Skip to main content
  • Research article
  • Open access
  • Published:

Automated, cassette-based isolation and formulation of high-purity [61Cu]CuCl2 from solid Ni targets

Abstract

Background

A need for improved, cassette-based automation of 61Cu separation from irradiated Ni targets was identified given the growing interest in theranostics, and generally lengthy separation chemistries for 64Cu/64Ni, upon which 61Cu chemistry is often based.

Methods

A method for separating 61Cu from irradiated natNi targets was therefore developed, with provision for target recycling. Following deuteron irradiation, electroplated natNi targets were remotely transferred from the cyclotron and dissolved in acid. The dissolved target solution was then transferred to an automated FASTlab chemistry module, where sequential TBP and TK201 (Triskem) resins isolated the [61Cu]CuCl2, removed Ni, Co, and Fe, and concentrated the product into a formulation suitable for anticipated radiolabelling reactions.

Results

61Cu saturation yields of 190 ± 33 MBq/μA from energetically thick natNi targets were measured. The average, decay-corrected, activity-based dissolution efficiency was 97.5 ± 1.4% with an average radiochemical yield of 90.4 ± 3.2% (N = 5). The isolated activity was collected approximately 65 min post end of bombardment in ~ 2 mL of 0.06 M HCl (HCl concentration was verified by titration). Quality control of the isolated [61Cu]CuCl2 (N = 5) measured 58Co content of (8.3 ± 0.6) × 10− 5% vs. 61Cu by activity, Ni separation factors ≥ (2.2 ± 1.8) × 106, EoB molar activities 85 ± 23 GBq/μmol and NOTA-based EoB apparent molar activities of 31 ± 8 MBq/nmol and 201 MBq/nmol for the 30 min and 3.3 h (N = 1) irradiations, respectively.

Conclusion

High purity 61Cu was produced with the developed automated method using a single-use, cassette-based approach. It was also applicable for 64Cu, as demonstrated with a single proof-of-concept 64Ni target production run.

Highlight statements

  • A rapid, automated, and highly repeatable procedure for isolating [61Cu]CuCl2 from cyclotron-irradiated Ni targets was developed.

  • The suitable formulation of the product facilitates labelling experiments.

  • Enables routine production of 61Cu with NOTA-based AMAs of 200 MBq/nmol [5 Ci/μmol] from natNi solid targets.

Introduction

Within the field of positron emission tomography (PET), radiometals research has increased during the past decade (e.g. publications per year with the keyword 68Ga have increased from 69 to 506 between 2009 and 2019, Scopus). These radiometals complement the traditional PET nuclide 18F thanks to differences in their chemical and radioactive decay properties. This opens a path toward more personalised medicine as an array of biomolecules and biodistribution mechanisms can be used to adapt a treatment to a specific disease case. Radiometals are widely used for radiopharmaceuticals, in part due to available chelation chemistry and labelling with biomolecules (Krasikova et al. 2016; Price and Orvig 2014; Aluicio-Sarduy et al. 2018; Vivier et al. 2018; Tsai and Wu 2018; Liu 2008). Among the radiometals, following Ga, Cu is one of the most extensively investigated for PET radiopharmaceutical purposes (Mikolajczak et al. 2019; Brandt et al. 2018; Wadas et al. 2006; McCarthy et al. 1999). One reason for this is the well-understood coordination chemistry and biodistribution of Cu (Wadas et al. 2006; Wadas et al. 2010; Wallhaus et al. 1998; Jalilian et al. 2009; Woo et al. 2019), which has resulted in a multitude of chelators and biomolecule options being available for Cu isotopes.

While 64Cu has been suggested for theranostic applications, including pairing with 177Lu (Song et al. 2016), several Cu radioisotopes are suitable for both imaging and therapy. This creates an opportunity for Cu to be used as a “true” (i.e. identical element) theranostic pair: 61Cu (t½ = 3.34 h, 61% β+, EMax = 1216 keV) is suitable for PET imaging; only 5.9 and 2.1% of its two major gammas (282.956 KeV, Iγ = 12.2%; 656.008 keV, Iγ = 10.8%, respectively), are coincident with β+ decay (IAEA 2020a). 64Cu (t½ = 12.7 h) is more commonly used in PET imaging (18% β+, EMax = 653 keV) (Singh et al. 2017; Follacchio et al. 2017) with negligible gamma emissions (1345.77 keV, Iγ = 0.475%), however, applications in β and Auger emission therapy have been reported (39% β, EMax = 580 keV) (Obata et al. 2005; Lewis et al. 1999, 2001; Gutfilen et al. 2018), and 67Cu (t½ = 61.83 h, 100% β, Emax = 562 keV) is a therapeutic radionuclide suitable for SPECT imaging (Eγ = 184.577 keV, Iγ = 48.7%).

61Cu and 64Cu share several physical properties; they are imageable with PET, have half-lives that allow for regional distribution and can both be made from Ni starting material. However, several production paths exist for these isotopes, and depending on the enrichment of the starting material, the production cost will vary significantly. Several 61Cu production routes start from Ni target materials, including the natNi(d,x)61Cu, 60Ni(d,n)61Cu, and 61Ni(p,n)61Cu reactions, as seen in Table 1 alongside typical production routes for 64Cu and 67Cu. Each of these Ni-based 61Cu production routes, e.g. from natNi to 60Ni to 61Ni, increase in theoretical thick target yields, but are met with increased cost of target material (estimated ~$1–2 USD/mg for 60Ni [nat ab. 26.223%], and ~ $30–40 USD/mg for 61Ni [nat ab. 1.1399%]). Enriched Ni options for producing 64Cu are limited to 64Ni (~$30–40 USD/mg for 64Ni [nat ab. 0.9255%]) through 64Ni(p,n)64Cu. This results in a possible lower production cost for 61Cu, compared with 64Cu, by using less expensive enriched options, at the cost of lower theoretical thick target yields.

Table 1 Common production paths and decay properties of 61Cu, 64Cu and 67Cu. All nuclear data was acquired from NuDat (Brookhaven National Laboratory 2020a) and Qcalc (Brookhaven National Laboratory 2020b). “Q” notes q-value

The relatively long, but different, half-lives of 61/64Cu provide an opportunity to study biodistributions of larger molecules with slower kinetics, such as peptides or antibodies (with 61Cu perhaps better suited for same-day imaging and 64Cu allowing for later time-point imaging). However, their application in PET imaging will depend on the purpose of the study, as 61Cu has a higher sensitivity, i.e. 3.43 vs 0.98 cps/Bq/mL (Williams et al. 2005), but slightly lower spatial resolution compared with 64Cu (Williams et al. 2005). The physical decay properties of 64Cu can also result in a relatively larger effective dose compared with 61Cu. E.g. the effective dose for [64Cu]Cu-PTSM and [61Cu]Cu-PTSM as perfusion imaging agents were 3.8 vs 2.5 mSv per 100 MBq respectively, according to Williams et al. (2005). Thus, 61Cu and 64Cu are both valuable diagnostic radionuclides whose applications should be tailored to their physical decay characteristics. However, relatively more papers are published on 64Cu than on 61Cu; 1288 vs 113 hits between 2009 and 2019 for the keywords 64Cu and 61Cu respectively (Scopus) – which likely arises from simplified, distribution-friendly 64Cu logistics. Nevertheless, the increased availability of cyclotron production facilities, increasing solid target infrastructure, and automated radiochemistry systems compel reconsideration of the utility of 61Cu.

Additionally, regardless of the production route, several Co radioisotopic contaminants will be produced, with 55Co, 57Co and 58Co being of greatest interest due to their relatively long half-lives (55Co t½ = 17.5 h, 57Co t½ = 272 d, 58Co t½ = 71 d). Their quantities will depend on the selected reaction, irradiation conditions, target thickness, and isotopic abundance of Ni in the target material. Consequently, in addition to separating Cu from the stable Ni target material, efficient separation of Co is also necessary. As enriched Ni may be cost-prohibitive to implement as single-use, especially 61Ni, target recycling is imperative. Efficient (> 96%) recovery and re-plating processes have been described (Avila-Rodriguez et al. 2007). For this reason, though 61Ni targets were not employed in this study, target recovery/recycling was also investigated, including a preliminary production using 64Ni.

This paper focuses on obtaining high-purity 61Cu via a cassette-based automated separation method using a two-column approach implemented on the FASTlab chemistry platform. A proof-of-concept 64Cu production (N = 1) was performed to demonstrate applicability of this method to the 64Ni(p,n)64Cu reaction. This is interesting as [64Cu]Cu-DOTA-TATE has recently been granted FastTrack review by the US FDAFootnote 1 and has been determined to produce more true-positive lesion detections for neuroendocrine tumours than [68Ga]Ga-DOTA-TOC (Johnbeck et al. 2017). Regardless of radioisotope, the Cu product must consistently be of high radionuclidic purity with a high apparent molar activity (AMA). Thus, it is important to have a robust separation chemistry and rigorous quality control (QC) process to ensure the high quality of the product.

Materials and methods

Bench-top pre-studies

Prior to irradiating electroplated Ni, target dissolution and chemical separation processes were investigated at the bench using a heater block and FASTlab. Dissolution studies, initially on Ni foil, probed the effects of acid concentration, H2O2 to HCl ratio, and temperature on dissolution efficiency, speed, and compatibility with downstream separation chemistry. Separation studies focused on minimizing Ni, Co, and Fe in the final product. These separation studies were performed by spiking dissolved stable Ni with ppm Cu, Co, and Fe, and analysing samples via semi-quantitative colorimetric tests (such as Merck’s Mquant® colorimetric Ni kit, part number 1.14420, which allow for sub-ppm analysis), or, with low activity (kBq) spikes of 61Cu and 55Co, produced by proton (≤ 5 μAmin) and deuteron (≤ 5 μAmin) irradiations of natNi foil. During the benchtop tests, the relative distribution of 61Cu (and similarly 55Co) were determined for the collected fractions with an Ortec LaBr (digibase, brilliance 380) gamma spectrometer.

Target preparation and recycling

Electroplated targets ranging from 70 to 120 mg were prepared by first dissolving natural Ni powder (Alfa Aesar, 99.8%, 325 mesh) in 2 mL of concentrated HNO3 (Optima Grade, Fisher Chemical) and drying down under N2 gas flow at 85 °C. Next, the dried Ni was prepared into an electroplating solution following the method of Piel et al. (1992) to electrodeposit Ni onto gold plating substrates. We adapted this method to electrodeposit onto 99.9% silver plating substrates (10 mm deposited diameter). Briefly, the dried Ni was reconstituted in 2.3 mL of 2.4 M H2SO4 (made from concentrated, Optima Grade H2SO4, Fisher Scientific, and 18 MΩ-cm milli-Q water) and the pH of the solution was brought to ~ 9.1 using ~ 2.5 mL concentrated NH4OH (28%, Optima Grade, Fisher Scientific). To the pH-adjusted solution 250–300 mg of (NH4)2SO4 (99.9999%, Puratronic, Alfa Aesar) was added, and the solution was quantitatively transferred to an electrolytic cell. With a platinum wire cathode and laboratory DC power supply, constant currents of 40–90 mA were tested for optimization purposes, with a voltage of 6–7.5 V applied through the static electrochemical cell for 1–4 days.

A similar setup was used for re-plating of targets following irradiation. Namely, the Ni collection fraction following purification was then dried down under N2 gas flow. As above with the fresh target material, 2 mL of concentrated HNO3 was added and the solution dried again. The dried, recycled Ni was then electroplated on a silver substrate as above for subsequent irradiation. The 64Ni (84.8 mg) was electroplated on a gold plating substrate, according to the method of Piel et al. (1992) directly.

Target irradiation

Electroplated, natNi targets were irradiated with 8.4 MeV deuterons on a PETtrace 800 cyclotron (GE Healthcare) equipped with a QIS (ARTMS) automated target handling system with typical beam currents of 20–30 μA, and typical irradiation times being either ≤30 min for initial tests (N = 3), or, up to 3.3 h (i.e. one half-life) for scaled-up demonstration (N = 1). To enable recycling comparison, both 1× recycled targets (N = 3) and 2× recycled targets (N = 2) were evaluated. For the preliminary enriched target production, the 64Ni was irradiated with nominally 13.1 MeV protons at 20 μA for 1 hour. These experiments are summarized in Table 2.

Table 2 Information compilation of the ten electroplated target irradiations

For optimisation purposes, natNi targets were additionally irradiated for approximately 1 min with 1 μA of protons to produce a radiocobalt tracer via the natNi(p,x)55Co reaction. However, for quantitative analysis, 58Co was used due to its longer half-life compared to 55Co (58Co t½ = 71 d, 55Co t½ = 17.5 h), as 58Co will be predominantly formed through the 60Ni(d,α)58Co reaction.

Dissolution

Following automated transportation of the irradiated natNi target from the cyclotron to the hot cell docking station, the target capsule was transferred to the QIS dissolution unit with tongs. Based on preliminary bench experiments, dissolution was in 3 mL 1:1 7 M HCl (Ultrapur, Merck): 30% H2O2 (ultratrace analysis, Merck) whereby H2O2 was added both to improve the dissolution and to oxidize Fe ions to Fe3+. These two reagents were mixed on-line and circulated over the target surface at 2 mL/min for ~ 23 min at an estimated solution temperature of ~ 60 °C (based on a heater sleeve set point of 111 °C and probing of the heated capsule exterior with a thermocouple). Finally, 3 mL 11.1 M HCl was automatically added to the dissolution solution, with 90 s bubbling with air to mix. Prior to transferring the solution to the FASTlab at 1 mL/min, the sequence was momentarily paused (N = 5), and ~ 200 μL of the nominal 6 mL solution was removed for pre-purification analysis. The irradiated 64Ni target was manually dissolved in a benchtop dissolution block and transferred to the FASTlab using a peristaltic pump at 1 mL/min, as with the automated target handling.

[61Cu]CuCl2 separation

Separation was implemented on a cassette-based FASTlab platform using a 1 mL TBP column (a tributyl-phosphate-based resin, particle size 50–100 μm, pre-packed, Triskem, Britany, Fr) and 2 mL TK201 column (a tertiary-amine-based weak ionic exchange resin containing a small amount of a long-chained alcohol, particle size 50–100 μm, pre-packed, Triskem, Brittany, Fr) automatically conditioned in series with 7 mL H2O and 6 mL 11.1 M HCl. The cassette reagent vials were prepared using concentrated HCl (Optima Grade, Fischer Scientific), NaCl (ACS, Fischer Scientific) and milli-Q water (Millipore system, 18 MΩ-cm resistivity).

A general schematic of the resin loading, washing, and elution steps is given in Fig. 1, with detailed process steps described below:

Fig. 1
figure 1

Two-column approach for 61Cu separation. Process steps described below

Process steps:

  1. 1)

    The acid-adjusted dissolution solution (approx. 6 mL) was loaded over both columns in series and directed into a “Ni collection vial”. The TBP resin acted as a guard column as it quantitatively retained Fe3+ ions, while the Cu and Co chloride complexes were quantitatively retained on the TK201 resin.

  2. 2)

    Both columns were washed with 4 mL 6 M HCl to maximize Ni recovery for future recycling.

  3. 3)

    The TK201 column was washed with 5.5 mL 4.5 M HCl to elute the majority of Co (Waste #1)

  4. 4)

    The TK201 column was washed with 4 mL of 5 M NaCl in 0.05 M HCl to decrease residual acid on the resin and further remove any Co (Waste #2). In the longer term, Waste 1 and Waste 2 will be combined, but were separated for this study for analytical/optimization purposes.

  5. 5)

    The TK201 column was washed with 2 mL of 0.05 M HCl to quantitatively elute the [61Cu]CuCl2

Analysis

Gamma ray spectrometry

For the electroplated target irradiations, 61Cu and 58Co activities were quantified by high purity germanium (HPGe) gamma spectrometry with an Al-windowed Canberra Model GC1519 (15% relative efficiency, full-width at half-maximum at 1173 keV = 1.8 keV) and used to determine the distribution in the samples and fractions. The gammas used for analysis were: 58Co (810.759 keV, Iγ = 99.45%) and 61Cu (primarily 282.956 keV, Iγ = 12.2% and 656.008 keV, Iγ = 10.8%). Samples were counted at a range of distances to the front face of the cylindrical detector body and were selected to maintain the dead time to ≤15%. The energy and efficiency calibration of the detector were performed using a 5-source method: 241Am, 133Ba, 137Cs, 152Eu, and 60Co.

Microwave plasma atomic emission spectrometry

Trace metal standards for Co, Cu, Fe, Ni, and Zn (1000 mg/L) were purchased from Sigma Aldrich. Trace metal analysis was conducted on aliquots of the collected sample fractions using an Agilent Technologies Model 4200 Microwave Plasma Atomic Emission Spectrometer (MP-AES). The concentration of HCl in each analysed sample aliquot was adjusted to 0.5 M. Calibration standards of 10, 50, 100, and 500 ppb and 1, 5, 10, and 50 ppm concentrations containing Co, Cu, Fe, Ni, and Zn were prepared in 0.5 M HCl and quantified using two atomic emission wavelengths for analysis of each element. The optimum wavelengths, depending on signal intensities, limits of detection, standard deviations and obvious interferants, were determined for the different elements post analysis.

The molar activity (MA) and separation factors (SF) of/for the [61Cu]CuCl2 product can be calculated from MP-AES quantified stable Cu and Ni. However, when assessing the separation factors, some samples contained Ni in concentrations below the method detection limit (MDL). When this was the case, our calculations assumed the MDL of 10 ppb as a conservative estimate of the Ni quantity.

Apparent molar activity

The apparent molar activity (AMA) was determined by titration with NOTA and adapted from the process described by McCarthy et al. (1997). Namely, 500 μL of [61Cu]CuCl2 was added to 0.6 mL 0.25 M Sodium acetate (anhydrous, 99% pure, Fisher Scientific), a solution pH of 4–5 was verified using Whatman pH strips, and the solution was vortexed. Next, 100 μL of this activity mixture was added to each of ten vials pre-loaded with both 40 μL 0.25 M sodium acetate (pH 4–5) and 100 μL of NOTA (~ 0.001–10 nmol). Samples were vortexed, individually assayed for activity, and incubated at room temperature for 15 min. Thin-layer chromatography (TLC) was performed by spotting each sample onto an aluminium-backed silica plate, developing in 1:1 MeOH:(10% w/v) NH4OAc and scanning on an OptiQuant autoradiography system (Perkin Elmer Cyclone Plus Storage Phosphor System). In plotting the sigmoidal curve of percent binding vs. NOTA concentration, the NOTA concentration required for 50% binding was identified. The AMA was then calculated as the average sample activity (decay corrected to EoB) and divided by 2× the NOTA concentration required for 50% binding.

Product HCl titration

To evaluate the suitability of the product formulation, and ensure residual HCl was minimized, titrations were performed to determine the HCl concentration of the product fraction. To this end, 0.5 mL of product fraction (N = 5) was added to approximately 10 mL milli-Q water with 100 μL of phenolphthalein in an Erlenmeyer flask with magnetic stir bar, and 5.8 mM NaOH was added dropwise from a burette until a faint pink colour visually persisted in the solution.

Results and discussion

Yields: 61Cu and [61Cu]CuCl2

The overall 61Cu yield was assessed by assaying the activity of the isolated product, waste vials, Ni recovery vial, resins, and target plate post dissolution. This resulted in a saturation yield (± SD) of 207 ± 26 MBq/μA (N = 3) for the initial low current optimization runs and 190 ± 33 MBq/μA (N = 5) for the subsequent 1×/2 × −recycled material. No significant difference was noted in saturation yield of the 61Cu between initial low current and recycled target irradiations. These saturation yields are also a conservative estimate due to the likely presence of unquantified residual activity in the lines and manifold, and possible fractional intercept of the 10 mm diameter Ni plating by the deuteron beam. Nevertheless these saturation yields are considered to be in reasonable agreement with the reported saturation yield (IAEA 2020b) of 248 MBq/μA for 8.4 MeV deuterons on natNi. With regards to chemical processing, the average activity-based dissolution efficiency was 97.5 ± 1.4% (N = 5) with an average radiochemical yield of 90.4 ± 3.2% (N = 5) from the separation and an average dissolution + separation processing time of 65 ± 3 min. Plating efficiencies of 96.0 ± 0.9% (N = 3) were demonstrated for the fresh natNi targets onto silver backings, as calculated from the natNi input and plating masses. For the recycled natNi targets, overall recycling efficiencies, i.e. the percent of Ni recovered and re-plated, were collectively determined to be 88% and 92% respectively for the first and second round of recycling.

Formulation

Select recent examples of [6xCu]CuCl2 purification methods are given in Table 3. Many suffer from lack of automation or have final formulations in large volumes or high acid concentrations. A product with a large volume or high acid concentration may need large buffer quantities or potentially time-consuming reformulation steps which are also subject to transfer losses. The motivation of this work was to address these concerns by developing a fast, efficient, automated process with attractive final formulation qualities.

Table 3 Comparison of requirements for [6xCu]CuCl2 separation vs. other recent literature [reagents for target preparation and recycling not considered]

When stating formulation, one must not assume that the acid concentration of the product is identical to that used for elution. As such, we titrated the HCl concentration of the [61Cu]CuCl2 product to assess the formulation directly. The HCl concentration of the product fraction was assessed on a subset (N = 5) of the product Cu vials (i.e. from the recycled productions) and determined to be 0.057 ± 0.002 M. This low product HCl concentration may circumvent the need for further product processing, such as roto-evaporation and reconstitution. This, in combination with the small product volume (2 mL), facilitates downstream radiolabelling by reducing the need for buffering. As a result, we are able to plan radiolabelling chemistries on the same single-use cassette using the FASTlab radiochemistry system.

Compared with the works cited in Table 3, our separation method is relatively fast – surpassed only by that of Matarrese et al. (2010). However, to achieve adequate labelling conditions, the method of Matarrese et al. would require ~ 40× the buffer than what was used in this work. Our method and the method presented by Ohya et al. (2016, 2019) have similar radiochemical yields. However, their method requires evaporation and has a considerably longer preparation and processing time. Overall, compared with the other methods in Table 3, our automated cassette-based purification method generally has a shorter preparation and process time, lower reagent consumption and more suitable product formulation.

Competing radiocobalt production

As mentioned above, various Co radioisotopes will be produced during irradiations of Ni. McCarthy et al. (1999) notes, for example, 0.05, 0.04, and 0.04% of produced 58Co activity relative to 61Cu for the natNi(d,x)61Cu, enriched 60Ni(d,n)61Cu, and 61Ni(p,n)61Cu reactions, respectively, whereas Strangis and Lepera (2007) notes 0.11% and 0.27% of produced 58Co and 56Co, respectively, relative to 61Cu, for the natNi(d,x)61Cu reaction. The production of different Co isotopes, and relative production vs. 61Cu will depend on various parameters, including the nuclear reaction, the particle irradiation energy and time, target thickness, and isotopic composition. For the 200 μL pre-processed aliquots of dissolved target solution assayed in this study, a pre-purified 58Co to 61Cu activity ratio of 0.0465 ± 0.0046% at EoB, when irradiating for 30 min at 30 μA was determined. This is in line with previous reports (McCarthy et al. 1999; Strangis and Lepera 2007).

A radionuclidic impurity limit of 0.1% by activity at time of validity is currently noted in the European Pharmacopoeia for FDG (European Pharmacopoeia 2014), and (aside for 66/67Ga), for direct accelerator-produced 68Ga (European Pharmacopoeia 2020). Assuming a similar limit of ≤0.1% at time of validity for 61Cu, radiocobalt must be isolated from the 61Cu product to provide [61Cu]CuCl2 with a reasonable shelf-life. Measured 58Co and 61Cu content in the five recycled target separations are presented in Fig. 2 and Table 4 for the collected Ni/waste/production fractions, and resins. The distribution of each nuclide in Fig. 2 has been normalized individually. In this figure, we see that the majority of the 58Co (97.87 ± 0.86%) and 61Cu (90.4 ± 3.2%) are found in the waste and product vials, respectively. From an absolute perspective, the 58Co activity content in the purified [61Cu]CuCl2 product at EoB is (8.3 ± 0.6) × 10− 5%, resulting in a reduction of the 58Co to 61Cu ratio by more than a factor of 500 following purification.

Fig. 2
figure 2

Bar diagram presenting the normalized activity distribution of 58Co and 61Cu. Note that the 58Co and 61Cu have been normalised individually

Table 4 The normalized activity distribution data of 58Co and 61Cu (N = 5), which is visually represented in Fig. 2. Note that the 58Co and 61Cu have been normalised individually

MAs, AMAs, and SFs

For the selected optimum wavelengths (Co 350 nm, Cu 324 nm, Fe 260 nm, Ni 352 nm and Zn 213 nm), the results of the MP-AES analysis largely generated concentrations below the method detection limit (MDL). For the analysed [61Cu]CuCl2 samples, only Cu produced consistent signals above the MDL, and the average product solution’s Cu concentration was determined to be 176 ± 37 ng/mL (N = 5). Co, Zn and Fe all produced signals < MDL (i.e. 100, 100, and 500 ppb respectively).

To assess the chemical purity and applicability of the product, MAs, AMAs, and Ni separation factors were determined and compiled, as reported in Table 5.

Table 5 Compilation of product MAs, AMAs and Ni separation factors for short and long irradiations

For longer irradiations, the MA is anticipated to increase proportionally with the produced activity, as the amount of stable Cu would not be expected to change significantly. Although not explicitly measured, if similar stable Cu (to the 30 min irradiations) is assumed for the 3.3 h irradiation, an estimated MA of 479 GBq/μmol [12.9 Ci/μmol] (N = 1) is calculated from the measured activity of that run at EoB. Additionally, assuming similar starting Ni quality, the MA would be expected to increase nearly 4-fold for enriched 60Ni targets. Therefore, MAs approaching 2 GBq/nmol [50 Ci/μmol] are not unrealistic for enriched 60Ni targets, and even higher MAs may be possible for enriched 61Ni irradiation. With this in mind, the results in Table 5 are promising, particularly when considering that the theoretical maximum MA for 61Cu is 35 GBq/nmol [939 Ci/μmol].

While a high MA is certainly desirable, it does not guarantee a chemically pure product capable of labelling, as other contaminants could be present and compete with the 61Cu for the chelator. A potentially more informative value is the AMA, which considers not only stable Cu, but also competing stable contaminants other than Cu. However, the AMA will depend on the chelator and labelling conditions, so comparing AMAs between different experimental setups is not straightforward.

Proof of concept – 64Cu

To demonstrate applicability to 64Cu, a single test irradiation on enriched 64Ni was additionally performed. We report herein a NOTA-based AMA at EoB of 179 GBq/μmol [4.8 Ci/ μmol], and radiochemical yield of 93%, with an overall processing time of 65 min. Additionally, the 64Ni was recovered post-dissolution and re-plated, resulting in a recycling efficiency of 95%. This was on par with our noted natNi recycling efficiencies between 88 and 92%.

Conclusion

An automated method capable of producing high purity 61Cu from recycled natNi targets has been developed. In summary, the product formulation is 2 mL of < 0.06 M HCl and the entire process can be achieved in ~ 65 min from EoB to EoS, with average radiochemical yields of 90.4 ± 3.2%, and AMAs > 5 Ci/μmol for NOTA when irradiating with 30 μA for 3.3 h. The [61Cu]CuCl2 isolation radiochemistry reported here has a measured separation factor ≥ (2.2 ± 1.8) × 106 for Ni and 58Co radionuclidic impurity of < 0.0001% relative to 61Cu activity. Our process is also directly applicable to the production of 64Cu via the 64Ni(p,n)64Cu reaction, as was demonstrated in a preliminary, enriched 64Ni target irradiation and purification test. Additionally, the repeatable, single-use, cassette-based method is facile to introduce into GMP environments; its final 61Cu radiochemical yield and HCl concentration have relative standard deviations of 3.2 and 3.6%, respectively (N = 5). Finally, the method and FASTlab platform offer the possibility to perform radiolabelling on the same cassette as the separation.

Availability of data and materials

The datasets used and/or analysed during the current study are available from the corresponding author on reasonable request.

Notes

  1. Granted FDA approval September 2020

Abbreviations

AMA:

Apparent Molar Activity

MA:

Molar Activity

MDL:

Method Detection Limit

MP-AES:

Microwave Plasma Atomic Emission Spectrometer

PET:

Positron Emission Tomography

QC:

Quality Control

SF:

Separation Factor

TLC:

Thin-layer Chromatography

References

  • Aluicio-Sarduy E, Ellison PA, Barnhart TE, Cai W, Nickles RJ, Engle JW. PET radiometals for antibody labeling. J Label Compd Radiopharm. 2018;61(9):636–51.

    CAS  Google Scholar 

  • Avila-Rodriguez MA, Nye JA, Nickles RJ. Simultaneous production of high specific activity 64Cu and 61Cu with 11.4 MeV protons on enriched 64Ni nuclei. Appl Radiat Isot. 2007;65(10):1115–20.

    CAS  Google Scholar 

  • Brandt M, Cardinale J, Aulsebrook ML, Gasser G, Mindt TL. An overview of PET radiochemistry, part 2: Radiometals. J Nucl Med. 2018;59(10):1500–6.

    CAS  Google Scholar 

  • Brookhaven National Laboratory. NuDat [Internet]. 2020a [cited 2020 Jun 5]. Available from: https://www.nndc.bnl.gov/nudat2/.

  • Brookhaven National Laboratory. QCalc [Internet]. 2020b [cited 2020 Jun 5]. Available from: https://www.nndc.bnl.gov/qcalc/.

  • European Pharmacopoeia. Fludeoxglucose (18F) injection; 2014. p. 1190–2.

    Google Scholar 

  • European Pharmacopoeia. Monograph 01/2021:3109 gallium (68Ga) chloride (accelerator-produced) solution for radiolabelling. 2020.

    Google Scholar 

  • Follacchio GA, De Feo MS, De Vincentis G, Monteleone F, Liberatore M. Radiopharmaceuticals labelled with copper radionuclides: clinical results in human beings. Curr Radiopharm. 2017;11(1):22–33.

    Google Scholar 

  • Gutfilen B, Souza SAL, Valentini G. Copper-64: a real theranostic agent. Drug Des Devel Ther. 2018;12:3235–45.

    CAS  Google Scholar 

  • IAEA. IAEA nuclear data section, Accessed 18 May 2020 [Internet]. 2020a [cited 2020 May 18]. Available from: https://nds.iaea.org/relnsd/vcharthtml/VChartHTML.html#lastnuc=61Cu.

  • IAEA. IAEA nuclear data section, medical isotope browser, Accessed 5 May 2020 [Internet]. 2020b [cited 2020 May 18]. Available from: https://www-nds.iaea.org/relnsd/isotopia/isotopia.html.

  • Jalilian AR, Rowshanfarzad P, Kamrani YY, Shafaii K, Mirzaii M. Production and tumour uptake of [64Cu]pyruvaldehyde-bis (N 4-methylthiosemicarbazone) for PET and/or therapeutic purposes. Nucl Med Rev. 2007;10(1):6–11.

    Google Scholar 

  • Jalilian AR, Yousefnia H, Faghihi R, Akhlaghi M, Zandi H. Preparation, quality control and biodistribution of [61Cu]-doxorubicin for PET imaging. Nucl Sci Tech. 2009;20(3):157–62.

    CAS  Google Scholar 

  • Johnbeck CB, Knigge U, Loft A, Berthelsen AK, Mortensen J, Oturai P, et al. Head-to-head comparison of 64Cu-DOTATATE and 68Ga-DOTATOC PET/CT: a prospective study of 59 patients with neuroendocrine tumors. J Nucl Med. 2017;58(3):451–7.

    Article  CAS  Google Scholar 

  • Krasikova RN, Aliev RA, Kalmykov SN. The next generation of positron emission tomography radiopharmaceuticals labeled with non-conventional radionuclides. Mendeleev Commun. 2016;26(2):85–94. https://doi.org/10.1016/j.mencom.2016.03.001.

    Article  CAS  Google Scholar 

  • Lewis JS, Laforest R, Buettner TL, Song SK, Fujibayashi Y, Connett JM, et al. Copper-64-diacetyl-bis(N4-methylthiosemicarbazone): an agent for radiotherapy. Proc Natl Acad Sci U S A. 2001;98(3):1206–11.

    Article  CAS  Google Scholar 

  • Lewis JS, Lewis MR, Cutler PD, Srinivasan A, Schmidt MA, Schwarz SW, et al. Radiotherapy and dosimetry of 64Cu-TETA-Tyr3-octreotate in a somatostatin receptor-positive, tumor-bearing rat model. Clin Cancer Res. 1999;5(11):3608–16.

    CAS  Google Scholar 

  • Liu S. Bifunctional coupling agents for radiolabeling of biomolecules and target-specific delivery of metallic radionuclides. Adv Drug Deliv Rev. 2008;60(12):1347–70.

    Article  CAS  Google Scholar 

  • Matarrese M, Bedeschi P, Scardaoni R, Sudati F, Savi A, Pepe A, et al. Automated production of copper radioisotopes and preparation of high specific activity [64Cu]Cu-ATSM for PET studies. Appl Radiat Isot. 2010;68(1):5–13. https://doi.org/10.1016/j.apradiso.2009.08.010.

    Article  CAS  Google Scholar 

  • McCarthy DW, Bass LA, Cutler PD, Shefer RE, Klinkowstein RE, Herrero P, et al. High purity production and potential applications of copper-60 and copper-61. Nucl Med Biol. 1999;26(4):351–8.

    CAS  Google Scholar 

  • McCarthy DW, Shefer RE, Klinkowstien RE, Bass LA, Margeneau WH, Cutler CS, et al. Efficient production of high specific activity 64Cu using a biomedical cyclotron. Nucl Med Biol. 1997;24(1):35–43.

    CAS  Google Scholar 

  • Mikolajczak R, van der Meulen NP, Lapi SE. Radiometals for imaging and theranostics, current production, and future perspectives. J Label Compd Radiopharm. 2019;62(10):615–34. https://doi.org/10.1002/jlcr.3770.

    Article  CAS  Google Scholar 

  • Obata A, Kasamatsu S, Lewis JS, Furukawa T, Takamatsu S, Toyohara J, et al. Basic characterization of 64Cu-ATSM as a radiotherapy agent. Nucl Med Biol. 2005;32(1):21–8.

    CAS  Google Scholar 

  • Ohya T, Minegishi K, Suzuki H, Nagatsu K, Fukada M, Hanyu M, et al. Development of a remote purification apparatus with disposable evaporator for the routine production of high-quality 64Cu for clinical use. Appl Radiat Isot. 2019;146(October 2018):127–32. https://doi.org/10.1016/j.apradiso.2019.01.024.

    Article  CAS  Google Scholar 

  • Ohya T, Nagatsu K, Suzuki H, Fukada M, Minegishi K, Hanyu M, et al. Efficient preparation of high-quality 64Cu for routine use. Nucl Med Biol. 2016;43(11):685–91.

    CAS  Google Scholar 

  • Piel H, Qain SM, Stücklin G. Excitation functions of (p,xn)-reactions on natNi and highly enriched 62Ni: possibility of production of medically important radioisotope 62Cu at a small cyclotron. Radiochim Acta. 1992;57(1):1–6.

    CAS  Google Scholar 

  • Price EW, Orvig C. Matching chelators to radiometals for radiopharmaceuticals. Chem Soc Rev. 2014;43(1):260–90.

    CAS  Google Scholar 

  • Singh A, Kulkarni HR, Baum RP. Imaging of prostate cancer using 64Cu-labeled prostate-specific membrane antigen ligand. PET Clin. 2017;12(2):193–203. https://doi.org/10.1016/j.cpet.2016.12.001.

    Article  Google Scholar 

  • Song IH, Lee TS, Park YS, Lee JS, Lee BC, Moon BS, et al. Immuno-PET imaging and radioimmunotherapy of 64Cu−/177Lu-labeled anti-EGFR antibody in esophageal squamous cell carcinoma model. J Nucl Med. 2016;57(7):1105–11.

    CAS  Google Scholar 

  • Strangis R, Lepera CG. Production of 61Cu by deuteron irradiation of natural Ni. Proc 18th Int Conf Cyclotrons their Appl 2007. Cyclotrons. 2007;2007:246–7.

    Google Scholar 

  • Thieme S, Walther M, Pietzsch HJ, Henniger J, Preusche S, Mäding P, et al. Module-assisted preparation of 64Cu with high specific activity. Appl Radiat Isot. 2012;70(4):602–8.

    CAS  Google Scholar 

  • Tsai WTK, Wu AM. Aligning physics and physiology: engineering antibodies for radionuclide delivery. J Label Compd Radiopharm. 2018;61(9):693–714.

    CAS  Google Scholar 

  • Vivier D, Sharma SK, Zeglis BM. Understanding the in vivo fate of radioimmunoconjugates for nuclear imaging. J Label Compd Radiopharm. 2018;61(9):672–92.

    CAS  Google Scholar 

  • Wadas T, Wong E, Weisman G, Anderson C. Copper chelation chemistry and its role in copper radiopharmaceuticals. Curr Pharm Des. 2006;13(1):3–16.

    Google Scholar 

  • Wadas TJ, Wong EH, Weisman GR, Anderson CJ. Coordinating radiometals of copper, gallium, indium, yttrium, and zirconium for PET and SPECT imaging of disease. Chem Rev. 2010;110(5):2858–902.

    CAS  Google Scholar 

  • Wallhaus TR, Lacy J, Whang J, Green MA, Nickles RJ, Stone CK. Human biodistribution and dosimetry of the PET perfusion agent copper- 62-PTSM. J Nucl Med. 1998;39(11):1958–64.

    CAS  Google Scholar 

  • Williams HA, Robinson S, Julyan P, Zweit J, Hastings D. A comparison of PET imaging characteristics of various copper radioisotopes. Eur J Nucl Med Mol Imaging. 2005;32(12):1473–80.

    Article  Google Scholar 

  • Woo SK, Jang SJ, Seo MJ, Park JH, Kim BS, Kim EJ, et al. Development of 64Cu-NOTA-trastuzumab for HER2 targeting: a radiopharmaceutical with improved pharmacokinetics for human studies. J Nucl Med. 2019;60(1):26–33.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We gratefully acknowledge the support of Swiss Nuclides, and in particular, Leila Jaafar, for constructive feedback during the development process and providing valuable insight regarding drafting of the manuscript. Provision of samples and constructive input from Steffen Happel, Triskem is also acknowledged.

Funding

Grant funding not applicable.

Author information

Authors and Affiliations

Authors

Contributions

JP/KG implemented and optimized the 61Cu purification scheme; CK performed the target preparation, including high activity 61Cu irradiations and extensive and detailed quality control testing and analysis; JWE evaluated the results/provided critical feedback to optimization. All four authors were significant contributors to writing and editing of the manuscript and have read and approved the final manuscript.

Corresponding author

Correspondence to Katherine Gagnon.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

JP and KG are employees of GE Healthcare. CJK and JWE declare no conflict of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Svedjehed, J., Kutyreff, C.J., Engle, J.W. et al. Automated, cassette-based isolation and formulation of high-purity [61Cu]CuCl2 from solid Ni targets. EJNMMI radiopharm. chem. 5, 21 (2020). https://doi.org/10.1186/s41181-020-00108-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s41181-020-00108-7

Keywords