to precipitate serum proteins. The suspension was shaken for 5 min and then centrifuged (5 min, 13 200 rpm). The radioactivity of the supernatant and pellet was recorded. The supernatant (20 μL) was analyzed by analytical RP-HPLC (Column 7, System C) for UV and radioactivity analysis and the pellet were washed with acetonitrile (3 × 0.1 mL), and radioactivity levels were again recorded (radioactivity levels of the pellet were negligible).
Acknowledgments
Professor Jonathan R. Dilworth (University of Oxford) is acknowledged for the central role he played in the early aspects of this research. We acknowledge the generous support of Prof. Andrew M. Scott, A/Prof. Henri Tochon-Danguy, Dr. FT Lee, Nick Alexopoulos, and David Thomas of the Austin Hospital, Heidelberg, Victoria, Australia, for providing access to 99mTc and radiochemistry facilities. This research was supported by the Australian Research Council, FT130100204 (PSD). M.T.M. acknowledges the support of the People Programme (Marie Curie Actions) of the European Union’s Seventh Framework Programme (FP7/2007-2013) under REA Grant Agreement Number 299009. This research was supported by the Centre of Excellence in Medical Engineering Centre funded by the Wellcome Trust and EPSRC (WT088641/Z/09/Z), the KCL and UCL Comprehensive Cancer Imaging Centre funded by CRUK and EPSRC in association with the MRC and DoH (England), and by the NIHR Biomedical Research Centre at Guy’s and St Thomas’ NHS Foundation Trust and King’s College London. The views expressed are those of the authors and not necessarily those of the NHS, the NIHR or the DoH.
Supporting Information Available
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b01247.
Representative RP-HPLC chromatograms (PDF)
Accession Codes
CCDC 1543360 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
Australian Research Council FT130100204 (P.S.D.). People Programme (Marie Curie Actions) of the European Union’s Seventh Framework Programme (FP7/2007–2013) under REA Grant Agreement Number 299009 (MTM). The Centre of Excellence in Medical Engineering Centre funded by the Wellcome Trust and EPSRC (WT088641/Z/09/Z), the KCL and UCL Comprehensive Cancer Imaging Centre funded by CRUK and EPSRC in association with the MRC and DoH (England), and the NIHR Biomedical Research Centre at Guy’s and St Thomas’ NHS Foundation Trust and King’s College London.
The authors declare no competing financial interest.
Supplementary Material
References
- Bartholoma M. D.; Louie A. S.; Valliant J. F.; Zubieta J. Technetium and Gallium Derived Radiopharmaceuticals: Comparing and Contrasting the Chemistry of Two Important Radiometals for the Molecular Imaging Era. Chem. Rev. 2010, 110, 2903–2920. 10.1021/cr1000755. [DOI] [PubMed] [Google Scholar]
- Donnelly P. S. The role of coordination chemistry in the development of copper and rhenium radiopharmaceuticals. Dalton Trans. 2011, 40, 999–1010. 10.1039/c0dt01075h. [DOI] [PubMed] [Google Scholar]
- Cutler C. S.; Hennkens H. M.; Sisay N.; Huclier-Markai S.; Jurisson S. S. Radiometals for Combined Imaging and Therapy. Chem. Rev. 2013, 113, 858–883. 10.1021/cr3003104. [DOI] [PubMed] [Google Scholar]
- Ramogida C. F.; Orvig C. Tumour targeting with radiometals for diagnosis and therapy. Chem. Commun. 2013, 49, 4720–4739. 10.1039/c3cc41554f. [DOI] [PubMed] [Google Scholar]
- Blower P. J. A nuclear chocolate box: the periodic table of nuclear medicine. Dalton Trans. 2015, 44, 4819–4844. 10.1039/C4DT02846E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramogida C. F.; Orvig C. Tumour targeting with radiometals for diagnosis and therapy. Chem. Commun. 2013, 49, 4720–4739. 10.1039/c3cc41554f. [DOI] [PubMed] [Google Scholar]
- Pierschbacher M. D.; Ruoslahti E. Cell attachment activity of fibronectin can be duplicated by small synthetic fragments of the molecule. Nature 1984, 309, 30–33. 10.1038/309030a0. [DOI] [PubMed] [Google Scholar]
- Plow E. F.; Pierschbacher M. D.; Ruoslahti E.; Marguerie G. A.; Ginsberg M. H. The effect of Arg-Gly-Asp-containing peptides on fibrinogen and von Willebrand factor binding to platelets. Proc. Natl. Acad. Sci. U. S. A. 1985, 82, 8057–8061. 10.1073/pnas.82.23.8057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pytela R.; Pierschbacher M. D.; Ginsberg M. H.; Plow E. F.; Ruoslahti E. Platelet membrane glycoprotein IIb/IIIa: member of a family of Arg-Gly-Asp-specific adhesion receptors. Science 1986, 231, 1559–1562. 10.1126/science.2420006. [DOI] [PubMed] [Google Scholar]
- Ruoslahti E.; Pierschbacher M. D. Arg-Gly-Asp: a versatile cell recognition signal. Cell 1986, 44, 517–518. 10.1016/0092-8674(86)90259-X. [DOI] [PubMed] [Google Scholar]
- Su Z.-F.; Liu G.; Gupta S.; Zhu Z.; Rusckowski M.; Hnatowich D. J. In Vitro and in Vivo Evaluation of a Technetium-99m-Labeled Cyclic RGD Peptide as a Specific Marker of αVβ3 Integrin for Tumor Imaging. Bioconjugate Chem. 2002, 13, 561–570. 10.1021/bc0155566. [DOI] [PubMed] [Google Scholar]
- Su Z.-F.; He J.; Rusckowski M.; Hnatowich D. J. In vitro cell studies of technetium-99m labeled RGD-HYNIC peptide, a comparison of tricine and EDDA as co-ligands. Nucl. Med. Biol. 2003, 30, 141–149. 10.1016/S0969-8051(02)00390-6. [DOI] [PubMed] [Google Scholar]
- Decristoforo C.; Faintuch-Linkowski B.; Rey A.; von Guggenberg E.; Rupprich M.; Hernandez-Gonzales I.; Rodrigo T.; Haubner R. [99mTc]HYNIC-RGD for imaging integrin αvβ3 expression. Nucl. Med. Biol. 2006, 33, 945–952. 10.1016/j.nucmedbio.2006.09.001. [DOI] [PubMed] [Google Scholar]
- Ji S.; Zhou Y.; Shao G.; Liu S. Evaluation of K(HYNIC)2 as a Bifunctional Chelator for 99mTc-Labeling of Small Biomolecules. Bioconjugate Chem. 2013, 24, 701–711. 10.1021/bc3006896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma M. T.; Neels O. C.; Denoyer D.; Roselt P.; Karas J. A.; Scanlon D. B.; White J. M.; Hicks R. J.; Donnelly P. S. Gallium-68 Complex of a Macrobicyclic Cage Amine Chelator Tethered to Two Integrin-Targeting Peptides for Diagnostic Tumor Imaging. Bioconjugate Chem. 2011, 22, 2093–2103. 10.1021/bc200319q. [DOI] [PubMed] [Google Scholar]
- Ma M. T.; Donnelly P. S. Peptide targeted copper-64 radiopharmaceuticals. Curr. Top. Med. Chem. 2011, 11, 500–520. 10.2174/156802611794785172. [DOI] [PubMed] [Google Scholar]
- Weckbecker G.; Lewis I.; Albert R.; Schmid H.; Hoyer D.; Bruns C. Opportunities in somatostatin research: Biological, chemical and therapeutic aspects. Nat. Rev. Drug Discovery 2003, 2, 999–1017. 10.1038/nrd1255. [DOI] [PubMed] [Google Scholar]
- Maina T.; Nock B.; Nikolopoulou A.; Sotiriou P.; Loudos G.; Maintas D.; Cordopatis P.; Chiotellis E. [99mTc]demotate, a new 99mTc-based [Tyr3]-octreotate analogue for the detection of somatostatin receptor-positive tumours: synthesis and preclinical results. Eur. J. Nucl. Med. Mol. Imaging 2002, 29, 742–753. 10.1007/s00259-002-0782-9. [DOI] [PubMed] [Google Scholar]
- Bigott-Hennkens H. M.; Dannoon S. F.; Noll S. M.; Ruthengael V. C.; Jurisson S. S.; Lewis M. R. Labeling, stability and biodistribution studies of 99mTc-cyclized Tyr3-octreotate derivatives. Nucl. Med. Biol. 2011, 38, 549–555. 10.1016/j.nucmedbio.2010.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahmadihosseini H.; Abedi J.; Ghodsi Rad M. A.; Zakavi S. R.; Knoll P.; Mirzaei S.; Sadeghi R. Diagnostic utility of 99mTc-EDDA-tricine-HYNIC-Tyr3-octreotate SPECT for differentiation of active from inactive pulmonary tuberculosis. Nucl. Med. Commun. 2014, 35, 1262–1267. 10.1097/MNM.0000000000000206. [DOI] [PubMed] [Google Scholar]
- de Jong M.; Breeman W. A. P.; Kwekkeboom D. J.; Valkema R.; Krenning E. P. Tumor Imaging and Therapy Using Radiolabeled Somatostatin Analogues. Acc. Chem. Res. 2009, 42, 873–880. 10.1021/ar800188e. [DOI] [PubMed] [Google Scholar]
- Abrams M. J.; Juweid M.; tenKate C. I.; Schwartz D. A.; Hauser M. M.; Gaul F. E.; Fuccello A. J.; Rubin R. H.; Strauss H. W.; Fischman A. J. Technetium-99m-Human Polyclonal IgG Radiolabeled via the Hydrazino Nicotinamide Derivative for Imaging Focal Sites of Infection in Rats. J. Nucl. Med. 1990, 31, 2022–2028. [PubMed] [Google Scholar]
- Schwartz D. A.; Abrams M. J.; Hauser M. M.; Gaul F. E.; Larsen S. K.; Rauh D.; Zubieta J. A. Preparation of hydrazino-modified proteins and their use for the synthesis of technetium-99m-protein conjugates. Bioconjugate Chem. 1991, 2, 333–336. 10.1021/bc00011a007. [DOI] [PubMed] [Google Scholar]
- Schwartz D. A.; Abrams M. J.; Hauser M. M.; Gaul F. E.; Larsen S. K.; Rauh D.; Zubieta J. A. Preparation of hydrazino-modified proteins and their use for the synthesis of 99mTc-protein conjugates. Bioconjugate Chem. 1991, 2, 333–336. 10.1021/bc00011a007. [DOI] [PubMed] [Google Scholar]
- Babich J. W.; Solomon H.; Pike M. C.; Kroon D.; Graham W.; Abrams M. J.; Tompkins R. G.; Rubin R. H.; Fischman A. J. Technetium-99m-Labeled Hydrazino Nicotinamide Derivatized Chemotactic Peptide Analogs for Imaging Focal Sites of Bacterial Infection. J. Nucl. Med. 1993, 34, 1964–1974. [PubMed] [Google Scholar]
- Nicholson T.; Cook J.; Davison A.; Rose D. J.; Maresca K. P.; Zubieta J. A.; Jones A. G. The synthesis, characterization and X-ray crystal structure of the rhenium organodiazenido, organodiazene complex [ReCl2(PPh3) (NNC5H4N) (HNNC5H4N)]. Inorg. Chim. Acta 1996, 252, 427–430. 10.1016/S0020-1693(96)05310-8. [DOI] [Google Scholar]
- Liu S.; Edwards D. S.; Looby R. J.; Harris A. R.; Poirier M. J.; Barrett J. A.; Heminway S. J.; Carroll T. R. Labeling a Hydrazino Nicotinamide-Modified Cyclic IIb/IIIa Receptor Antagonist with 99mTc Using Aminocarboxylates as Coligands. Bioconjugate Chem. 1996, 7, 63–71. 10.1021/bc950069+. [DOI] [PubMed] [Google Scholar]
- Hirsch-Kuchma M.; Nicholson T.; Davison A.; Davis W. M.; Jones A. G. Synthesis and Characterization of Rhenium(III) and Technetium(III) Organohydrazide Chelate Complexes. Reactions of 2-Hydrazinopyridine with Complexes of Rhenium and Technetium. Inorg. Chem. 1997, 36, 3237–3241. 10.1021/ic961489t. [DOI] [PubMed] [Google Scholar]
- King R. C.; Surfraz M. B.-U.; Biagini S. C. G.; Blower P. J.; Mather S. J. How do HYNIC-conjugated peptides bind technetium? Insights from LC-MS and stability studies. Dalton Trans. 2007, 4998–5007. 10.1039/b705111e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Greenland W. E. P.; Howland K.; Hardy J.; Fogelman I.; Blower P. J. Solid-phase synthesis of peptide radiopharmaceuticals using Fmoc-N-ε-(hynic-Boc)-lysine, a technetium-binding amino acid: application to Tc-99m-labeled salmon calcitonin. J. Med. Chem. 2003, 46, 1751–1757. 10.1021/jm030761n. [DOI] [PubMed] [Google Scholar]
- Dilworth J. R. Diazene, diazenido, isodiazene and hydrazido complexes. Coord. Chem. Rev. 2017, 330, 53–94. 10.1016/j.ccr.2016.09.006. [DOI] [Google Scholar]
- Meszaros L. K.; Dose A.; Biagini S. C. G.; Blower P. J. Synthesis and evaluation of analogues of HYNIC as bifunctional chelators for technetium. Dalton Trans. 2011, 40, 6260–6267. 10.1039/c0dt01608j. [DOI] [PubMed] [Google Scholar]
- Meszaros L. K.; Dose A.; Biagini S. C. G.; Blower P. J. Hydrazinonicotinic acid (HYNIC) - Coordination chemistry and applications in radiopharmaceutical chemistry. Inorg. Chim. Acta 2010, 363, 1059–1069. 10.1016/j.ica.2010.01.009. [DOI] [Google Scholar]
- Liu S.; Hsieh W.-Y.; Kim Y.-S.; Mohammed S. I. Effect of Coligands on Biodistribution Characteristics of Ternary Ligand 99mTc Complexes of a HYNIC-Conjugated Cyclic RGDfK Dimer. Bioconjugate Chem. 2005, 16, 1580–1588. 10.1021/bc0501653. [DOI] [PubMed] [Google Scholar]
- Edwards D. S.; Liu S.; Barrett J. A.; Harris A. R.; Looby R. J.; Ziegler M. C.; Heminway S. J.; Carroll T. R. New and Versatile Ternary Ligand System for Technetium Radiopharmaceuticals: Water Soluble Phosphines and Tricine as Coligands in Labeling a Hydrazinonicotinamide-Modified Cyclic Glycoprotein IIb/IIIa Receptor Antagonist with 99mTc. Bioconjugate Chem. 1997, 8, 146–154. 10.1021/bc970002h. [DOI] [PubMed] [Google Scholar]
- Liu S.; Edwards D. S.; Harris A. R. A Novel Ternary Ligand System for 99mTc-Labeling of Hydrazino Nicotinamide-Modified Biologically Active Molecules Using Imine-N-Containing Heterocycles as Coligands. Bioconjugate Chem. 1998, 9, 583–595. 10.1021/bc9800116. [DOI] [PubMed] [Google Scholar]
- Liu S.; Hsieh W.-Y.; Jiang Y.; Kim Y.-S.; Sreerama S. G.; Chen X.; Jia B.; Wang F. Evaluation of a 99mTc-labeled cyclic RGD tetramer for noninvasive imaging integrin alpha(v)beta3-positive breast cancer. Bioconjugate Chem. 2007, 18, 438–446. 10.1021/bc0603081. [DOI] [PubMed] [Google Scholar]
- Wang L.; Shi J.; Kim Y.-S.; Zhai S.; Jia B.; Zhao H.; Liu Z.; Wang F.; Chen X.; Liu S. Improving tumor-targeting capability and pharmacokinetics of 99mTc-labeled cyclic RGD dimers with PEG(4) linkers. Mol. Pharmaceutics 2009, 6, 231–245. 10.1021/mp800150r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jia B.; Liu Z.; Zhu Z.; Shi J.; Jin X.; Zhao H.; Li F.; Liu S.; Wang F. Blood clearance kinetics, biodistribution, and radiation dosimetry of a kit-formulated integrin αvβ3-selective radiotracer 99mTc-3PRGD2 in non-human primates. Mol. Imag. Biol. 2011, 13, 730–736. 10.1007/s11307-010-0385-y. [DOI] [PubMed] [Google Scholar]
- Rose D. J.; Maresca K. P.; Nicholson T.; Davison A.; Jones A. G.; Babich J.; Fischman A.; Graham W.; DeBord J. R. D.; Zubieta J. Synthesis and Characterization of Organohydrazino Complexes of Technetium, Rhenium, and Molybdenum with the {M(η1-HNNR)(η2-HyNNR)} Core and Their Relationship to Radiolabeled Organohydrazine-Derivatized Chemotactic Peptides with Diagnostic Applications. Inorg. Chem. 1998, 37, 2701–2716. 10.1021/ic970352f. [DOI] [PubMed] [Google Scholar]
- Kovacs M. S.; Hein P.; Sattarzadeh S.; Patrick B. O.; Emge T. J.; Orvig C. Complexes of phosphine-phenolate ligands with the [ReO]3+ and [Re(HNNC5HN)(NNCH4N)]2+ cores. J. Chem. Soc., Dalton Trans. 2001, 3015–3024. 10.1039/b104274m. [DOI] [Google Scholar]
- Cowley A. R.; Dilworth J. R.; Donnelly P. S. A Mono-Diazenide Complex from Perrhenate: Toward a New Core for Rhenium Radiopharmaceuticals. Inorg. Chem. 2003, 42, 929–931. 10.1021/ic025995w. [DOI] [PubMed] [Google Scholar]
- Cowley A. R.; Dilworth J. R.; Donnelly P. S.; Ross S. J. Rhenium diazenide ternary complexes with dithiocarbamate ligands: Towards new rhenium radiopharmaceuticals. Dalton Trans. 2007, 73–82. 10.1039/B611403B. [DOI] [PubMed] [Google Scholar]
- Clarke C.; Cowley A. R.; Dilworth J. R.; Donnelly P. S. Pyridylthiocarbazide complexes of rhenium with potential radiopharmaceutical applications. Dalton Trans. 2004, 2402–2403. 10.1039/b406439a. [DOI] [PubMed] [Google Scholar]
- Alberto R.; Cowley A. R.; Dilworth J. R.; Donnelly P. S.; Pratt J. Diazenide and hydrazide(2-) derivatives of the [Re(CO)3]+ core. Dalton Trans. 2004, 2610–2611. 10.1039/B407708C. [DOI] [PubMed] [Google Scholar]
- Torres J. B.; Andreozzi E. M.; Dunn J. T.; Siddique M.; Szanda I.; Howlett D. R.; Sunassee K.; Blower P. J. PET imaging of copper trafficking in a mouse model of Alzheimer disease. J. Nucl. Med. 2016, 57, 109–114. 10.2967/jnumed.115.162370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teng B.; Bai Y.; Chang Y.; Chen S.; Li Z. Technetium-99m-labeling and synthesis of thymidine analogs: Potential candidates for tumor imaging. Bioorg. Med. Chem. Lett. 2007, 17, 3440–3444. 10.1016/j.bmcl.2007.03.086. [DOI] [PubMed] [Google Scholar]
- Santos I. G.; Abram U. Oxorhenium(V) Complexes with Thiosemicarbazones. Z. Anorg. Allg. Chem. 2004, 630, 697–700. 10.1002/zaac.200400004. [DOI] [Google Scholar]
- Bandoli G.; Gerber T. I. A. Synthesis and characterization of tetrabutylammonium trans-bis(2-amidobenzenethiolato(2-)-S,N)oxotechnetate(V) and its reduction by 2-aminobenzenethiol. Inorg. Chim. Acta 1987, 126, 205–208. 10.1016/S0020-1693(00)84436-9. [DOI] [Google Scholar]
- Chi D. Y.; Katzenellenbogen J. A. Selective formation of heterodimeric bis-bidentate aminothiol-oxometal complexes or rhenium(V). J. Am. Chem. Soc. 1993, 115, 7045–7046. 10.1021/ja00068a101. [DOI] [Google Scholar]
- Konno T.; Shimazaki Y.; Kawai M.; Hirotsu M. Synthesis, Characterization, and Stereochemistry of Oxorhenium(V) Complexes with 2-Aminoethanethiolate. Inorg. Chem. 2001, 40, 4250–4256. 10.1021/ic010326o. [DOI] [PubMed] [Google Scholar]
- Mayer J. M. Metal-oxygen multiple bond lengths: a statistical study. Inorg. Chem. 1988, 27, 3899–3903. 10.1021/ic00295a006. [DOI] [Google Scholar]
- Nugent W. A.; Mayer J. M.: Metal-Ligand Multiple Bonds: The Chemistry of Transition Metal Complexes Containing Oxo, Nitrido, Imido, Alkylidene, or Alkylidyne Ligands; Wiley: Ann Arbor, MI, 1988. [Google Scholar]
- Alvarez Luna S.; Bolzati C.; Duatti A.; Zucchini G. L.; Bandoli G.; Refosco F. Formation of the trans-[dioxorhenium(1+)] group from the reactions of ReNCl2(PPh3)2, ReCl4(PPh3)2, and ReCl3(CH3CN)(PPh3)2 with chelating amines and tetraazamacrocycles. Crystal structure of trans-[ReO2(cyclam)](PF6) (cyclam = 1,4,8,11-tetraazacyclotetradecane). Inorg. Chem. 1992, 31, 2595–2598. 10.1021/ic00038a052. [DOI] [Google Scholar]
- Wang Y. P.; Che C. M.; Wong K. Y.; Peng S. M. Spectroscopy, molecular structure, and electrochemistry of rhenium(V) oxo and imido complexes of 1,4,8,11-tetraazacyclotetradecane (cyclam). Inorg. Chem. 1993, 32, 5827–5832. 10.1021/ic00077a029. [DOI] [Google Scholar]
- Le Gal J.; Tisato F.; Bandoli G.; Gressier M.; Jaud J.; Michaud S.; Dartiguenave M.; Benoist E. Synthesis and structural characterization of new oxorhenium and oxotechnetium complexes with XN2S-tetradentate semi-rigid ligands (X = O, S, N). Dalton Trans. 2005, 3800–3807. 10.1039/b508661b. [DOI] [PubMed] [Google Scholar]
- Paterson B. M.; White J. M.; Donnelly P. S. A hexadentate bis(thiosemicarbazonato) ligand: rhenium(V), iron(III) and cobalt(III) complexes. Dalton Trans. 2010, 39, 2831–2837. 10.1039/b922127a. [DOI] [PubMed] [Google Scholar]
- Cowley A. R.; Dilworth J. R.; Donnelly P. S.; Woollard-Shore J. Synthesis and characterisation of new homoleptic rhenium thiosemicarbazone complexes. Dalton Trans. 2003, 748–754. 10.1039/b210540n. [DOI] [Google Scholar]
- Greenland W. E. P.; Howland K.; Hardy J.; Fogelman I.; Blower P. J. Solid-Phase Synthesis of Peptide Radiopharmaceuticals Using Fmoc-N-ε-(Hynic-Boc)-Lysine, a Technetium-Binding Amino Acid: Application to Tc-99m-Labeled Salmon Calcitonin. J. Med. Chem. 2003, 46, 1751–1757. 10.1021/jm030761n. [DOI] [PubMed] [Google Scholar]
- Stephenson K. A.; Zubieta J.; Banerjee S. R.; Levadala M. K.; Taggart L.; Ryan L.; McFarlane N.; Boreham D. R.; Maresca K. P.; Babich J. W.; Valliant J. F. A New Strategy for the Preparation of Peptide-Targeted Radiopharmaceuticals Based on an Fmoc-Lysine-Derived Single Amino Acid Chelate (SAAC). Automated Solid-Phase Synthesis, NMR Characterization, and in Vitro Screening of fMLF(SAAC)G and fMLF[(SAAC-Re(CO)3)+]G. Bioconjugate Chem. 2004, 15, 128–136. 10.1021/bc034128s. [DOI] [PubMed] [Google Scholar]
- James S.; Maresca K. P.; Allis D. G.; Valliant J. F.; Eckelman W.; Babich J. W.; Zubieta J. Extension of the Single Amino Acid Chelate Concept (SAAC) to Bifunctional Biotin Analogues for Complexation of the M(CO)3+1 Core (M = Tc and Re): Syntheses, Characterization, Biotinidase Stability, and Avidin Binding. Bioconjugate Chem. 2006, 17, 579–589. 10.1021/bc050297w. [DOI] [PubMed] [Google Scholar]
- Maresca K. P.; Hillier S. M.; Femia F. J.; Zimmerman C. N.; Levadala M. K.; Banerjee S. R.; Hicks J.; Sundararajan C.; Valliant J.; Zubieta J.; Eckelman W. C.; Joyal J. L.; Babich J. W. Comprehensive Radiolabeling, Stability, and Tissue Distribution Studies of Technetium-99m Single Amino Acid Chelates (SAAC). Bioconjugate Chem. 2009, 20, 1625–1633. 10.1021/bc900192b. [DOI] [PubMed] [Google Scholar]
- Bartholomae M.; Valliant J.; Maresca K. P.; Babich J.; Zubieta J. Single amino acid chelates: a strategy for the design of technetium and rhenium radiopharmaceuticals. Chem. Commun. 2009, 493–512. 10.1039/B814903H. [DOI] [PubMed] [Google Scholar]
- Haubner R.; Gratias R.; Diefenbach B.; Goodman S.; Jonczyk A.; Kessler H. Structural and functional aspects of RGD-containing cyclic pentapeptides as highly potent and selective integrin αvβ3 antagonists. J. Am. Chem. Soc. 1996, 118, 7461–7472. 10.1021/ja9603721. [DOI] [Google Scholar]
- Khorana H. G. The chemistry of carbodiimides. Chem. Rev. 1953, 53, 145–166. 10.1021/cr60165a001. [DOI] [Google Scholar]
- Petersen A. L.; Binderup T.; Jølck R. I.; Rasmussen P.; Henriksen J. R.; Pfeifer A. K.; Kjær A.; Andresen T. L. Positron emission tomography evaluation of somatostatin receptor targeted 64Cu-TATE-liposomes in a human neuroendocrine carcinoma mouse model. J. Controlled Release 2012, 160, 254–263. 10.1016/j.jconrel.2011.12.038. [DOI] [PubMed] [Google Scholar]
- Shi J.; Wang L.; Kim Y.-S.; Zhai S.; Liu Z.; Chen X.; Liu S. Improving Tumor Uptake and Excretion Kinetics of 99mTc-Labeled Cyclic Arginine-Glycine-Aspartic (RGD) Dimers with Triglycine Linkers. J. Med. Chem. 2008, 51, 7980–7990. 10.1021/jm801134k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang L.; Shi J.; Kim Y.-S.; Zhai S.; Jia B.; Zhao H.; Liu Z.; Wang F.; Chen X.; Liu S. Improving Tumor-Targeting Capability and Pharmacokinetics of 99mTc-Labeled Cyclic RGD Dimers with PEG4 Linkers. Mol. Pharmaceutics 2009, 6, 231–245. 10.1021/mp800150r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Z.; Jia B.; Shi J.; Jin X.; Zhao H.; Li F.; Liu S.; Wang F. Tumor Uptake of the RGD Dimeric Probe 99mTc-G3-2P4-RGD2 is Correlated with Integrin αvβ3 Expressed on both Tumor Cells and Neovasculature. Bioconjugate Chem. 2010, 21, 548–555. 10.1021/bc900547d. [DOI] [PubMed] [Google Scholar]
- Zhou Y.; Kim Y.-S.; Lu X.; Liu S. Evaluation of 99mTc-Labeled Cyclic RGD Dimers: Impact of Cyclic RGD Peptides and 99mTc Chelates on Biological Properties. Bioconjugate Chem. 2012, 23, 586–595. 10.1021/bc200631g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Y.; Ji S.; Liu S. Impact of Multiple Negative Charges on Blood Clearance and Biodistribution Characteristics of 99mTc-Labeled Dimeric Cyclic RGD Peptides. Bioconjugate Chem. 2014, 25, 1720–1729. 10.1021/bc500309r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imberti C.; Terry S. Y. A.; Cullinane C.; Clarke F.; Cornish G. H.; Ramakrishnan N. K.; Roselt P.; Cope A. P.; Hicks R. J.; Blower P. J.; Ma M. T. Enhancing PET Signal at Target Tissue in Vivo: Dendritic and Multimeric Tris(hydroxypyridinone) Conjugates for Molecular Imaging of αvβ3 Integrin Expression with Gallium-68. Bioconjugate Chem. 2017, 28, 481–495. 10.1021/acs.bioconjchem.6b00621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Atherton E.; Fox H.; Harkiss D.; Logan C. J.; Sheppard R. C.; Williams B. J. A mild procedure for solid phase peptide synthesis: use of fluorenylmethoxycarbonylamino-acids. J. Chem. Soc., Chem. Commun. 1978, 537–539. 10.1039/c39780000537. [DOI] [Google Scholar]
- Sheldrick G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112–122. 10.1107/S0108767307043930. [DOI] [PubMed] [Google Scholar]
- Sheldrick G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3–8. 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farrugia L. J. WinGX and ORTEP for Windows: an update. J. Appl. Crystallogr. 2012, 45, 849–854. 10.1107/S0021889812029111. [DOI] [Google Scholar]
- Spek A. L. PLATON SQUEEZE: a tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 9–18. 10.1107/S2053229614024929. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.










