Next Article in Journal
Repurposing Therapeutic Drugs Complexed to Vanadium in Cancer
Previous Article in Journal
Anti-Inflammatory Activity of Pequi Oil (Caryocar brasiliense): A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Trace Metal Impurities Effects on the Formation of [64Cu]Cu-diacetyl-bis(N4-methylthiosemicarbazone) ([64Cu]Cu-ATSM)

1
Faculty of Science, Toho University, Funabashi 274-8510, Japan
2
Institute for Quantum Medical Science, National Institutes for Quantum Science and Technology, Chiba 263-8555, Japan
3
Kanagawa Cancer Center, Kanagawa 241-8515, Japan
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2024, 17(1), 10; https://doi.org/10.3390/ph17010010 (registering DOI)
Submission received: 15 November 2023 / Revised: 15 December 2023 / Accepted: 19 December 2023 / Published: 21 December 2023
(This article belongs to the Special Issue Copper Radiopharmaceuticals for Theranostic Applications)

Abstract

:
[64Cu]Cu-diacetyl-bis(N4-methylthiosemicarbazone) ([64Cu]Cu-ATSM) is a radioactive hypoxia-targeting therapeutic agent being investigated in clinical trials for malignant brain tumors. For the quality management of [64Cu]Cu-ATSM, understanding trace metal impurities’ effects on the chelate formation of 64Cu and ATSM is important. In this study, we conducted coordination chemistry studies on metal–ATSM complexes. First, the effects of nonradioactive metal ions (Cu2+, Ni2+, Zn2+, and Fe2+) on the formation of [64Cu]Cu-ATSM were evaluated. When the amount of Cu2+ or Ni2+ added was 1.2 mol or 288 mol, equivalent to ATSM, the labeling yield of [64Cu]Cu-ATSM fell below 90%. Little effect was observed even when excess amounts of Zn2+ or Fe2+ were added to the ATSM. Second, these metals were reacted with ATSM, and chelate formation was measured using ultraviolet–visible (UV-Vis) absorption spectra. UV-Vis spectra showed a rapid formation of Cu2+ and the ATSM complex upon mixing. The rate of chelate formation by Ni2+ and ATSM was lower than that by Cu-ATSM. Zn2+ and Fe2+ showed much slower reactions with the ATSM than Ni2+. Trace amounts of Ni2+, Zn2+, and Fe2+ showed little effect on [64Cu]Cu-ATSM’ quality, while the concentration of impurity Cu2+ must be controlled. These results can provide process management tools for radiopharmaceuticals.

1. Introduction

Hypoxia has been implicated in the poor prognosis of malignant tumors, such as recurrent brain tumors [1,2,3,4]. Extensive hypoxic areas are associated with a poor prognosis in patients, as malignant factors are promoted in hypoxic cancer cells [2,3,4,5]. Therefore, new tumor therapeutics targeting hypoxia have extensively been investigated to improve patient outcomes. We developed a promising radioactive hypoxia-targeting theranostic agent, [64Cu]Cu-diacetyl-bis(N4-methylthiosemicarbazone) ([64Cu]Cu-ATSM). Clinical studies on patients with recurrent brain tumors are underway.
Cu-ATSM (Figure 1) is labeled with several copper isotopes, such as 60Cu, 61Cu, 62Cu, and 64Cu, for positron emission tomography (PET) imaging under hypoxia [6,7,8,9,10,11,12,13,14]. Cu-ATSM rapidly permeates cells, is reduced, and is trapped within cells under hypoxia [15,16,17,18,19]. In hypoxic cells, the levels of the biological reductant NAD(P)H and the activity of NAD(P)H-dependent reductive enzymes are upregulated, and highly reduced conditions lead to Cu-ATSM retention in the cells [16,20,21,22]. To develop hypoxia-targeting radiotherapeutics, 64Cu was chosen to label ATSM. 64Cu emits β+ for PET imaging (0.655 MeV, 17.8%), β- (0.574 MeV, 40%), and particularly, Auger electrons bearing a high potential for killing cancer cells [21].
In our previous study [22], using a xenograft mouse model of glioblastoma, multiple doses of [64Cu]Cu-ATSM significantly prolonged survival, but no toxicity was reported [22]. We stabilized the therapeutic dose of [64Cu]Cu-ATSM by adding sodium L-ascorbate to our investigational drug formulation [23]. We also evaluated trace amounts of chemical impurities derived from the degradation of ATSM in our formulation of [64Cu]Cu-ATSM [24]. We identified chemical impurities derived from the degradation of ATSM by using liquid chromatography with tandem mass spectrometry (LC-MS/MS). We also assessed their chemical hazards using quantitative structure–activity relationship (QSAR) applications and concluded that the potential risk posed by chemical impurities contained in the therapeutic dose of [64Cu]Cu-ATSM is negligible.
Cu-64 was obtained by the 64Ni(p,n)64Cu reaction using a cyclotron. After the dissolution of the 64Ni target with hydrochloric acid, no carrier added (n.c.a.) 64Cu was obtained by chemical separation using an anionic exchange resin or cationic exchange resin [25,26,27,28]. Although the 64Ni target was carefully purified before irradiation, minimal amounts (100–102 ppb) of impurities such as Cu, Ni, Fe, and Zn remained in the irradiated sample solution. Most metal ions are present as divalent ions in solution and may react competitively with ATSM to form [64Cu]Cu-ATSM chelates, thereby inhibiting [64Cu]Cu-ATSM’s formation.
Our formulation of the [64Cu]Cu-ATSM investigational drug contains 2.5 µg/mL of ATSM and up to 1.5 GBq/mL of 64Cu [23]. The molar ratio of ATSM to n.c.a. 64Cu in the therapeutic formulation was approximately 60. Therefore, sub-ppm-level contamination of metal ions may negatively affect the radiochemical yield of [64Cu]Cu-ATSM. As of date, the effect of trace metal impurities on chelate formation by 64Cu has been reported for some ligands by Ferreira et al. (DOTA-DBCB) [29], Boswell et al. (H2CB-TE2A) [30], and Zeng et al. (EdF-DOTA) [31]; however, that for ATSM remains unclear. Therefore, control of trace metal impurities would be critical for the quality management of 64Cu radiopharmaceutical agents. To obtain quantitative insights into the chelate formulations of ATSM and metal ions, we conducted studies on the coordination chemistry of metal–ATSM complexes in this study. First, we investigated the effect of non-radioactive metal ions on the formation of [64Cu]Cu-ATSM. Copper, nickel, zinc, and iron were chosen for this experiment because they are possible contaminants in the target system of the cyclotron [32]. Next, these metals were made to react with ATSM, and the chelate formation rate was measured using ultraviolet–visible (UV-Vis) absorption spectra. These results can provide useful process management tools for this promising radiopharmaceutical for the treatment of recurrent brain tumors in future clinical studies.

2. Results

2.1. Experiment 1: Effect of Trace Metal Impurities on the Formation of [64Cu]Cu-ATSM Complex

We investigated the radiochemical yield of [64Cu]Cu-ATSM when metal solutions (Cu2+, Ni2+, Zn2+, and Fe2+) at nine different concentrations (0.0125, 0.025, 0.05, 0.1, 0.2, 0.4, 4, 10, and 100 ppm in 40.5 μL of the reaction mixture) were added during the synthesis. The concentration range was set at approximately 100 times the amount considered to be contaminated based on the literature [26,32]. The results are summarized in Figure 2 and Table 1 and Table 2. Figure 2 shows the plots of the radiochemical yield of [64Cu]Cu-ATSM against the concentration of metal ions in the solutions. The top axis shows the molar ratio of the metal ions to ATSM. Table 1 and Table 2 show the concentration of metal ions, radiochemical yield, and molar ratio of M2+ to ATSM (i.e., M2+/ATSM). As shown in Figure 2a, Cu2+ inhibited the formation of the [64Cu]Cu-ATSM complex at lower concentrations compared to other metal ions. When 1.2 mol equivalent of Cu2+ or more was added to ATSM, the radiochemical yield of [64Cu]Cu-ATSM was reduced to lower than 90%, which is regarded as a quality standard for radiopharmaceuticals.
Figure 2b shows the concentration-dependent effects of Ni2+ on [64Cu]Cu-ATSM complex formation. The radiochemical yield of [64Cu]Cu-ATSM fell below 90% upon adding 288 equivalent Ni2+ ions to ATSM. Thus, the effect of Ni2+ on [64Cu]Cu-ATSM formation was 240 times weaker than that of Cu2+. Little effects of Zn2+ and Fe2+ were observed, even with the addition of 248 mol or 290 mol equivalents of ATSM (Figure 2c and 2d, respectively, and Table 2).

2.2. Experiment 2: Chelate Formation of ATSM with Metal Ions

In Experiment 2, we examined chelate formation when metal ions (Cu2+, Ni2+, Zn2+, and Fe2+) were added to the ATSM with UV-Vis spectra over time. Figure 3 shows spectral changes in the ATSM after adding transition metal ions. Cu2+ chelated ATSM within 5 min. Ni2+ also formed chelates with ATSM; however, the reaction rate was slower than that with Cu2+. In the case of Fe2+, the ATSM peak at 337 nm gradually decreased, and a weak shoulder appeared at around 400 nm, indicating a slower reaction rate of Fe2+ than that of Ni2+. The spectrum of Zn2+ remained unchanged after 24 h, indicating a prolonged reaction rate.
The reactions for Cu2+ and Ni2+ were studied semi-quantitatively; the concentrations of Cu- and Ni-ATSM were estimated using the molar extinction coefficients, ε, which were determined at 7680 (Cu-ATSM at 477 nm) and 14,300 cm M–1 (Ni-ATSM at 402 nm) in this study (Figure S1). Although the ε values in DMSO have not been reported, the values in DMF (Cu-ATSM) [33] and pyridine (Ni-ATSM) [34] are similar to our findings. Figure 4 shows the time-dependent increase in the M-ATSM concentration; the vertical axis represents the concentration of M-ATSM relative to the complete complex concentration. The concentration of Cu-ATSM reached 100% immediately after 5 min, suggesting the completion of the chelate formation. In contrast, the complexation yield of Ni2+ with ATSM gradually increased from 5.8% at 5 min to over 50% at 2 h, indicating that the formation of Ni-ATSM is slower and weaker than that of Cu-ATSM.

2.3. Experiment 3: Chelate Formation of Zn-ATSM in the Presence of Brønsted Base

In Experiment 2, the UV-Vis spectrum of ATSM showed that only Zn2+ did not react under this formulation condition. To understand the underlying reason, we hypothesized that deprotonation of ATSM would be necessary for coordination with Zn2+, and, therefore, we checked whether the use of a Brønsted base changes the reaction rate of Zn-ATSM. Figure 5 shows the changes in the UV-Vis spectrum of ATSM after mixing with ZnCl2 and sodium methoxide as a Brønsted base. In contrast to the Zn-ATSM spectrum shown in Figure 3c, Zn-ATSM began to form 15 min after the initiation of the reaction. The ATSM peak slightly decreased for 20 h, but that of Zn-ATSM did not change from 15 min to 20 h.

2.4. Experiment 4: Differences in the Formation Constants of Cu-ATSM and Ni-ATSM

To confirm the difference of the formation constants between Cu2+ and Ni2+ with ATSM, 150 μM CuCl2, 150 μM NiCl2, and 150 μM, ATSM were mixed in a total volume of 600 µL, and the UV-Vis spectra change was obtained (Figure 6). Compared to Figure 3a,b, the Cu-ATSM spectrum was observed within 5 min of the start of the reaction, and there was no significant change in the spectrum after 2 h, suggesting that the formation constant of Cu-ATSM was much larger than that of Ni-ATSM. The results showed that upon mixing Cu2+ and Ni2+, the chelate formation rates of ATSM and Cu2+ were much higher than those of Ni2+, and the formation of Cu-ATSM was thermodynamically predominant.

3. Discussion

[64Cu]Cu-ATSM is a promising therapeutic radiopharmaceutical for clinical trials. The 64Cu used in [64Cu]Cu-ATSM was prepared using a cyclotron and contained trace metal impurities, mainly Cu2+, Ni2+, Zn2+, and Fe2+, derived from cyclotron target systems and production lines. Therefore, to obtain insights into the effects of trace metal impurities on the chelate formation of 64Cu and ATSM ligands, we conducted coordination chemistry studies on metal–ATSM complexes.
Experiment 1 indicates that trace amounts of Cu2+ and Ni2+ affect the radiochemical yield of [64Cu]Cu-ATSM. In contrast, Fe2+ and Zn2+ did not affect radiochemical yield, even when the concentration of these metals reached up to 100 ppm. This result is consistent with those of Experiment 2 (vide infra). Defined as a specification for the investigational drug formulation of [64Cu]Cu-ATSM, the radiochemical yield of the chelate should be not less than 90%. Considering this criterion, the maximum metal ion concentrations for [64Cu]Cu-ATSM formation were 0.26 ppm (260 ppb) for Cu2+, 10 ppm for Ni2+, and 100 ppm for Fe2+ and Zn2+. These values were defined as the highest experimental concentrations at which radiochemical yield could be maintained above 90% in the presence of the impurities.
The metal impurities in the irradiated 64Ni target were estimated using previously reported values [26,32,35,36]. The concentration of metal impurities can differ among laboratories because it depends on the cyclotron irradiation, mainly on the target and subsequent separation process. The 64Cu source was produced at our institute, QST, and we referred to the values of Ohya et al. [32]. The reported concentrations in 10 mL solutions were Cu: 36 ± 22, Ni: 12 ± 27, Fe: 64 ± 48, and Zn: 157 ± 42 ppb. Assuming that the impurities came from only 10 mL of the solution recovered from the target and that all impurities were in the final product, the estimated concentrations of the metal impurities in our products were well below the recommended levels mentioned above, indicating that the metal impurities from the target did not affect the radiochemical yields in our formulation. In this study, the concentration of Cu2+ impurities in the reaction mixture derived from the target system was estimated to be 60.8 ppb in the reaction mixture. This value was approximately 20% of the allowed value (260 ppb). In the production process of 64Cu, the 64Ni fraction is usually recovered to regenerate as the target; as a result, the purity of the 64Ni target increases, and control of the copper impurities in the target is essential. According to the values of metal impurities reported by other institutes, the copper content reaches the permissible value in one case; thus, in such a case, the removal of Cu2+ from the 64Ni target using a chelate resin, for instance, is required [37,38].
The chelate formation reaction was examined qualitatively from a kinetic point of view, with contrasting results obtained for Cu2+ and Ni2+. Cu2+ was made to quantitatively react with ATSM within 5 min, whereas Ni2+ showed a slow reaction. The formation of Ni-ATSM complexes was almost complete after 24 h, but the amount of formed Ni-ATSM was only 86% of that of the theoretical amount from the quantitative reaction. This observation suggests that the system was in chemical equilibrium, and Ni2+ did not react quantitatively with ATSM. Thus, the reaction of metal ions with the ATSM is more favorable with Cu2+ than with Ni2+. The low inhibition of Ni2+ and ATSM can be explained by both thermodynamic and kinetic consequences.
The reaction of Zn2+ with ATSM showed little spectral change, suggesting a slow reaction rate for Zn2+. One reason for the slow reaction may be that deprotonation of ATSM is required to coordinate with Zn2+. The reaction proceeded more rapidly when we added two equivalents of sodium methoxide, a strong Brønsted base (Figure 5). This finding suggests that a kinetic factor is involved, whereby Zn2+ does not inhibit the formation of the Cu-ATSM complex. Therefore, deprotonating agents may inhibit the ions involved in [64Cu]Cu-ATSM formation, and the concentration of deprotonating agents in the [64Cu]Cu-ATSM formulation must be strictly controlled.
The kinetic experiments indicated that the reaction rates are different between Cu2+ and Ni2+. This phenomenon is frequently observed in the formation of metal complexes in aqueous solutions. The exchange reaction rate of the solvent molecule dominated most of the formation reaction rates, and no ligand could react faster than the solvent molecule. A significant difference in the solvent exchange reaction in aqueous solutions was observed between Cu2+ and Ni2+; the exchange rate constant, kex, of Cu2+ was 105 times larger than that of Ni2+ in water [39]. We expected a similar relationship with DMSO concentration.
Under the conditions of Experiment 1, the maximally allowed Ni2+ concentration of 10 ppm corresponded to 170 μM, and the molar ratio Ni2+/ATSM was 29. This indicated that Ni2+ did not inhibit [64Cu]Cu-ATSM formation up to 29 equivalents to ATSM. This observation suggests the low chelate formation ability of Ni2+, although the coexisting glycine, added as an auxiliary complexing agent to maintain Cu2+ in the solution, can weaken the coordination of ATSM to some extent through the competition of ATSM with glycine. This coexisting glycine also competed with [64Cu]Cu-ATSM formation because the molar ratio of total Cu, the sum of cold Cu and 64Cu, to ATSM was only 0.68 at 91% radiochemical yield.
The effect of metal ions on the formation of [44Sc]Sc-DOTATATE (Figure S2) has been reported [40]. Two nmol/mL of Fe2+/3+ and Zn2+ and 10 nmol/mL of Cu2+ significantly reduced the labeling yield, while Ca2+ and Al3+ did not affect it up to 2 μmol/mL. The formation of [44Sc]Sc-DOTATATE was more sensitive to metal ion contamination, in contrast to our [64Cu]Cu-ATSM. This high inhibition of DOTATATE results from the fact that the DO3A moiety of DOTATATE can coordinate to a wide range of M2+ and M3+ ions with high formation constants, reducing the selectivity of DOTATATE.
A limitation of this study is that the formation constant of the ATSM complex was not determined. However, these values can be referred to for the same coordination atom. The formation constant, Kf, for Ni2+ and Cu2+ complexes of 1,10-diaza-4,7-dithiadecane (Figure S3) has been reported; the log Kf value is 7.41 for Ni2+ and 10.70 for Cu2+ [41]. These values indicated that the ligand formed a more stable (approximately 1000 times) complex with Cu2+ than with Ni2+. The values of ATSM can be larger than 1,10-dizaza-4,7-dithiodecane because of the rigid framework of the ligand, and the same stability relationship holds for ATSM. The Irving–Williams series can interpret this stability relationship. The Irving–Williams series shows that the order of stability of the divalent transition metal complex of a given ligand is Mn2+ < Fe2+ < Co2+ < Ni2+ < Cu2+ > Zn2+ [42]. The hard and soft acid and base theory [43] can also explain the large formation constant of Cu-ATSM. Fe2+, Ni2+, Cu2+, and Zn2+ are Lewis acids that are intermediates between hard and soft acids, and Cu2+ has the highest electronegativity and polarizability among these ions. This indicates that Cu2+ is the softest acid in the series and has the highest affinity for soft Lewis bases containing sulfur atoms. Thus, ATSM has the highest affinity for Cu2+, and the inhibition of the formation of [64Cu]Cu-ATSM by Ni2+, Fe2+, and Zn2+ can be negligible in future manufacturing practices.

4. Materials and Methods

4.1. Reagents and Materials

All reagents and solvents from commercial sources (Fujifilm Wako Pure Chemical, Osaka, Japan) were of high quality and were used without further purification. ATSM and Cu-ATSM were purchased from ABX Advanced Biochemical Laboratories (Radeberg, Germany). Japan Pharmacopoeia grade distilled water (Otsuka Pharmaceutical Factory, Tokyo, Japan) was used for preparation of aqueous solutions. Spectrochemical-analysis-grade dimethyl sulfoxide (DMSO) was purchased from Fujifilm Wako Pure Chemical Co., Ltd., Osaka, Japan. The Ni2+ complex of ATSM was prepared following a previously reported method [44].

4.2. Determination of Radiochemical Yield

The radiochemical yield of [64Cu]Cu-ATSM was determined by radio-TLC. One microliter of the sample solution was spotted onto an HPTLC silica gel 60 F254 glass plate (Merck Millipore, Burlington, MA, USA) and developed using methanol. Radioactivity on the TLC plates was measured using a radio-TLC system (Raytest PET miniGITA Star; Elysia s.a., Liège, Belgium).

4.3. Experiment 1: Effect of Trace Metal Impurities on the Formation of [64Cu]Cu-ATSM Complex

The experiments were performed on a scale of 1/400 of the clinical formulation, and the concentrations of the reagents, including 64Cu and ATSM, were kept constant.
Cyclotron-produced 64CuCl2 (29.25 MBq; 3.20 pmol) was dissolved in 20 μL of 0.125 M aqueous glycine solution, and 0.5 μL of 0.5 mM ATSM in DMSO was added and mixed well. Then, 20 μL of each metal aqueous solution (Cu2+, Ni2+, Zn2+, and Fe2+) at nine different concentrations (0.0125, 0.025, 0.05, 0.1, 0.2, 0.4, 4, 10, and 100 ppm in 40.5 μL of the reaction mixture) were added. The mixed solutions were incubated at room temperature for 10 min. The radiochemical yield of [64Cu]Cu-ATSM was determined by radio-TLC. The effect of adding these metal solutions was evaluated by plotting the radiochemical yield against the concentration of metal ions as the sum of the added metal ions and metal impurities derived from the target system.

4.4. Experiment 2: Chelate Formation of ATSM with Metal Ions

Briefly, 300 μM of metal solutions (Cu2+, Ni2+, Zn2+, and Fe2+) were prepared by dissolving their chlorides into DMSO, and 300 µL of each solution was mixed with 300 µL of 300 μM ATSM in DMSO. UV-Vis spectra were measured at 5, 15, 30, 1, 2, 5, and 24 h after mixing. Spectra were measured using a NanoDrop One Microdrop UV-Vis Spectrometer (ND-ONE-W; Thermo Fisher Scientific, Waltham, MA, USA) with 1 μL of each mixed solution. The molar absorption coefficient (ε) of Cu-ATSM and Ni-ATSM was determined on 100 μM DMSO solutions of Cu-ATSM and Ni-ATSM. The concentrations of the Cu-ATSM and Ni-ATSM in each reaction mixture were analyzed based on the obtained ε values.

4.5. Experiment 3: Chelate Formation of Zn-ATSM in the Presence of Brønsted Base

The spectrum of the reaction between ATSM and Zn2+ remained almost unchanged for 24 h in Experiment 2. One possible reason for the slow reaction is that deprotonation of ATSM is required to coordinate with Zn2+. Therefore, we confirmed whether the reaction rate of Zn-ATSM is changed by using the Brønsted base. A total of 600 µL of DMSO solution containing 200 µL of a ZnCl2 (300 µM), 200 µL of ATSM (300 µM), and 200 µL of sodium methoxide (600 µM) was stirred at room temperature. The UV-Vis spectra were measured at 15 min, 30 min, 1 h, and 20 h after mixing. The spectra were measured using the same method as in Experiment 2.

4.6. Experiment 4: Differences in the Formation Constants of M-ATSM in the Reaction of ATSM with Cu2+ and Ni2+ Mixtures

Specifically, DMSO solutions of 150 µM CuCl2 (200 µL) and 150 µM NiCl2 (200 µL) were mixed, and 200 µL solution of ATSM in DMSO (150 µM) was added at room temperature. Changes in UV-Vis spectra were measured at 5 min, 1 h, and 2 h after mixing. The spectra were measured using the same method as in Experiment 2.

5. Conclusions

This study showed that 1.2 mol Cu2+ or 288 mol Ni2+, equivalent to ATSM, decreased the labeling yield of [64Cu]Cu-ATSM below 90%. Zn2+ and Fe2+ had little effect on the labeling of [64Cu]Cu-ATSM, even with the addition of 248 or 290 mol equivalents of ATSM. Based on these data, we conclude that trace amounts of Ni2+, Zn2+, and Fe2+ have little effect on the quality of [64Cu]Cu-ATSM; however, the concentration of Cu2+ must be controlled. These results can provide process management tools for radiopharmaceuticals.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph17010010/s1, Figure S1: Molar absorption coefficients (ε) of Cu-ATSM and Ni-ATSM were determined by using UV-Vis absorption spectra. (a): UV-Vis absorption spectra of Cu-ATSM, (b) UV-Vis absorption spectra of Ni-ATSM; Figure S2: Structure of DOTATATE; Figure S3: The structure of 1,10-dizaza-4,7-dithiodecane (log Kf value is 7.41 for Ni2+ and 10.70 for Cu2+).

Author Contributions

Conceptualization, M.S., H.M. and Y.Y.; methodology, M.S., H.M. and Y.Y.; formal analysis, M.S., H.S. (Hisashi Suzuki), M.H., C.I., H.M., M.T., F.H., T.T., C.S., M.-R.Z., H.K. and Y.Y.; investigation, M.S., C.I., H.M., M.T., F.H., T.T. and Y.Y.; writing—original draft preparation, M.S., H.M. and Y.Y.; writing—review and editing, H.S. (Hisashi Suzuki), M.H., M.T. M.-R.Z., T.H., H.S. (Hidemitsu Sato), H.K. and Y.D.; supervision, Y.Y. and Y.D.; funding acquisition, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the JST FOREST Program, Grant Number JPMJFR2116, and the Japan Agency for Medical Research and Development (AMED) programs of Practical Research for Innovative Cancer Control [grant number 20ck0106567s0201].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Material.

Conflicts of Interest

H.M., H.K., and Y.Y. are cofounders and board members of LinqMed Inc. C.I., F.H., and T.T. are employees of LinqMed Inc. The other authors declare no conflicts of interest. The funders had no role in the study design; collection, analyses, or interpretation of data; writing of the manuscript; or decision to publish the results.

References

  1. Hsieh, C.H.; Shyu, W.C.; Chiang, C.Y.; Kuo, J.W.; Shen, W.C.; Liu, R.S. NADPH oxidase subunit 4-mediated reactive oxygen species contribute to cycling hypoxia-promoted tumor progression in glioblastoma multiforme. PLoS ONE 2011, 6, e23945. [Google Scholar] [CrossRef] [PubMed]
  2. Rong, Y.; Durden, D.L.; Van Meir, E.G.; Brat, D.J. “Pseudopalisading” necrosis in glioblastoma: A familiar morphologic feature that links vascular pathology, hypoxia, and angiogenesis. J. Neuropathol. Exp. Neurol. 2006, 65, 529–539. [Google Scholar] [CrossRef] [PubMed]
  3. Merighi, S.; Benini, A.; Mirandola, P.; Gessi, S.; Varani, K.; Leung, E.; Maclennan, S.; Borea, P.A. Adenosine modulates vascular endothelial growth factor expression via hypoxia-inducible factor-1 in human glioblastoma cells. Biochem. Pharmacol. 2006, 72, 19–31. [Google Scholar] [CrossRef] [PubMed]
  4. Lopez-Lazaro, M. Hypoxia-inducible factor 1 as a possible target for cancer chemoprevention. Cancer Epidemiol. Biomark. Prev. 2006, 15, 2332–2335. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, W.J.; Chen, W.W.; Zhang, X. Glioblastoma multiforme: Effect of hypoxia and hypoxia inducible factors on therapeutic approaches. Oncol. Lett. 2016, 12, 2283–2288. [Google Scholar] [CrossRef] [PubMed]
  6. Takahashi, N.; Fujibayashi, Y.; Yonekura, Y.; Welch, M.J.; Waki, A.; Tsuchida, T.; Sadato, N.; Sugimoto, K.; Itoh, H. Evaluation of 62Cu labeled diacetyl-bis(N4-methylthiosemicarbazone) as a hypoxic tissue tracer in patients with lung cancer. Ann. Nucl. Med. 2000, 14, 323–328. [Google Scholar] [CrossRef] [PubMed]
  7. Lewis, J.S.; Laforest, R.; Dehdashti, F.; Grigsby, P.W.; Welch, M.J.; Siegel, B.A. An imaging comparison of 64Cu-ATSM and 60Cu-ATSM in cancer of the uterine cervix. J. Nucl. Med. 2008, 49, 1177–1182. [Google Scholar] [CrossRef] [PubMed]
  8. Floberg, J.M.; Wang, L.; Bandara, N.; Rashmi, R.; Mpoy, C.; Garbow, J.R.; Rogers, B.E.; Patti, G.J.; Schwarz, J.K. Alteration of Cellular Reduction Potential Will Change 64Cu-ATSM Signal with or without Hypoxia. J. Nucl. Med. 2020, 61, 427–432. [Google Scholar] [CrossRef]
  9. Green, M.A.; Mathias, C.J.; Smith, N.J.; Cheng, M.; Hutchins, G.D. In Vivo Quantitative Whole-Body Perfusion Imaging Using Radiolabeled Copper(II) Bis(Thiosemicarbazone) Complexes and Positron Emission Tomography (PET). Methods Mol. Biol. 2022, 2393, 751–771. [Google Scholar] [CrossRef]
  10. Liu, T.; Karlsen, M.; Karlberg, A.M.; Redalen, K.R. Hypoxia imaging and theranostic potential of [64Cu][Cu(ATSM)] and ionic Cu(II) salts: A review of current evidence and discussion of the retention mechanisms. EJNMMI Res. 2020, 10, 33. [Google Scholar] [CrossRef]
  11. Maitz, C.A.; Tate, D.; Bechtel, S.; Lunceford, J.; Henry, C.; Flesner, B.; Collins, A.; Varterasian, M.; Tung, D.; Zhang, L.; et al. Paired 18F-Fluorodeoxyglucose (18F-FDG), and 64Cu-Copper(II)-diacetyl-bis(N(4)-methylthiosemicarbazone) (64Cu-ATSM) PET Scans in Dogs with Spontaneous Tumors and Evaluation for Hypoxia-Directed Therapy. Radiat. Res. 2022, 197, 253–260. [Google Scholar] [CrossRef] [PubMed]
  12. Okazawa, H.; Ikawa, M.; Tsujikawa, T.; Mori, T.; Makino, A.; Kiyono, Y.; Nakamoto, Y.; Kosaka, H.; Yoneda, M. Cerebral Oxidative Stress in Early Alzheimer’s Disease Evaluated by 64Cu-ATSM PET/MRI: A Preliminary Study. Antioxidants 2022, 11, 1022. [Google Scholar] [CrossRef]
  13. Peres, E.A.; Toutain, J.; Paty, L.P.; Divoux, D.; Ibazizene, M.; Guillouet, S.; Barre, L.; Vidal, A.; Cherel, M.; Bourgeois, M.; et al. 64Cu-ATSM/64Cu-Cl2 and their relationship to hypoxia in glioblastoma: A preclinical study. EJNMMI Res. 2019, 9, 114. [Google Scholar] [CrossRef] [PubMed]
  14. Tsuchiya, J.; Yoneyama, T.; Ohtake, M.; Tateishi, K.; Bae, H.; Kishino, M.; Tateishi, U. Redox reaction and clinical outcome of primary diffuse large B-cell lymphoma of the central nervous system: Prognostic role of metabolic and textural parameters of 62Cu-diacetyl-bis (N4-methylthiosemicarbazone) PET/computed tomography in a small patient cohort. Nucl. Med. Commun. 2020, 41, 567–574. [Google Scholar] [CrossRef] [PubMed]
  15. Dearling, J.L.; Packard, A.B. Some thoughts on the mechanism of cellular trapping of Cu(II)-ATSM. Nucl. Med. Biol. 2010, 37, 237–243. [Google Scholar] [CrossRef] [PubMed]
  16. Obata, A.; Yoshimi, E.; Waki, A.; Lewis, J.S.; Oyama, N.; Welch, M.J.; Saji, H.; Yonekura, Y.; Fujibayashi, Y. Retention mechanism of hypoxia selective nuclear imaging/radiotherapeutic agent cu-diacetyl-bis(N4-methylthiosemicarbazone) (Cu-ATSM) in tumor cells. Ann. Nucl. Med. 2001, 15, 499–504. [Google Scholar] [CrossRef] [PubMed]
  17. Holland, J.P.; Giansiracusa, J.H.; Bell, S.G.; Wong, L.L.; Dilworth, J.R. In vitro kinetic studies on the mechanism of oxygen-dependent cellular uptake of copper radiopharmaceuticals. Phys. Med. Biol. 2009, 54, 2103–2119. [Google Scholar] [CrossRef] [PubMed]
  18. Lewis, J.S.; McCarthy, D.W.; McCarthy, T.J.; Fujibayashi, Y.; Welch, M.J. Evaluation of 64Cu-ATSM in vitro and in vivo in a hypoxic tumor model. J. Nucl. Med. 1999, 40, 177–183. [Google Scholar]
  19. Maurer, R.I.; Blower, P.J.; Dilworth, J.R.; Reynolds, C.A.; Zheng, Y.; Mullen, G.E. Studies on the mechanism of hypoxic selectivity in copper bis(thiosemicarbazone) radiopharmaceuticals. J. Med. Chem. 2002, 45, 1420–1431. [Google Scholar] [CrossRef]
  20. Anderson, C.J.; Ferdani, R. Copper-64 radiopharmaceuticals for PET imaging of cancer: Advances in preclinical and clinical research. Cancer Biother. Radiopharm. 2009, 24, 379–393. [Google Scholar] [CrossRef]
  21. McMillan, D.D.; Maeda, J.; Bell, J.J.; Genet, M.D.; Phoonswadi, G.; Mann, K.A.; Kraft, S.L.; Kitamura, H.; Fujimori, A.; Yoshii, Y.; et al. Validation of 64Cu-ATSM damaging DNA via high-LET Auger electron emission. J. Radiat. Res. 2015, 56, 784–791. [Google Scholar] [CrossRef] [PubMed]
  22. Yoshii, Y.; Matsumoto, H.; Yoshimoto, M.; Zhang, M.R.; Oe, Y.; Kurihara, H.; Narita, Y.; Jin, Z.H.; Tsuji, A.B.; Yoshinaga, K.; et al. Multiple Administrations of 64Cu-ATSM as a Novel Therapeutic Option for Glioblastoma: A Translational Study Using Mice with Xenografts. Transl. Oncol. 2018, 11, 24–30. [Google Scholar] [CrossRef] [PubMed]
  23. Matsumoto, H.; Igarashi, C.; Kaneko, E.; Hashimoto, H.; Suzuki, H.; Kawamura, K.; Zhang, M.R.; Higashi, T.; Yoshii, Y. Process development of [64Cu]Cu-ATSM: Efficient stabilization and sterilization for therapeutic applications. J. Radioanal. Nucl. Chem. 2019, 322, 467–475. [Google Scholar] [CrossRef]
  24. Igarashi, C.; Matsumoto, H.; Takahashi, M.; Hihara, F.; Tachibana, T.; Zhang, M.R.; Kurihara, H.; Higashi, T.; Yoshii, Y. Identification and quantitative structure-activity relationship assessment of trace chemical impurities contained in the therapeutic formulation of [64Cu]Cu-ATSM. Nucl. Med. Biol. 2022, 108–109, 10–15. [Google Scholar] [CrossRef] [PubMed]
  25. Alliot, C.; Michel, N.; Bonraisin, A.-C.; Bossé, V.; Laizé, J.; Bourdeau, C.; Mokili, B.M.; Haddad, F. One step purification process for no-carrier-added 64Cu produced using enriched nickel target. Radiochim. Acta 2011, 99, 627–630. [Google Scholar] [CrossRef]
  26. Avila-Rodriguez, M.A.; Nye, J.A.; Nickles, R.J. Simultaneous production of high specific activity 64Cu and 61Co with 11.4 MeV protons on enriched 64Ni nuclei. Appl. Radiat. Isot. 2007, 65, 1115–1120. [Google Scholar] [CrossRef]
  27. McCarthy, D.W.; Shefer, R.E.; Klinkowstein, R.E.; Bass, L.A.; Margeneau, W.H.; Cutler, C.S.; Anderson, C.J.; Welch, M.J. Efficient production of high specific activity 64Cu using a biomedical cyclotron. Nucl. Med. Biol. 1997, 24, 35–43. [Google Scholar] [CrossRef]
  28. Obata, A.; Kasamatsu, S.; McCarthy, D.W.; Welch, M.J.; Saji, H.; Yonekura, Y.; Fujibayashi, Y. Production of therapeutic quantities of 64Cu using a 12 MeV cyclotron. Nucl. Med. Biol. 2003, 30, 535–539. [Google Scholar] [CrossRef]
  29. Ferreira, C.L.; Yapp, D.T.; Lamsa, E.; Gleave, M.; Bensimon, C.; Jurek, P.; Kiefer, G.E. Evaluation of novel bifunctional chelates for the development of Cu-64-based radiopharmaceuticals. Nucl. Med. Biol. 2008, 35, 875–882. [Google Scholar] [CrossRef]
  30. Boswell, C.A.; McQuade, P.; Weisman, G.R.; Wong, E.H.; Anderson, C.J. Optimization of labeling and metabolite analysis of copper-64-labeled azamacrocyclic chelators by radio-LC-MS. Nucl. Med. Biol. 2005, 32, 29–38. [Google Scholar] [CrossRef]
  31. Zeng, D.; Anderson, C.J. Rapid and sensitive LC-MS approach to quantify non-radioactive transition metal impurities in metal radionuclides. Chem. Commun. 2013, 49, 2697–2699. [Google Scholar] [CrossRef] [PubMed]
  32. Ohya, T.; Nagatsu, K.; Suzuki, H.; Fukada, M.; Minegishi, K.; Hanyu, M.; Fukumura, T.; Zhang, M.R. Efficient preparation of high-quality 64Cu for routine use. Nucl. Med. Biol. 2016, 43, 685–691. [Google Scholar] [CrossRef] [PubMed]
  33. Holland, J.P.; Barnard, P.J.; Collison, D.; Dilworth, J.R.; Edge, R.; Green, J.C.; McInnes, E.J. Spectroelectrochemical and computational studies on the mechanism of hypoxia selectivity of copper radiopharmaceuticals. Chemistry 2008, 14, 5890–5907. [Google Scholar] [CrossRef] [PubMed]
  34. Jones, C.J.; McCleverty, J.A. Complexes of transition metals with Schiff bases and the factors influencing their redox properties. Part I. Nickel and copper complexes of some diketone bisthiosemicarbazones. J. Chem. Soc. A Inorg. Phys. Theor. 1970, 2829–2836. [Google Scholar] [CrossRef]
  35. Thisgaard, H.; Jensen, M.; Elema, D.R. Medium to large scale radioisotope production for targeted radiotherapy using a small PET cyclotron. Appl. Radiat. Isot. 2011, 69, 1–7. [Google Scholar] [CrossRef] [PubMed]
  36. Tochon-Danguy, H.; Poniger, S.; Sachinidis, J.; Alt, K.; Scott, A.; Hagemeyer, C. Reducing Metal Contamination in Cu-64 Production; Helmholtz-Zentrum Dresden: Rossendorf, Germany, 2014. [Google Scholar]
  37. Szajek, L.P.; Meyer, W.; Plascjak, P.; Eckelman, W.C. Semi-remote production of [64Cu]CuCl2 and preparation of high specific activity [64Cu]Cu-ATSM for PET studies. Radiochim. Acta 2005, 93, 239–244. [Google Scholar] [CrossRef]
  38. Thieme, S.; Walther, M.; Pietzsch, H.J.; Henniger, J.; Preusche, S.; Mading, P.; Steinbach, J. Module-assisted preparation of 64Cu with high specific activity. Appl. Radiat. Isot. 2012, 70, 602–608. [Google Scholar] [CrossRef]
  39. Shigenobu, F.; Yasuhiro, I. Solvation Structures and Solvent Exchange Reactions of Metal Ions in Various Coordinating Solvents. Bull. Chem. Soc. Jpn. 2002, 75, 1901–1925. [Google Scholar] [CrossRef]
  40. Walczak, R.; Gawęda, W.; Dudek, J.; Choiński, J.; Bilewicz, A. Influence of metal ions on the 44Sc-labeling of DOTATATE. J. Radioanal. Nucl. Chem. 2019, 322, 249–254. [Google Scholar] [CrossRef]
  41. Martell, A.E.; Smith, R.M. Critical Stability Constants: First Supplement; Plenum Press: New York, NY, USA, 1974; pp. 132–200. [Google Scholar]
  42. Irving, H.; Williams, R.J.P. 637. The stability of transition-metal complexes. J. Chem. Soc. (Resumed) 1953, 5, 3192–3210. [Google Scholar] [CrossRef]
  43. Pearson, R.G. Hard and Soft Acids and Bases. J. Am. Chem. Soc. 1963, 85, 3533–3539. [Google Scholar] [CrossRef]
  44. Douglas, X.W.; Janeice, S.I.; Gordon, A.B.; Anthony, E.L.; Jesús, V.-M.; Klaus, H.E.; Simón, H.-O. Copper(II) and nickel(II) complexes of 2,3-butanedione bis(N(3)-substituted thiosemicarbazones). Polyhedron 1997, 16, 1895–1905. [Google Scholar] [CrossRef]
Figure 1. ATSM (free ligand form) and its Cu2+ and Ni2+ complexes.
Figure 1. ATSM (free ligand form) and its Cu2+ and Ni2+ complexes.
Pharmaceuticals 17 00010 g001
Figure 2. Effect of radiochemical yield of [64Cu]Cu-ATSM on coexisting metal ions. (a): Cu2+, (b): Ni2+ (c): Zn2+, (d): Fe2+.
Figure 2. Effect of radiochemical yield of [64Cu]Cu-ATSM on coexisting metal ions. (a): Cu2+, (b): Ni2+ (c): Zn2+, (d): Fe2+.
Pharmaceuticals 17 00010 g002
Figure 3. Variation in UV-Vis spectra of DMSO solution containing 300 µM MCl2 and 300 µM-ATSM at room temperature with time. A 300 µM (90 nmol) MCl2 concentration equals an M2+ concentration of Cu2+ (19.1 ppm), Ni2+ (17.6 ppm), Zn2+(19.6 ppm), and Fe2+ (16.8 ppm). (a): Cu-ATSM, (b): Ni-ATSM, (c): Zn-ATSM, (d): Fe-ATSM.
Figure 3. Variation in UV-Vis spectra of DMSO solution containing 300 µM MCl2 and 300 µM-ATSM at room temperature with time. A 300 µM (90 nmol) MCl2 concentration equals an M2+ concentration of Cu2+ (19.1 ppm), Ni2+ (17.6 ppm), Zn2+(19.6 ppm), and Fe2+ (16.8 ppm). (a): Cu-ATSM, (b): Ni-ATSM, (c): Zn-ATSM, (d): Fe-ATSM.
Pharmaceuticals 17 00010 g003
Figure 4. Time-dependent increase in M-ATSM concentration during the reaction of 300 μM MCl2 with 300 μM ATSM in DMSO at room temperature. (a) Cu-ATSM, (b) Ni-ATSM.
Figure 4. Time-dependent increase in M-ATSM concentration during the reaction of 300 μM MCl2 with 300 μM ATSM in DMSO at room temperature. (a) Cu-ATSM, (b) Ni-ATSM.
Pharmaceuticals 17 00010 g004
Figure 5. Time-dependent change in the UV-Vis spectra of DMSO solution containing 300 µM ZnCl2, 300 µM ATSM, and 600 µM sodium methoxide at room temperature.
Figure 5. Time-dependent change in the UV-Vis spectra of DMSO solution containing 300 µM ZnCl2, 300 µM ATSM, and 600 µM sodium methoxide at room temperature.
Pharmaceuticals 17 00010 g005
Figure 6. Time-dependent changes in the UV-Vis spectra of ATSM after simultaneously mixing with Cu2+ and Ni2+.
Figure 6. Time-dependent changes in the UV-Vis spectra of ATSM after simultaneously mixing with Cu2+ and Ni2+.
Pharmaceuticals 17 00010 g006
Table 1. Effect of coexisting transition metal ions (Cu2+ and Ni2+) on radiochemical yield of [64Cu]Cu-ATSM. The contents of metal ions in the reaction mixture, shown as the unit of ppm and mol in parenthesis, are indicated as the sum of added metal ions and metal impurities derived from the target system.
Table 1. Effect of coexisting transition metal ions (Cu2+ and Ni2+) on radiochemical yield of [64Cu]Cu-ATSM. The contents of metal ions in the reaction mixture, shown as the unit of ppm and mol in parenthesis, are indicated as the sum of added metal ions and metal impurities derived from the target system.
Cu2+Ni2+
Metal Ion in ppm
(mol)
Radiochemical Yield (%)Ratio (1)Metal Ion in ppm
(mol)
Radiochemical Yield (%)Ratio
0.073
(1.9 × 10−8)
98.80.190.0125
(3.4 × 10−9)
99.00.036
0.086
(2.2 × 10−8)
97.40.220.025
(6.9 × 10−9)
98.80.072
0.11
(2.9 × 10−8)
96.60.280.05
(1.4 × 10−8)
97.70.14
0.16
(4.1 × 10−8)
94.80.410.1
(2.8 × 10−8)
98.20.29
0.26
(6.7 × 10−8)
91.00.660.2
(5.5 × 10−8)
98.40.58
0.46
(1.2 × 10−7)
88.81.20.4
(1.1 × 10−7)
98.21.2
4.1
(1.0 × 10−6)
25.0104
(1.1 × 10−6)
98.312
10
(2.6 × 10−6)
8.32610
(2.8 × 10−6)
94.929
100
(2.6 × 10−5)
1.0255100
(2.8 × 10−5)
85.9288
(1) Metal ion (mol)/ATSM (mol).
Table 2. Effect of coexisting transition metal ions (Zn2+ and Fe2+) on radiochemical yield of [64Cu]Cu-ATSM. The contents of metal ions in the reaction mixture, shown as the unit of ppm and mol in parenthesis, are indicated as the sum of added metal ions and metal impurities derived from the target system.
Table 2. Effect of coexisting transition metal ions (Zn2+ and Fe2+) on radiochemical yield of [64Cu]Cu-ATSM. The contents of metal ions in the reaction mixture, shown as the unit of ppm and mol in parenthesis, are indicated as the sum of added metal ions and metal impurities derived from the target system.
Zn2+Fe2+
Metal Ion in ppm
(mol)
Radiochemical Yield (%)RatioMetal Ion in ppm
(mol)
Radiochemical Yield (%)Ratio
0.064
(1.6 × 10−8)
93.50.160.030
(8.8 × 10−9)
98.50.088
0.077
(1.9 × 10−8)
94.90.190.043
(1.2 × 10−8)
96.40.12
0.10
(2.5 × 10−8)
93.70.250.068
(2.0 × 10−8)
96.70.20
0.15
(3.8 × 10−8)
96.10.380.12
(3.4 × 10−8)
94.60.34
0.25
(6.3 × 10−8)
95.70.620.22
(6.3 × 10−8)
96.60.63
0.45
(1.1 × 10−7)
94.11.10.42
(1.2 × 10−7)
95.41.2
4.1
(1.0 × 10−6)
98.7104.0
(1.2 × 10−6)
96.412
10
(2.5 × 10−6)
98.02510
(2.9 × 10−6)
96.429
100
(2.5 × 10−5)
96.1248100
(2.9 × 10−5)
96.4290
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shinada, M.; Suzuki, H.; Hanyu, M.; Igarashi, C.; Matsumoto, H.; Takahashi, M.; Hihara, F.; Tachibana, T.; Sogawa, C.; Zhang, M.-R.; et al. Trace Metal Impurities Effects on the Formation of [64Cu]Cu-diacetyl-bis(N4-methylthiosemicarbazone) ([64Cu]Cu-ATSM). Pharmaceuticals 2024, 17, 10. https://doi.org/10.3390/ph17010010

AMA Style

Shinada M, Suzuki H, Hanyu M, Igarashi C, Matsumoto H, Takahashi M, Hihara F, Tachibana T, Sogawa C, Zhang M-R, et al. Trace Metal Impurities Effects on the Formation of [64Cu]Cu-diacetyl-bis(N4-methylthiosemicarbazone) ([64Cu]Cu-ATSM). Pharmaceuticals. 2024; 17(1):10. https://doi.org/10.3390/ph17010010

Chicago/Turabian Style

Shinada, Mitsuhiro, Hisashi Suzuki, Masayuki Hanyu, Chika Igarashi, Hiroki Matsumoto, Masashi Takahashi, Fukiko Hihara, Tomoko Tachibana, Chizuru Sogawa, Ming-Rong Zhang, and et al. 2024. "Trace Metal Impurities Effects on the Formation of [64Cu]Cu-diacetyl-bis(N4-methylthiosemicarbazone) ([64Cu]Cu-ATSM)" Pharmaceuticals 17, no. 1: 10. https://doi.org/10.3390/ph17010010

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop