Next Article in Journal
Assessment Tools for the Admission of Older Adults to Inpatient Rehabilitation: A Scoping Review
Previous Article in Journal
Survival of Myocardial Infarction Patients with Diabetes Mellitus at the Invasive Era (Results from the Városmajor Myocardial Infarction Registry)
Previous Article in Special Issue
Turbo Gradient and Spin-Echo BLADE-DWI for Extraocular Muscles in Thyroid-Associated Ophthalmopathy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microspheres as a Carrier System for Therapeutic Embolization Procedures: Achievements and Advances

by
Mick. M. Welling
1,
Nikolas Duszenko
1,2,
Maarten P. van Meerbeek
1,
Tom J. M. Molenaar
1,3,
Tessa Buckle
1,
Fijs W. B. van Leeuwen
1 and
Daphne D. D. Rietbergen
1,4,*
1
Interventional Molecular Imaging Laboratory, Department of Radiology, Leiden University Medical Center, 2333 ZA Leiden, The Netherlands
2
Departments of Parasitology and Infectious Diseases, Leiden University Medical Center, 2333 ZA Leiden, The Netherlands
3
Radiochemistry Facility, Department of Radiology, Leiden University Medical Center, 2333 ZA Leiden, The Netherlands
4
Section of Nuclear Medicine, Department of Radiology, Leiden University Medical Center, 2333 ZA Leiden, The Netherlands
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2023, 12(3), 918; https://doi.org/10.3390/jcm12030918
Submission received: 30 December 2022 / Revised: 17 January 2023 / Accepted: 20 January 2023 / Published: 24 January 2023
(This article belongs to the Special Issue 10th Anniversary of JCM - Nuclear Medicine & Radiology)

Abstract

:
The targeted delivery of anti-cancer drugs and isotopes is one of the most pursued goals in anti-cancer therapy. One of the prime examples of such an application is the intra-arterial injection of microspheres containing cytostatic drugs or radioisotopes during hepatic embolization procedures. Therapy based on the application of microspheres revolves around vascular occlusion, complemented with local therapy in the form of trans-arterial chemoembolization (TACE) or radioembolization (TARE). The broadest implementation of these embolization strategies currently lies within the treatment of untreatable hepatocellular cancer (HCC) and metastatic colorectal cancer. This review aims to describe the state-of-the-art TACE and TARE technologies investigated in the clinical setting for HCC and addresses current trials and new developments. In addition, chemical properties and advancements in microsphere carrier systems are evaluated, and possible improvements in embolization therapy based on the modification of and functionalization with therapeutical loads are explored.

1. Introduction

The targeted delivery of anti-cancer drugs and isotopes is the most pursued goal in anti-cancer therapy. However, a major disadvantage of the systemic application of these drugs is their poor selectivity for cancer cells and their general distribution to non-cancerous tissues, causing toxic side effects to otherwise healthy tissue [1,2]. To this end, novel therapies are constantly being discovered and applied for anti-cancer interventions in pursuit of avoiding toxic side effects. Theoretically, the involvement of non-cancerous tissues can also be reduced via disease-targeted therapeutic delivery. Herein, disease-targeting can be guided by either receptor targeting or via selective administration [3]. Unique for the hepatic system is that cancerous diseases such as HCC and mCRC alter the vascularization of the liver. Normal liver tissue will receive oxygen from the portal system; HCC and mCRC receive oxygen from the artery and are therefore suitable for trans-arterial therapy [4]. Over the past two decades, microspheres have proven themselves effective trans-arterial drug delivery systems [5]. Such microspheres have broad applications in both life science research and medicine [6], e.g., contrast agents [7], tissue fillers [8], and drug delivery vehicles [9]. Stable or biodegradable microspheres are generally utilized to direct drugs to organs by taking advantage of vascular physical restraints [10,11,12,13,14]. Therapeutic microspheres are characteristically small particles or beads with a well-defined size distribution of 50–750 mm, consisting of either glass, synthetic polymers, or proteins [15]. By selectively blocking the blood supply of the targeted tissue via so-called trans-arterial embolization (TAE), the deprived tissue will be starved of nutrients, ultimately resulting in a therapeutic benefit. When lodged in the end vasculature that surrounds the tumor, microspheres can also release encapsulated drugs (trans-arterial chemoembolization (TACE)) or irradiate surrounding tissue via embedded radioisotopes (trans-arterial radioembolization (TARE)).
Being a medicinal product, microspheres are considered medical devices rather than drugs. From a practical perspective, this means that such agents have to comply with different regulations compared to more standard therapeutics. Independent of this classification, the various clinical requirements ultimately drive therapeutical applications, of which the most important ones are stated in Table 1. Caine et al. extensively reviewed more detailed information on the microspheres used for TAE [16].
Therapeutical microspheres have been widely implemented for hepatocellular cancer (HCC) and hepatic metastasized disease of different kinds of solid cancers, including colorectal, lung, and breast cancer that metastasize to the liver [5,17]. Non-treatable advanced neoplastic diseases and the development of hepatic metastasis have poor prognoses. Only 10–20% of these patients are suitable candidates for radical resection, as surgical excision cannot be applied in grade 3–5 staged HCC [18]. A substantial portion of inoperable patients who present metastatic liver tumors needs alternative treatment strategies or therapy for bridging surgery [19,20]. Alternative therapies encompass minimally invasive techniques, including percutaneous ablative treatments (radiofrequency ablation, microwave ablation) and trans-catheter intra-arterial therapies [21,22].
With a focus on HCC, this review provides an overview of the properties and applications of various microspheres and microsphere carrier systems’ chemical properties and advancements. In addition, the latest developments in embolization strategies and alternative technologies are discussed (Scheme 1).

2. Trans-Arterial Embolization for Vascular Occlusion

Trans-arterial embolization (TAE) is a technique wherein inert microspheres are used to block the blood supply around the tumor. The use of large microspheres, e.g., Embosphere® (⌀ 300–750 mm, Merit Medical Systems, South Jordan, UT, USA) in tumor-affected liver lobe(s) allows blockage of the blood supply towards the tumor (Figure 1B), ultimately resulting in reduced tumor growth [10,11,12,13,14]. An overview of the different microspheres currently used for hepatic trans-arterial embolization is provided in Table 2. The clinically applied glass and synthetic spheres are non-degradable and remain in the vasculature for life. Recently, in a pre-clinical phase, biodegradable spheres were developed, which allow local occlusion of the vasculature, and after degradation they allow a follow-up injection of embolization treatment. The clinical unfavorable short- and long-term outcomes of patients with large HCCs (≥50 mm) were revealed compared to those with small HCCs (<50 mm). Detailed analyses revealed that the average rates of change in tumor size and shrinkage after TAE were 48.6 ± 35.6 mm and 30.7 ± 17.0%, respectively [23]. Additional details on the chemical properties and advancements of microsphere carrier systems are evaluated in Section 4.
Instead of focusing on major vascular occlusion, chemo- or radio-embolic targeting is an alternative treatment option with microspheres that have the potential to manage hepatic tumors effectively [18].

3. Therapeutic Loads Employed during Microsphere-Trans-Arterial Embolization Therapy

3.1. Trans-Arterial Chemoembolization

Trans-arterial chemoembolization (TACE) uses embolization to deliver chemotherapy locally, thus limiting systemic exposure (Figure 1) [30,31]. One prime example are acrylic co-polymer microspheres (Hepasphere™, ⌀ 50–100 µm, Merit Medical Systems, Inc., South Jordan, UT, USA) that can absorb cytostatic drugs such as doxorubicin, irinotecan, epirubicin, mitomycin, cisplatin, and oxaliplatin (Table 3).
After contact with either an ionized environment, such as 0.9% NaCl and blood, or nonionic contrast media, acrylic co-polymer microspheres expand to 200–400 µm and slowly release their cytostatic payload [50]. The advantage of TACE is that it maximizes the concentration of chemotherapeutic agents within the tumor for up to seven days while keeping a minimal concentration in the systemic circulation. This approach reduces systemic side effects and the toxicity of cytostatic drugs compared to systemic chemotherapy. Furthermore, the malignancy’s arterial supply is occluded like in TAE, thus limiting nutrient availability to the tumor [51,52]. In HCC patients [5], this chemoembolization strategy has proven beneficial for patients’ survival, increasing the survival time by up to 12 months [53] and reducing the symptoms related to chemotherapy [28].

3.2. Trans-Arterial Radioembolization (TARE)

Trans-arterial radioembolization (TARE) uses radioisotopes embedded in microspheres to locally irradiate tissue after the vascular occlusion of blood vessels surrounding the tumor [54]. Herein, microspheres carrying b-emitting radioisotopes enable a more pinpointed delivery of radiation to liver tumors than other radiotherapy techniques. Although these options extend patient survival, most remain palliative [10,53]. With 166Ho-microspheres and 3.8 GBq/kg liver tissue in a Phase II study including 38 patients, the target lesions showed complete response or stabilized disease for 27 patients (73%), with a median survival of 15 months [55]. Given these results, a more extensive, randomized Phase III study appears to be required. Although 90Y, 188Re, and 166Ho in microspheres effectively reduce tumor size and patients’ survival, data from large Phase III trials are warranted to prove their benefits compared to other treatment modalities. In addition, the cost-effectiveness between the various radioisotopes and types of microspheres has to be determined. Besides the application of TARE in HCC, patients with intrahepatic cholangiocarcinoma (CC), which is a rare but very aggressive neoplasia with limited therapeutic options, and patients with to liver metastasized colorectal (mCRC) and neuroendocrine disease also benefit from therapy using 90Y-loaded glass or resin microspheres, with a response of more prolonged overall survival of at least 6 months [56,57,58,59]. Various radioembolization microspheres carrying radioisotopes to be delivered in the tumor-bearing hepatic segments of patients are summarized in Table 4. This table is specified in non-degradable and degradable microspheres. As we noted for TAE and TACE, the non-degradable microspheres allow only one treatment session due to the permanent occlusion of the vasculature. In pre-clinical settings, biodegradable microspheres have been evaluated for their effectiveness in TARE, as described in Section 4.
An essential issue in radioembolization studies is preventing shunting to normal tissues, such as the lungs. Shunting displaces a fraction of the administered particles towards the microvasculature of other tissues, mainly the lung, instead of the liver, leading to ineffective dose distribution and irreversible severe adverse effects such as radiation pneumonitis [88,94]. To avoid the shunting of microspheres, radioembolization is performed in a theranostic setting. In this setting, a catheter is (selectively) placed to deliver radiolabeled macro-aggregated albumin (99mTc-MAA; ⌀ 10–40 mm) to the affected tissues. Via a scout scan using single-positron emission computed tomography (SPECT) imaging, the localization and distribution of the 99mTc-MAA are visualized, which is an approach that helps assess the degree of shunting and, at the same time, facilitates dosimetry measurements. When this has been done, therapeutic loads of β-emitting glass or resin microparticles (⌀ 15–25 mm) [95] containing, e.g., 90Y (SIR-Spheres®, Sirtex Medical Inc. Woburn, MA, USA; Therasphere, Boston Scientific, Marlborough, MA, USA) or 166Ho (QuiremSpheres, Quirem Medical Deventer, The Netherlands) are injected via a catheter positioned in the same way [96]. Given the overlap in size and retention properties between 99mTc-MAA and microparticles, a 99mTc-MAA scout scan has been deemed a sufficient standard requirement in the clinical guidelines to predict the accurate delivery of therapeutic microspheres. Despite these guidelines, a mismatch between the scout and therapeutic is inevitable, i.e., given the time span that separates these two procedures. As such, the delivery of microspheres can still lead to adverse side effects and suboptimal dose delivery in about 30% of cases [97,98,99], a complication which highlights the need for innovative solutions that help refine the correlation between the scout scan and therapeutic delivery. In this respect, the physical properties of the used radioisotope, 199Ho as a b/g/paramagnetic element [100] and 90Y as a PET/SPECT imaging agent, post-TARE imaging facilitates mapping of the dose delivery. Where the therapy is insufficient, adjuvant therapy can be considered [101,102]. Furthermore, microspheres containing these isotopes can also serve as a scout scan. Another drawback is the delay of 2 weeks between the execution of the scout scan and the therapeutic intervention due to the need for dosimetry [103,104,105] and the production/delivery time of the β-emitting microspheres [96].

3.3. TACE vs. TARE

Instead of focusing on major vascular occlusion with TAE, in this section, we focus on comparing chemo- or radio-embolic targeting that has the potential to facilitate vascular occlusion and realize chemo- or radio-embolic treatment of hepatic tumors [18].
Four TARE studies determined the overall survival median at 9–11 months [50]. Based on these findings, TARE was not recommended as a first-line therapy for patients with non-resectable colorectal liver metastasis. For HCC, however, the overall survival in a study with unresectable HCC patients was 19.9 months in the 90Y-resin TARE group, which was an improvement compared to the 14 months of survival in a matching TACE group [106]. Recently, the efficacy of TARE combined with TACE was determined in 19 patients with bi-lobar HCC, and no procedure-related major clinical complications were observed, and the mean overall survival yielded a promising 27.3 months compared to untreated patients [107]. Differences between studies comparing TACE and TARE indicate that the outcome in the benefits of treatment may be related to the type of carcinoma, an observation that needs additional research. TARE also proved superior in safety regarding post-embolization syndrome, hospitalization days, and outpatient-based therapy [108,109]. TARE was a safe alternative treatment to TACE [110], especially as using a scout scan helps prevent complications related to shunting with TACE [111]. Applying the scout scan also helps personalize the dosing, a concept that could extend to TACE. Compared to TACE, TARE had a longer time-to-progression, greater ability to downsize tumors, and less post-embolization syndrome [112]. For that reason, it could be an alternative to ablation, surgical resection, or portal vein embolization [113]. On the other hand, TACE is the trans-arterial treatment of choice for patients with marginal hepatic reserve (i.e., hyperbilirubinemia, ascites) or candidates for transplantation [114].

4. Chemical Properties and Advancements of Microsphere Carrier Systems

This section evaluates the chemical properties and advancements of radioactive functionalized microsphere carrier systems. A summary of these properties and modifications of microspheres is depicted in Figure 2 and Figure 3.

4.1. Non-Degradable Microspheres

Several examples are available for the use of non-degradable microspheres in embolization procedures. Such particles are generally chemically stable and can be sterilized and monodispersed with a very tight particle-size distribution. Therapy with non-degradable microspheres is only suitable for a single embolization intervention and remains in the vasculature indefinitely. Examples of such non-degradable particles are discussed below.
  • Chemically inert glass—microspheres are non-porous, do not induce immunological effects, and are FDA-approved for application in humans for embolization therapy. Glass microspheres have a primary yttria–alumina–silica system (YAS), and a ternary YAS composition (40% Y2O3, 40% SiO2, and 20% Al2O3) [66]. These microspheres, on average, have a diameter of 50–150 mm. Glass is used in clinical TAE and TARE therapies [115,116]. Radioisotopes can be embedded during preparation into the glass via thermal neutron irradiation in a nuclear reactor (Figure 2). In this process, irradiation of stable Y—through an 89Y (n, γ)90Y reaction—produces 90Y [43,117]. The amount can be fine-tuned between 0.5 and 11 GBq per treatment according to need [118]. These glass microspheres cannot be used for TACE as elution of embedded drugs and surface modifications are impossible.
  • Ion-exchange resin-based microspheres SIR-Spheres® are porous, have a lower density/weight than glass, and are regularly used for TAE and TARE [54]. These polymers do not contain any groups amenable to covalent conjugation. Various resins were investigated for TARE only after including stable Y, Ho, or Sm, preparing microspheres during the synthesis process (Figure 2). As for glass microspheres, neutron-activated formulations in a research reactor after irradiation of stable Y, Ho, or SM isotopes through (n, γ) reaction procedures yielded 90Y, 166Ho, or 153Sm [43,54]. Bio-Rex 70 (Bio-Rad Inc. Veenendaal, The Netherlands) proved to have the best properties in stability, loading capacity, and sterilization [119]. Next to high-energy beta-radiation, 166Ho also emits gamma-radiation, which allows for imaging by gamma scintigraphy, thus helping determine treatment dose. 153Sm and 166Ho have a theragnostic advantage as they emit both therapeutic beta and diagnostic gamma radiations, allowing both imaging and therapy in one. In combination with 153Sm, this resin was pursued as an alternative microsphere in light of production time, stability, and costs [120,121].
  • The discovery of the macroporous chelating ion exchanger G-Gel (Merck, Darmstad, Germany), consisting of poly(glycidyl methacrylate-co-ethylene dimethacrylate), helps provide a new class of 177Lu and 131I carriers [122]. Methacrylate is formed in beads that support high radionuclide loading due to macroporous structure and mechanically stable sphere-shaped particles of 20–40 μm. This concept is being implemented in G-Gel [43]. The material facilitates the conjugation of functional moieties, such as chelates and dyes, without negatively impacting the overall properties [70]. In this respect, DOTA and Quinoline-8-ol have been used as metal-binding ligands because they readily form stable complexes with nearly all therapeutically or diagnostically used metal ion radionuclides such as 90Y, 188Re, 166Ho, and 177Lu [54]. This stable complexation makes G-Gel less useful for drug release as with TACE.
  • The interest in using cellulose for TARE comes from its nontoxicity, biocompatibility, biodegradability, and amenable chemistry for functionalization with, e.g., chelating groups. Polyhydroxyamic acid polyacrylamide (PHA) has been chosen for its capacity to form complexes with a wide range of metallic radionuclides [71,123]. In one study on the efficacy of a PHA loaded with 177Lu [123], PHA microspheres were synthesized starting from polyacrylamide. Incorporating isotopes such as 177Lu seems straightforward; modification of the polymers to incorporate dyes/chelates/adamantane seems impossible due to the complicated chemistry. Thus, PHA-functionalized microspheres were not applied for TACE, as is the case for resin microspheres. Subsequently, experimental variables such as reaction pH, amount of PHA microspheres, carrier 177Lu content, and incubation time were optimized for maximum uptake of 177Lu on PHA microspheres (median particle size to be 54 μm, which is still suitable for TARE, but relatively small for an effective TAE). Under optimized conditions, >99% loading of 177Lu on PHA microspheres with high stability could be achieved. 177Lu-PHA microspheres exhibited excellent in vitro stability in sodium phosphate solutions, saline, and serum for up to 5 days at 37 °C. In animal studies, 93% of 177Lu-PHA microspheres were retained in the liver at 96 h post-injection without significant leakage to other organs. Although the latter is encouraging, this set-up has not yet been evaluated in patients for HCC [71].

4.2. Bio-Degradable Particles

Biodegradable microspheres have the potential to provide an alternative to stable microspheres, and they can potentially allow sequential microsphere administrations. Repeated injections are an advantage, considering that a single treatment of HCC using TACE or TARE may not be sufficient for a successful remission. Moreover, they can be armed with bifunctional payloads, e.g., radioisotopes and elution of anti-tumor drugs. The following section summarizes various microspheres’ characteristics and applications (Table 3, Figure 2) and mentions the FDA clearance. For TAE only, biodegradable microspheres have not been evaluated.
  • Poly-DL-lactic-co-glycolic acid (PLGA) particles can be formed to the size of microspheres by employing emulsion or microemulsion polymerization, interfacial polymerization, and precipitation polymerization, and a monomer as a starting point [124]. Thereafter, they can be modified into biodegradable carriers for the controlled delivery of drugs and isotopes. PLGA has one reactive COOH group per polymer chain. Functionalization of PLGA microspheres should be possible, although the influence on the hydrogel formation in combination with PEG is unclear. PLGA is widely used for TACE [44,45,46], although the process of drug release is complex [125]. In general, drug release occurs mainly via diffusion through pores, osmotic pumping, degradation, or erosion. More recently, modifications for TARE have been initiated [126]. An application of PLGA in TARE is in Radiogel® (Vivos Inc., Richland, WA, USA), registered as a medical device under the FDA. InjecTable 90Y-Radiogel® comprises an insoluble 90Y-phosphate (YPO4) radiation source mixed within an injectable, thermally reversible, temperature-sensitive polymer solution that includes polylactide, polyglycolide, and polylactic-co-glycolic acid co-polymers, all embedded in a microsphere [126]. This hydrogel is a liquid at temperatures below body temperature but begins to gel and harden upon injection as the temperature increases to normal body temperature, thereby locking the particles in place. RadioGel® is drained within tumor extracellular spaces after injection when it warms to body temperature and has a short half-life, delivering more than 90% of its therapeutic radiation within 10 days. Over time, natural breakdown products of RadioGel® include lactic acid and glycolic acid (also known as non-toxic natural byproducts of the Krebs cycle), and the remaining radioactivity is excreted via urine [127].
  • Hydroxyapatite (Ca10(PO4)6(OH)2) is a natural mineral constituent of bone matrix and, hence, is biocompatible. Hydroxyapatite particles can be easily synthesized in the desired particle-size range for embolization, and the abundant PO43− moieties can coordinate 166Ho. Earlier studies have investigated hydroxyapatite lanthanum oxide composites [128] and the effect of tissue engineering strategies on bone regeneration [129]. The synthesis of these particles requires heating in an oven at 1250 °C [130]. Hydroxyapatite particles were uniformly spherical and large (50 µm), with a high specific surface area, uniform mesopores, and a doxorubicin loading capacity of 460.8 µg mg−1. In vivo, hydroxyapatite particles could be smoothly delivered through an arterial catheter to achieve chemoembolization. Doxorubicin-loaded hydroxyapatite particles effectively inhibited liver cancer cell growth in a rabbit liver tumor model, demonstrating the efficacy of TACE [131]. Pre-clinical studies explored the possibility of using hydroxyapatite particles with a 20–60 μm size range for vascular occlusion [78,132]. After 6 weeks of therapy, the biodegradation of the hydroxyapatite particles was realized by metabolizing Ca2+ and PO43− ions [83,84].
  • 166Ho-poly L-lactic acid (PLLA) microspheres have been developed as a possible alternative to TARE with glass- or resin-containing 90Y, as PLLA is biocompatible with the human body and its degradation reaction is mainly due to hydrolysis to lactic acid [75,77]. The chelated and stable form of Ho, Sm, or Y is added as an acetylacetonate compound and mixed with L-Lactic acid (LLA) polymer during microsphere polymerization [77,133]. When the particles are formed and isolated, they are irradiated with neutrons, which form the radioactive 166Ho, 153Sm, or 90Y [75].
  • Macro-aggregate albumin particles from HSA-B20 (Rotop Pharmaka, Dresden, Germany), Vasculosis® (Global Medical Solutions, Auckland, New Zealand), MAA (DRAXIMAGE®, Kirkland, QC, Canada), Pulmocis® Curium (London, UK)) are prepared after heating albumin and can be labeled directly with 99mTc, a recipe routinely used for scout scans in TARE set-up [99,103,134]. Alternatively, different therapeutic approaches have been investigated for TACE [49] and TARE [86], but, to date, only 188Re-labeled human serum albumin (188Re-HSA) microspheres have made their way to the clinic [85]. One advantage of HSA is that it is an approved carrier molecule, with 99mTc-HSA (Vasculosis®, Nanocoll® (GE Healthcare Ltd., Milan, Italy), Nanoalbumon® (Radiopharmacy Laboratoy Ltd, Budaörs, Hungary), Magnevist® (Bayer Inc., Toronto, ON, Canada) routinely used in nuclear medicine centers, indicated for blood pool imaging, angiocardiography, and ventriculography [135]. Pre-clinical [87] and clinical feasibility studies with 188Re-MAA have been published [65,85]. Both clinical studies demonstrated high product stability, low urinary excretion, good tolerance, and acceptable toxicity. Larger cohorts are necessary to conclude the usefulness of this device, which seems to be the ideal match with 99mTc-MAA. More recently, 90Y-DTPA-HSA microspheres were successfully evaluated in rats [68]. In pre-clinical settings, MAA was functionalized with adamantane to allow a pre-targeting set-up in the liver of mice. With this, targeting and imaging with a radiolabeled CD-PIBMA-Cy5 polymer yielded uptake in the vasculature with the functionalized MAA in the liver of mice based on host–guest chemistry [136,137]. Recently, a similar pre-clinical setup was carried out using click chemistry based on the interaction between azide-functionalized MAA and a radiolabeled DBCO-carrying moiety [138]. Given these pre-clinical findings, forming HSA microspheres seems feasible and can be carried out at low costs in a GLP facility [85].
  • Chitosan, a polymer of 2-deoxy-2-amino-d-glucose obtained from the exoskeletons of crustaceans such as crabs and shrimps, transforms from a liquid to a gel state above pH 6 [99]. Chitosan is also regularly used in nanoparticle vaccines [139]. The feasibility of chitosan for TAE was assessed in the renal arteries of a rabbit model [140]. The renal arteries were still completely occluded after 8 weeks, and no inflammatory reaction was observed. Several strategies are available to modify chitosan, which can be used to couple additional moieties of interest [141,142]; for example, pre-activation of the COOH-bearing label (e.g., adamantane) with DIC or EDC in (acidic) water can be followed by the addition of chitosan PyBOP base to subsequently conjugate the COOH-bearing label. Kim et al. studied doxorubicin-loaded chitosan microcapsules in TACE in rabbits [143]. In a recent study, biodegradable chitosan was used to deliver and retain 166Ho at the tumor site [31,79]. Chitosan was complexed with 166Ho after mixing 166HoCl3 or 166Ho(NO3)3 at pH 3 for 30 min [144,145]. The holmium/chitosan complex (Millican, Dong Wha Pharmaceutical Co., Seoul, Korea) was effective in treating small HCCs in a novel study based on 40 patients with single HCC < 3 cm in size with satisfactory response rates and survival rates of 87.2%, 71.8%, and 65.3% at 1, 2, and 3 years, respectively [79,89,90].
  • Starch-based microparticles (SBMP) were proposed as a unique system for the pre-therapeutic step (scout scan) after 188Re or 68Ga radiolabeling and TARE after direct radiolabeling with 188Re using SnCl2 reduction with gluconate or with 68GaCl3 and sodium acetate [93]. SBMP appeared to be a promising theranostic agent for the internal radiation therapy of hepatocellular carcinoma. SBMP was first developed for lung perfusion scintigraphy and formulated as a ready-to-use 99mTc radiolabeling kit [91,92]. After selecting suitable size particles via mechanical filtration, an aldehyde is formed, followed by the attachment of a diamine-linker. This chemistry should be possible with an amine, for example, Ahx [91,92,93]. The in vivo stability of the compounds is of primary importance, especially considering the therapeutic one, i.e., the SBMP radiolabeled with 188Re, and further investigations in pre-clinical models are warranted.

5. Advancements and Future Perspectives for Therapeutic Loads

TACE has been employed as drug-eluting microspheres, where embolic microspheres loaded with positively charged drugs release them locally to the tumor site via ion exchange, thus reducing systemic drug exposure [146,147]. The drug of choice for loading into microspheres is doxorubicin, a cytotoxic agent that interferes with DNA tumor growth [148]. Anti-angiogenic strategies are being considered in combination with drug-eluting particles to combat resistance [149]. It has been shown in pre-clinical and clinical settings that hypoxic conditions can lead to doxorubicin resistance in HCC cells and angiogenic upregulation of the formation of new blood vessels feeding the tumor cells [150,151].
Regarding TARE, most of the innovations lie in further integrating scout scans and actual therapy delivery. One strategy that has been suggested is the use of low-dose therapeutic spheres for scout scans [152]. While ensuring that the particles behave identically, this strategy potentially puts the patient at risk. Alternatives have also been sought after. Recent pre-clinical studies report on the possibility of microsphere surface modifications in combination with host–guest [136,137,153,154] and “click” chemistry [155] strategies. In these pre-targeting approaches, after functionalization with either adamantane (guest-vector) or azide moieties, microspheres of aggregated albumin (MAA) create a platform for the introduction of therapeutic moieties. This feature further integrates the scout scan with the therapeutic delivery. Uniquely, this strategy allows for the exploratory use of cells as “bio-microspheres” [156].
Recently, developments have been made in the direction of targeted α-therapy (TAT), using α-emitting isotopes such as 225Ac [157], 211At, and 223Ra [158]. Generally, α-emitting isotopes are effective in tissue penetration up to less than 0.1 mm, whereas with β-emitting isotopes, the radiation range in tissue varies between 2 and 12 mm, depending on their β-energies. Due to their shorter range, higher linear energy transfer, and presence of therapeutic daughter nuclides, α-particles are ideal for local tumor treatment [159]. For embolization therapy with interest in the platform of TARE, advances can be made using a-emitting radioisotopes instead of b-emitting radioisotopes. A point to consider is the radiochemistry to prepare a-emitting isotopes inside microspheres.
The BCLC system (Barcelona Clinic Liver Cancer) is recommended as a staging and treatment algorithm for HCC. For patients who are not a candidate for curative treatments, locoregional therapies, as discussed in this review, are recommended. On the other hand, the BCLC system does not recommend systemic molecular therapy for early-stage HCC. The choice of treatment depends on the availability of the discussed treatment options, besides patients’ costs and toxicity. Where TARE and TACE can be considered, this will partly depend on the patient’s characteristics, tumor size and number, local availability and expertise, and, of course, the decisions made during the tumor board meeting. The meta-analysis of Chow et al. showed a similar overall survival (OS) for the use of RFA, TACE, and TARE. Further research is in a more homogeneous group [160].

6. Conclusions

A range of materials has been used for TAE, TACE, and TARE. Hereby, the chemical composition of the microspheres guided the presence of a therapeutic load, indicating that specific healthcare needs require specific chemical designs.

Author Contributions

Conceptualization, M.M.W., T.B., F.W.B.v.L. and D.D.D.R.; methodology, M.M.W. and F.W.B.v.L.; investigation and resources, M.M.W., N.D., M.P.v.M., T.J.M.M., F.W.B.v.L. and D.D.D.R.; writing—original draft preparation, M.M.W., T.B., D.D.D.R. and F.W.B.v.L.; writing—review and editing, M.M.W., N.D., M.P.v.M., T.J.M.M., F.W.B.v.L., T.B. and D.D.D.R.; supervision, M.M.W. and F.W.B.v.L.; funding acquisition, F.W.B.v.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Netherlands Organization for Scientific Research TTW-VICI grant (Grant No. TTW 16141).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pérez-López, A.; Martín-Sabroso, C.; Gómez-Lázaro, L.; Torres-Suárez, A.I.; Aparicio-Blanco, J. Embolization therapy with microspheres for the treatment of liver cancer: State-of-the-art of clinical translation. Acta Biomater. 2022, 149, 1–15. [Google Scholar] [CrossRef]
  2. Marchal, S.; El Hor, A.; Millard, M.; Gillon, V.; Bezdetnaya, L. Anticancer drug delivery: An update on clinically applied nanotherapeutics. Drugs 2015, 75, 1601–1611. [Google Scholar] [CrossRef] [PubMed]
  3. Lorscheider, M.; Gaudin, A.; Nakhlé, J.; Veiman, K.L.; Richard, J.; Chassaing, C. Challenges and opportunities in the delivery of cancer therapeutics: Update on recent progress. Ther. Deliv. 2021, 12, 55–76. [Google Scholar] [CrossRef]
  4. Gritzapis, A.D.; Mahaira, L.G.; Perez, S.A.; Cacoullos, N.T.; Papamichail, M.; Baxevanis, C.N. Vaccination with Human HER-2/neu (435-443) CTL Peptide Induces Effective Antitumor Immunity against HER-2/neu-Expressing Tumor Cells In vivo. Cancer Res. 2006, 66, 5452–5460. [Google Scholar] [CrossRef] [Green Version]
  5. Rajput, M.; Agrawal, P. Microspheres in cancer therapy. Indian J. Cancer 2010, 47, 458–468. [Google Scholar] [CrossRef] [PubMed]
  6. Sinha, V.R.; Goyal, V.; Bhinge, J.R.; Mittal, B.R.; Trehan, A. Diagnostic microspheres: An overview. Crit. Rev. Drug Carr. Syst. 2003, 20, 431–460. [Google Scholar] [CrossRef] [PubMed]
  7. Klibanov, A.L. Targeted delivery of gas-filled microspheres, contrast agents for ultrasound imaging. Adv. Drug Deliv. Rev. 1999, 37, 139–157. [Google Scholar] [CrossRef] [PubMed]
  8. Laeschke, K. Biocompatibility of microparticles into soft tissue fillers. Semin. Cutan Med. Surg. 2004, 23, 214–217. [Google Scholar] [CrossRef]
  9. Varde, N.K.; Pack, D.W. Microspheres for controlled release drug delivery. Expert Opin. Biol. Ther. 2004, 4, 35–51. [Google Scholar] [CrossRef] [PubMed]
  10. Niu, H.; Du, T.; Xiao, Q.; Hu, X.; Li, D.; Wang, C.; Gao, W.; Xing, T.; Xu, X. Application of embolization microspheres in interventional therapy of malignant non-hypervascular tumor of liver. Oncotarget 2017, 8, 55593–55599. [Google Scholar] [CrossRef] [Green Version]
  11. Osuga, K.; Nakajima, Y.; Sone, M.; Arai, Y.; Nambu, Y.; Hori, S. Transarterial embolization of hypervascular tumors using trisacryl gelatin microspheres (Embosphere): A prospective multicenter clinical trial in Japan. Jpn. J. Radiol. 2016, 34, 366–375. [Google Scholar] [CrossRef]
  12. Vaidya, S.; Tozer, K.R.; Chen, J. An overview of embolic agents. Semin. Interv. Radiol. 2008, 25, 204–215. [Google Scholar] [CrossRef] [Green Version]
  13. Granberg, D.; Eriksson, L.G.; Welin, S.; Kindmark, H.; Janson, E.T.; Skogseid, B.; Oberg, K.; Eriksson, B.; Nyman, R. Liver embolization with trisacryl gelatin microspheres (embosphere) in patients with neuroendocrine tumors. Acta Radiol. 2007, 48, 180–185. [Google Scholar] [CrossRef]
  14. Hiraki, T.; Koizumi, J.; Arai, Y.; Sakurai, Y.; Kumada, H.; Nambu, Y.; Hori, S. Transcatheter arterial embolization of hypervascular tumors with HepaSphere: Prospective multicenter open label clinical trial of microspheres in Japan. Jpn. J. Radiol. 2015, 33, 479–486. [Google Scholar] [CrossRef] [PubMed]
  15. Kettenbach, J.; Stadler, A.; Katzler, I.V.; Schernthaner, R.; Blum, M.; Lammer, J.; Rand, T. Drug-loaded microspheres for the treatment of liver cancer: Review of current results. Cardiovasc. Interv. Radiol. 2008, 31, 468–476. [Google Scholar] [CrossRef] [PubMed]
  16. Caine, M.; Carugo, D.; Zhang, X.; Hill, M.; Dreher, M.R.; Lewis, A.L. Review of the development of methods for characterization of microspheres for use in embolotherapy: Translating bench to cathlab. Adv. Healthc. Mater. 2017, 6, 1601291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Burrill, J.; Hafeli, U.; Liu, D. Advances in radioembolization—Embolics and isotopes. J. Nucl. Med. Radiat. Ther. 2011, 2, 1000107. [Google Scholar] [CrossRef]
  18. Duran, R.; Chapiro, J.; Schernthaner, R.E.; Geschwind, J.F. Systematic review of catheter-based intra-arterial therapies in hepatocellular carcinoma: State of the art and future directions. Br. J. Radiol. 2015, 88, 20140564. [Google Scholar] [CrossRef] [Green Version]
  19. Janevska, D.; Chaloska-Ivanova, V.; Janevski, V. Hepatocellular carcinoma: Risk factors, diagnosis and treatment. Open Access Maced. J. Med. Sci. 2015, 3, 732–736. [Google Scholar] [CrossRef] [Green Version]
  20. Coletta, M.; Nicolini, D.; Benedetti Cacciaguerra, A.; Mazzocato, S.; Rossi, R.; Vivarelli, M. Bridging patients with hepatocellular cancer waiting for liver transplant: All the patients are the same? Transl. Gastroenterol. Hepatol. 2017, 2, 78. [Google Scholar] [CrossRef] [Green Version]
  21. Li, D.; Kang, J.; Golas, B.J.; Yeung, V.W.; Madoff, D.C. Minimally invasive local therapies for liver cancer. Cancer Biol. Med. 2014, 11, 217–236. [Google Scholar] [CrossRef] [PubMed]
  22. Freedman, J.; Nilsson, H.; Jonas, E. New horizons in ablation therapy for hepatocellular carcinoma. Hepat. Oncol. 2015, 2, 349–358. [Google Scholar] [CrossRef] [PubMed]
  23. Saito, R.; Amemiya, H.; Hosomura, N.; Kawaida, H.; Shoda, K.; Furuya, S.; Akaike, H.; Kawaguchi, Y.; Inoue, S.; Kono, H.; et al. Intended preoperative trans-arterial embolization for large hepatocellular carcinoma: A retrospective cohort study. World J. Surg. Oncol. 2022, 20, 90. [Google Scholar] [CrossRef]
  24. Shimohira, M.; Sato, Y.; Yasumoto, T.; Kodama, Y.; Masada, T.; Inaba, Y.; Yamakado, K. Arterial embolization using microspheres for hypervascular liver metastases refractory to standard treatments: A multicenter prospective clinical trial. Cardiovasc. Interv. Radiol. 2021, 44, 392–400. [Google Scholar] [CrossRef] [PubMed]
  25. Laurent, A. Microspheres and Nonspherical Particles for Embolization. Tech. Vasc. Interv. Radiol. 2007, 10, 248–256. [Google Scholar] [CrossRef] [PubMed]
  26. Lee, K.H.; Liapi, E.A.; Cornell, C.; Reb, P.; Buijs, M.; Vossen, J.A.; Ventura, V.P.; Geschwind, J.F. Doxorubicin-loaded QuadraSphere microspheres: Plasma pharmacokinetics and intratumoral drug concentration in an animal model of liver cancer. Cardiovasc. Interv. Radiol. 2010, 33, 576–582. [Google Scholar] [CrossRef] [Green Version]
  27. Hecq, J.D.; Lewis, A.L.; Vanbeckbergen, D.; Athanosopoulos, A.; Galanti, L.; Jamart, J.; Czuczman, P.; Chung, T. Doxorubicin-loaded drug-eluting beads (DC Bead®) for use in transarterial chemoembolization: A stability assessment. J. Oncol. Pharm. Pract. 2013, 19, 65–74. [Google Scholar] [CrossRef] [PubMed]
  28. Lewis, A.L.; Gonzalez, M.V.; Lloyd, A.W.; Hall, B.; Tang, Y.; Willis, S.L.; Leppard, S.W.; Wolfenden, L.C.; Palmer, R.R.; Stratford, P.W.; et al. DC bead: In vitro characterization of a drug-delivery device for transarterial chemoembolization. J. Vasc. Interv. Radiol. 2006, 17, 335–342. [Google Scholar] [CrossRef]
  29. Wang, Y.; Molin, D.G.M.; Sevrin, C.; Grandfils, C.; van den Akker, N.M.S.; Gagliardi, M.; Knetsch, M.L.; Delhaas, T.; Koole, L.H. In vitro and in vivo evaluation of drug-eluting microspheres designed for transarterial chemoembolization therapy. Int. J. Pharm. 2016, 503, 150–162. [Google Scholar] [CrossRef]
  30. Liu, Y.S.; Lin, C.Y.; Chuang, M.T.; Lin, C.Y.; Tsai, Y.S.; Wang, C.K.; Ou, M.C. Five-year outcome of conventional and drug-eluting transcatheter arterial chemoembolization in patients with hepatocellular carcinoma. BMC Gastroenterol. 2018, 18, 124. [Google Scholar] [CrossRef]
  31. Ni, J.Y.; Xu, L.F.; Wang, W.D.; Sun, H.L.; Chen, Y.T. Conventional transarterial chemoembolization vs microsphere embolization in hepatocellular carcinoma: A meta-analysis. World J. Gastroenterol. 2014, 20, 17206–17217. [Google Scholar] [CrossRef]
  32. De Luis, E.; Bilbao, J.I.; de Ciercoles, J.A.; Martinez-Cuesta, A.; de Martino Rodriguez, A.; Lozano, M.D. In vivo evaluation of a new embolic spherical particle (HepaSphere) in a kidney animal model. Cardiovasc. Interv. Radiol. 2008, 31, 367–376. [Google Scholar] [CrossRef] [PubMed]
  33. Kennoki, N.; Hori, S.; Hori, A.; Takeo, Y.; Oshiro, H. Transcatheter arterial chemoembolization with spherical embolic material for locally advanced breast cancer: First report of HepaSphere treatment for primary breast cancer. BJR Case Rep. 2016, 2, 20150417. [Google Scholar] [CrossRef] [PubMed]
  34. Sottani, C.; Leoni, E.; Porro, B.; Montagna, B.; Amatu, A.; Sottotetti, F.; Quaretti, P.; Poggi, G.; Minoia, C. Validation of an LC-MS/MS method for the determination of epirubicin in human serum of patients undergoing drug eluting microsphere-transarterial chemoembolization (DEM-TACE). J. Chromatogr. B 2009, 877, 3543–3548. [Google Scholar] [CrossRef] [PubMed]
  35. Poggi, G.; Quaretti, P.; Minoia, C.; Bernardo, G.; Bonora, M.R.; Gaggeri, R.; Ronchi, A.; Saluzzo, C.M.; Azzaretti, A.; Rodolico, G.; et al. Transhepatic arterial chemoembolization with oxaliplatin-eluting microspheres (OEM-TACE) for unresectable hepatic tumors. Anticancer Res. 2008, 28, 3835–3842. [Google Scholar] [PubMed]
  36. Poursaid, A.; Jensen, M.M.; Huo, E.; Ghandehari, H. Polymeric materials for embolic and chemoembolic applications. J. Control. Release 2016, 240, 414–433. [Google Scholar] [CrossRef] [Green Version]
  37. Duan, X.H.; Li, H.; Ren, J.Z.; Han, X.W.; Chen, P.F.; Li, F.Y.; Huang, G.H.; Ju, S.G. Hepatic Arterial Chemoembolization With Arsenic Trioxide Eluting CalliSpheres Microspheres Versus Lipiodol Emulsion: Pharmacokinetics And Intratumoral Concentration In A Rabbit Liver Tumor Model. Cancer Manag. Res. 2019, 11, 9979–9988. [Google Scholar] [CrossRef] [Green Version]
  38. Amrein, M.L.; Soong, C.; Liang, N. Upregulated Membrane Expression of a Conserved Voltage—Gated Sodium Channel, Nav1.4a, and Electrical Organ Discharge in Electric Mouse, P. pikachu. PLoS Biol. 2013, 11, 1001501. [Google Scholar] [CrossRef]
  39. Wu, B.; Zhou, J.; Ling, G.; Zhu, D.; Long, Q. CalliSpheres drug-eluting beads versus lipiodol transarterial chemoembolization in the treatment of hepatocellular carcinoma: A short-term efficacy and safety study. World J. Surg. Oncol. 2018, 16, 69. [Google Scholar] [CrossRef]
  40. Ma, Y.; Zhao, C.; Zhao, H.; Li, H.; Chen, C.; Xiang, H.; Zheng, C.; Ma, C.; Luo, C.; Qiu, H.; et al. Comparison of treatment efficacy and safety between drug-eluting bead transarterial chemoembolization with CalliSpheres® microspheres and conventional transarterial chemoembolization as first-line treatment in hepatocellular carcinoma patients. Am. J. Transl. Res. 2019, 11, 7456–7470. [Google Scholar]
  41. Zhang, S.; Huang, C.; Li, Z.; Yang, Y.; Bao, T.; Chen, H.; Zou, Y.; Song, L. Comparison of pharmacokinetics and drug release in tissues after transarterial chemoembolization with doxorubicin using diverse lipiodol emulsions and CalliSpheres Beads in rabbit livers. Drug Deliv. 2017, 24, 1011–1017. [Google Scholar] [CrossRef] [Green Version]
  42. Pelage, J.P.; Fohlen, A.; Mitry, E.; Lagrange, C.; Beauchet, A.; Rougier, P. Chemoembolization of neuroendocrine liver metastases using streptozocin and tris-acryl microspheres: Embozar (EMBOsphere + ZAnosaR) study. Cardiovasc. Interv. Radiol. 2017, 40, 394–400. [Google Scholar] [CrossRef] [Green Version]
  43. Beaujeux, R.; Laurent, A.; Wassef, M.; Casasco, A.; Gobin, Y.P.; Aymard, A.; Rufenacht, D.; Merland, J.J. Trisacryl gelatin microspheres for therapeutic embolization, II: Preliminary clinical evaluation in tumors and arteriovenous malformations. Am. J. Neuroradiol. 1996, 17, 541–548. [Google Scholar] [PubMed]
  44. Qian, J.; Truebenbach, J.; Graepler, F.; Pereira, P.; Huppert, P.; Eul, T.; Wiemann, G.; Claussen, C. Application of poly-lactide-co-glycolide-microspheres in the transarterial chemoembolization in an animal model of hepatocellular carcinoma. World J. Gastroenterol. 2003, 9, 94. [Google Scholar] [CrossRef] [PubMed]
  45. Makadia, H.K.; Siegel, S.J. Poly lactic-co-glycolic acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef]
  46. Blasi, P. Poly(lactic acid)/poly(lactic-co-glycolic acid)-based microparticles: An overview. J. Pharm. Investig. 2019, 49, 337–346. [Google Scholar] [CrossRef] [Green Version]
  47. Fuchs, K.; Duran, R.; Denys, A.; Bize, P.E.; Borchard, G.; Jordan, O. Drug-eluting embolic microspheres for local drug delivery-State of the art. J. Control. Release 2017, 262, 127–138. [Google Scholar] [CrossRef] [PubMed]
  48. Fujimoto, S.; Miyazaki, M.; Endoh, F.; Takahashi, O.; Okui, K.; Morimoto, Y. Biodegradable mitomycin C microspheres given intra-arterially for inoperable hepatic cancer. With particular reference to a comparison with continuous infusion of mitomycin C and 5-fluorouracil. Cancer 1985, 56, 2404–2410. [Google Scholar] [CrossRef]
  49. Karimi, M.; Bahrami, S.; Ravari, S.B.; Zangabad, P.S.; Mirshekari, H.; Bozorgomid, M.; Shahreza, S.; Sori, M.; Hamblin, M.R. Albumin nanostructures as advanced drug delivery systems. Expert Opin. Drug Deliv. 2016, 13, 1609–1623. [Google Scholar] [CrossRef] [Green Version]
  50. Tsitskari, M.; Filippiadis, D.; Kostantos, C.; Palialexis, K.; Zavridis, P.; Kelekis, N.; Brountzos, E. The role of interventional oncology in the treatment of colorectal cancer liver metastases. Ann. Gastroenterol. 2019, 32, 147–155. [Google Scholar] [CrossRef]
  51. Forner, A.; Llovet, J.M.; Bruix, J. Hepatocellular carcinoma. Lancet 2012, 379, 1245–1255. [Google Scholar] [CrossRef] [PubMed]
  52. Kloeckner, R.; Weinmann, A.; Prinz, F.; Pinto dos Santos, D.; Ruckes, C.; Dueber, C.; Pitton, M.B. Conventional transarterial chemoembolization versus drug-eluting bead transarterial chemoembolization for the treatment of hepatocellular carcinoma. BMC Cancer 2015, 15, 465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sangro, B.; Carpanese, L.; Cianni, R.; Golfieri, R.; Gasparini, D.; Ezziddin, S.; Paprottka, P.M.; Fiore, F.; van Buskirk, M.; Ignacio Bilbao, J.; et al. Survival after Yttrium-90 resin microsphere radioembolization of hepatocellular carcinoma across Barcelona clinic liver cancer stages: A European evaluation. Hepatology 2011, 54, 868–878. [Google Scholar] [CrossRef] [PubMed]
  54. Bouvry, C.; Palard, X.; Edeline, J.; Ardisson, V.; Loyer, P.; Garin, E.; Lepareur, N. Transarterial radioembolization (TARE) agents beyond (90)Y-microspheres. Biomed. Res. Int. 2018, 2018, 1435302. [Google Scholar] [CrossRef] [Green Version]
  55. Prince, J.F.; van den Bosch, M.A.A.J.; Nijsen, J.F.W.; Smits, M.L.J.; van den Hoven, A.F.; Nikolakopoulos, S.; Wessels, F.J.; Bruijnen, R.C.G.; Braat, M.N.G.J.A.; Zonnenberg, B.A.; et al. Efficacy of Radioembolization with166Ho-Microspheres in Salvage Patients with Liver Metastases: A Phase 2 Study. J. Nucl. Med. 2018, 59, 582–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Filippi, L.; Schillaci, O.; Cianni, R.; Bagni, O. Yttrium-90 resin microspheres and their use in the treatment of intrahepatic cholangiocarcinoma. Future Oncol. 2018, 14, 809–818. [Google Scholar] [CrossRef]
  57. Robinson, T.J.; Du, L.; Matsuoka, L.; Sze, D.Y.; Kennedy, A.S.; Gandhi, R.T.; Kouri, B.E.; Collins, Z.S.; Kokabi, N.; Grilli, C.J.; et al. Survival and toxicities after Yttrium-90 transarterial radioembolization of Cholangiocarcinoma in the RESiN registry. J. Vasc. Interv. Radiol. 2022, in press. [CrossRef]
  58. Bargellini, I.; Bozzi, E.; Lorenzoni, G.; Boni, G.; Bianchi, F.; Traino, C.A.; Masi, G.; Cioni, R.; Crocetti, L. Role of Transhepatic Arterial Radioembolization in Metastatic Colorectal Cancer. Cardiovasc. Interv. Radiol. 2022, 45, 1579–1589. [Google Scholar] [CrossRef]
  59. Ingenerf, M.K.; Karim, H.; Fink, N.; Ilhan, H.; Ricke, J.; Treitl, K.M.; Schmid-Tannwald, C. Apparent diffusion coefficients (ADC) in response assessment of transarterial radioembolization (TARE) for liver metastases of neuroendocrine tumors (NET): A feasibility study. Acta Radiol. 2022, 63, 877–888. [Google Scholar] [CrossRef]
  60. Raoul, J.L.; Guyader, D.; Bretagne, J.F.; Heautot, J.F.; Duvauferrier, R.; Bourguet, P.; Bekhechi, D.; Deugnier, Y.M.; Gosselin, M. Prospective randomized trial of chemoembolization versus intra-arterial injection of 131I-labeled-iodized oil in the treatment of hepatocellular carcinoma. Hepatology 1997, 26, 1156–1161. [Google Scholar] [CrossRef]
  61. Pirayesh, E.; Amoui, M.; Akhlaghpoor, S.; Tolooee, S.; Khorrami, M.; Poorbeigi, H.; Sheibani, S.; Assadi, M. Technical considerations of phosphorous-32 Bremsstrahlung SPECT imaging after radioembolization of hepatic tumors: A clinical assessment with a review of imaging parameters. Radiol. Res. Pract. 2014, 2014, 407158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Raoul, J.L.; Duvauferrier, R.; Bourguet, P.; Bretagne, J.F.; Coornaert, S.; Darnault, P.; Deugnier, Y.; Herry, J.Y.; Gastard, J. Lipiodolized angiography in hepatocellular carcinomas. Contribution of iodine-131-labelled lipiodol. J. Radiol. 1986, 67, 797–801. [Google Scholar] [PubMed]
  63. Hafeli, U.O.; Casillas, S.; Dietz, D.W.; Pauer, G.J.; Rybicki, L.A.; Conzone, S.D.; Day, D.E. Hepatic tumor radioembolization in a rat model using radioactive rhenium (186Re/188Re) glass microspheres. Int. J. Radiat. Oncol. Biol. Phys. 1999, 44, 189–199. [Google Scholar] [CrossRef] [PubMed]
  64. Lepareur, N.; Lacoeuille, F.; Bouvry, C.; Hindre, F.; Garcion, E.; Cherel, M.; Noiret, N.; Garin, E.; Knapp, F.F.R., Jr. Rhenium-188 labeled radiopharmaceuticals: Current clinical applications in oncology and promising perspectives. Front. Med. 2019, 6, 00132. [Google Scholar] [CrossRef] [Green Version]
  65. Liepe, K.; Brogsitter, C.; Leonhard, J.; Wunderlich, G.; Hliscs, R.; Pinkert, J.; Folprecht, G.; Kotzerke, J. Feasibility of high activity rhenium-188-microsphere in hepatic radioembolization. Jpn. J. Clin. Oncol. 2007, 37, 942–950. [Google Scholar] [CrossRef] [Green Version]
  66. Poorbaygi, H.; Aghamiri, S.M.R.; Sheibani, S.; Kamali-asl, A.; Mohagheghpoor, E. Production of glass microspheres comprising 90Y and 177Lu for treating of hepatic tumors with SPECT imaging capabilities. Appl. Radiat. Isot. 2011, 69, 1407–1414. [Google Scholar] [CrossRef]
  67. Akram, A.R.; Avlonitis, N.; Scholefield, E.; Vendrell, M.; McDonald, N.; Aslam, T.; Craven, T.H.; Gray, C.; Collie, D.S.; Fisher, A.J.; et al. Enhanced avidity from a multivalent fluorescent antimicrobial peptide enables pathogen detection in a human lung model. Sci. Rep. 2019, 9, 8422. [Google Scholar] [CrossRef] [Green Version]
  68. Vukadinović, V.; Janković, D.; Radović, M.; Milanović, Z.; Mirković, M.; Stanković, D.; Vranješ-Đurić, S. Optimization of the radiolabelling method for improved in vitro and in vivo stability of 90Y-albumin microspheres. Appl. Radiat. Isot. 2020, 156, 108984. [Google Scholar] [CrossRef]
  69. Hashikin, N.A.A.; Yeong, C.-H.; Abdullah, B.J.J.; Ng, K.-H.; Chung, L.-Y.; Dahalan, R.; Perkins, A.C. Neutron activated samarium-153 microparticles for transarterial radioembolization of liver tumour with post-procedure imaging capabilities. PLoS ONE 2015, 10, 0138106. [Google Scholar] [CrossRef] [Green Version]
  70. Hruby, M.; Skodova, M.; Mackova, H.; Skopal, J.; Tomes, M.; Kropacek, M.; Zimova, J.; Kucka, J. Lutetium-177 and iodine-131 loaded chelating polymer microparticles intended for radioembolization of liver malignancies. React. Funct. Polym. 2011, 71, 1155–1159. [Google Scholar] [CrossRef]
  71. Pandey, U.; Subramanian, S.; Shaikh, S.; Gamre, N.; Kumar, S.; Dash, A. Synthesis and preliminary biological evaluation of 177Lu-labeled polyhydroxamic acid microparticles toward therapy of hepatocellular carcinoma. Cancer Biother. Radiopharm. 2019, 34, 306–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Saxena, S.K.; Kumar, Y.; Shaikh, S.H.; Pandey, U.; Kumar, S.A.; Dash, A. Preparation of radioactive skin patches using polyhydroxamic acid-grafted cellulose films toward applications in treatment of superficial tumors. Cancer Biother. Radiopharm. 2017, 32, 364–370. [Google Scholar] [CrossRef] [PubMed]
  73. Wong, Y.H.; Tan, H.Y.; Kasbollah, A.; Abdullah, B.J.J.; Yeong, C.H. Preparation and in vitro evaluation of neutron-activated, theranostic samarium-153-labeled microspheres for transarterial radioembolization of hepatocellular carcinoma and liver metastasis. Pharmaceutics 2019, 11, 596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Fisher, D.; Fidel, J.; Maitz, C. Direct interstitial treatment of solid tumors using an injectable yttrium-90-polymer composite. Cancer Biother. Radiopharm. 2020, 35, 2947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Mumper, R.J.; Ryo, U.Y.; Jay, M. Neutron-activated holmium-166-poly (L-lactic acid) microspheres: A potential agent for the internal radiation therapy of hepatic tumors. J. Nucl. Med. 1991, 32, 2139–2143. [Google Scholar]
  76. Vente, M.A.; Nijsen, J.F.; de Roos, R.; van Steenbergen, M.J.; Kaaijk, C.N.; Koster-Ammerlaan, M.J.; de Leege, P.F.; Hennink, W.E.; van Het Schip, A.D.; Krijger, G.C. Neutron activation of holmium poly(L-lactic acid) microspheres for hepatic arterial radio-embolization: A validation study. Biomed. Microdevices 2009, 11, 763–772. [Google Scholar] [CrossRef] [Green Version]
  77. Mumper, R.J.; Jay, M. Biodegradable radiotherapeutic polyester microspheres: Optimization and in-vitro/in-vivo evaluation. J. Control. Release 1992, 18, 193–203. [Google Scholar] [CrossRef]
  78. Das, T.; Chakraborty, S.; Sarma, H.D.; Venkatesh, M.; Banerjee, S. 166Ho-labeled hydroxyapatite particles: A possible agent for liver cancer therapy. Cancer Biother. Radiopharm. 2009, 24, 7–14. [Google Scholar] [CrossRef]
  79. Kim, J.K.; Han, K.H.; Lee, J.T.; Paik, Y.H.; Ahn, S.H.; Lee, J.D.; Lee, K.S.; Chon, C.Y.; Moon, Y.M. Long-term clinical outcome of phase IIb clinical trial of percutaneous injection with holmium-166/chitosan complex (Milican) for the treatment of small hepatocellular carcinoma. Clin. Cancer Res. 2006, 12, 543–548. [Google Scholar] [CrossRef] [Green Version]
  80. De la Vega, J.C.; Esquinas, P.L.; Rodríguez-Rodríguez, C.; Bokharaei, M.; Moskalev, I.; Liu, D.; Saatchi, K.; Häfeli, U.O. Radioembolization of hepatocellular carcinoma with built-in dosimetry: First in vivo results with uniformly-sized, biodegradable microspheres labeled with188re. Theranostics 2019, 9, 868–883. [Google Scholar] [CrossRef]
  81. Jamre, M.; Shamsaei, M.; Erfani, M.; Sadjadi, S.; Maragheh, M.G. Preparation and evaluation of 188Re sulfide colloidal nanoparticles loaded biodegradable poly (L-lactic acid) microspheres for radioembolization therapy. J. Label. Compd. Radiopharm. 2018, 61, 586–594. [Google Scholar] [CrossRef] [PubMed]
  82. Hrubý, M.; Hradil, J.; Beneš, M.J. Interactions of phenols with silver(I), copper(II) and iron(III) complexes of chelating methacrylate-based polymeric sorbent containing quinolin-8-ol groups. React. Funct. Polym. 2004, 59, 105–118. [Google Scholar] [CrossRef]
  83. Chinol, M.; Vallabhajosula, S.; Goldsmith, S.J.; Klein, M.J.; Deutsch, K.F.; Chinen, L.K.; Brodack, J.W.; Deutsch, E.A.; Watson, B.A.; Tofe, A.J.; et al. Chemistry and biological behavior of samarium-153 and rhenium-186-labeled hydroxyapatite particles: Potential radiopharmaceuticals for radiation synovectomy. J. Nucl. Med. 1993, 34, 1536–1542. [Google Scholar] [PubMed]
  84. Unni, P.R.; Chaudhari, P.R.; Venkatesh, M.; Ramamoorthy, N.; Pillai, M.R. Preparation and bioevaluation of 166Ho labelled hydroxyapatite (HA) particles for radiosynovectomy. Nucl. Med. Biol. 2002, 29, 199–209. [Google Scholar] [CrossRef]
  85. Nowicki, M.L.; Cwikla, J.B.; Sankowski, A.J.; Shcherbinin, S.; Grimmes, J.; Celler, A.; Buscombe, J.R.; Bator, A.; Pech, M.; Mikolajczak, R.; et al. Initial study of radiological and clinical efficacy radioembolization using 188Re-human serum albumin (HSA) microspheres in patients with progressive, unresectable primary or secondary liver cancers. Med. Sci. Monit. 2014, 20, 1353–1362. [Google Scholar] [CrossRef] [Green Version]
  86. Bhushan, B.; Khanadeev, V.; Khlebtsov, B.; Khlebtsov, N.; Gopinath, P. Impact of albumin based approaches in nanomedicine: Imaging, targeting and drug delivery. Adv. Colloid Interface Sci. 2017, 246, 13–39. [Google Scholar] [CrossRef]
  87. Chen, L.-C.; Lee, W.-C.; Ho, C.-L.; Chang, Y.-J.; Chen, S.-J.; Chang, C.-H. Biodistribution, pharmacokinetics and efficacy of (188)re(i)-tricarbonyl-labeled human serum albumin microspheres in an orthotopic hepatoma rat model. In Vivo 2018, 32, 567–573. [Google Scholar] [CrossRef]
  88. Cremonesi, M.; Chiesa, C.; Strigari, L.; Ferrari, M.; Botta, F.; Guerriero, F.; de Cicco, C.; Bonomo, G.; Orsi, F.; Bodei, L.; et al. Radioembolization of hepatic lesions from a radiobiology and dosimetric perspective. Front. Oncol. 2014, 4, 210. [Google Scholar] [CrossRef] [Green Version]
  89. Memon, K.; Lewandowski, R.J.; Kulik, L.; Riaz, A.; Mulcahy, M.F.; Salem, R. Radioembolization for primary and metastatic liver cancer. Semin. Radiat. Oncol. 2011, 21, 294–302. [Google Scholar] [CrossRef] [Green Version]
  90. Sohn, J.H.; Choi, H.J.; Lee, J.T.; Lee, J.D.; Kim, J.H.; Moon, Y.M.; Park, K.; Park, K.B.; Kim, E.; Yoo, N.C.; et al. Phase II study of transarterial holmium-166-chitosan complex treatment in patients with a single, large hepatocellular carcinoma. Oncology 2009, 76, 1–9. [Google Scholar] [CrossRef]
  91. Lacoeuille, F.; Hindré, F.; Denizot, B.; Bouchet, F.; Legras, P.; Couturier, O.; Askiénazy, S.; Benoit, J.P.; le Jeune, J.J. New starch-based radiotracer for lung perfusion scintigraphy. Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Lacoeuille, F.; Hindré, F.; Venier-Julienne, M.C.; Sergent, M.; Bouchet, F.; Jouaneton, S.; Denizot, B.; Askienazy, S.; Benoit, J.P.; Couturier, O.F.; et al. A starch-based microparticulate system dedicated to diagnostic and therapeutic nuclear medicine applications. Biomaterials 2011, 32, 7999–8009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Verger, E.; Drion, P.; Meffre, G.; Bernard, C.; Duwez, L.; Lepareur, N.; Couturier, O.; Hindré, F.; Hustinx, R.; Lacoeuille, F.; et al. 68Ga and 188Re starch-based microparticles as theranostic tool for the hepatocellular carcinoma: Radiolabeling and preliminary in vivo rat studies. PLoS ONE 2016, 11, e0164626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Braat, A.J.A.T.; Smits, M.L.J.; Braat, M.N.G.J.A.; van den Hoven, A.F.; Prince, J.F.; de Jong, H.W.A.M.; van den Bosch, M.A.A.J.; Lam, M.G.E.H. 90Y hepatic radioembolization: An update on current practice and recent developments. J. Nucl. Med. 2015, 56, 1079–1087. [Google Scholar] [CrossRef] [Green Version]
  95. Salem, R.; Lewandowski, R.J.; Gates, V.L.; Nutting, C.W.; Murthy, R.; Rose, S.C.; Soulen, M.C.; Geschwind, J.-F.H.; Kulik, L.; Kim, Y.H.; et al. Research reporting standards for radioembolization of hepatic malignancies. J. Vasc. Interv. Radiol. 2011, 22, 265–278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Hickey, R.; Lewandowski, R.J.; Prudhomme, T.; Ehrenwald, E.; Baigorri, B.; Critchfield, J.; Kallini, J.; Gabr, A.; Gorodetski, B.; Geschwind, J.-F.; et al. 90Y radioembolization of colorectal hepatic metastases using glass microspheres: Safety and survival outcomes from a 531-patient multicenter study. J. Nucl. Med. 2016, 57, 665–671. [Google Scholar] [CrossRef] [Green Version]
  97. Uliel, L.; Royal, H.D.; Darcy, M.D.; Zuckerman, D.A.; Sharma, A.; Saad, N.E. From the angio suite to the γ-camera: Vascular mapping and 99mTc-MAA hepatic perfusion imaging before liver radioembolization—A comprehensive pictorial review. J. Nucl. Med. 2012, 53, 1736–1747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Riaz, A.; Awais, R.; Salem, R. Side effects of Yttrium-90 radioembolization. Front. Oncol. 2014, 4, 00198. [Google Scholar] [CrossRef] [Green Version]
  99. Xing, M.; Lahti, S.; Kokabi, N.; Schuster, D.M.; Camacho, J.C.; Kim, H.S. 90Y Radioembolization lung shunt fraction in primary and metastatic liver cancer as a biomarker for survival. Clin. Nucl. Med. 2016, 41, 21–27. [Google Scholar] [CrossRef]
  100. Stella, M.; Braat, A.; van Rooij, R.; de Jong, H.; Lam, M. Holmium-166 Radioembolization: Current Status and Future Prospective. Cardiovasc. Interv. Radiol. 2022, 45, 1634–1645. [Google Scholar] [CrossRef]
  101. Deidda, D.; Denis-Bacelar, A.M.; Fenwick, A.J.; Ferreira, K.M.; Heetun, W.; Hutton, B.F.; Robinson, A.P.; Scuffham, J.; Thielemans, K. Hybrid kernelised expectation maximisation for Bremsstrahlung SPECT reconstruction in SIRT with (90)Y micro-spheres. EJNMMI Phys. 2022, 9, 25. [Google Scholar] [CrossRef]
  102. D’Arienzo, M.; Pimpinella, M.; Capogni, M.; de Coste, V.; Filippi, L.; Spezi, E.; Patterson, N.; Mariotti, F.; Ferrari, P.; Chiaramida, P.; et al. Phantom validation of quantitative Y-90 PET/CT-based dosimetry in liver radioembolization. EJNMMI Res. 2017, 7, 94. [Google Scholar] [CrossRef] [Green Version]
  103. Ilhan, H.; Goritschan, A.; Paprottka, P.; Jakobs, T.F.; Fendler, W.P.; Todica, A.; Bartenstein, P.; Hacker, M.; Haug, A.R. Predictive value of 99mTc-MAA SPECT for 90Y-labeled resin microsphere distribution in radioembolization of primary and secondary hepatic tumors. J. Nucl. Med. 2015, 56, 1654–1660. [Google Scholar] [CrossRef] [Green Version]
  104. Bult, W.; Vente, M.A.; Zonnenberg, B.A.; van Het Schip, A.D.; Nijsen, J.F. Microsphere radioembolization of liver malignancies: Current developments. Q. J. Nucl. Med. Mol. Imaging 2009, 53, 325–335. [Google Scholar]
  105. Elschot, M.; Nijsen, J.F.; Lam, M.G.; Smits, M.L.; Prince, J.F.; Viergever, M.A.; van den Bosch, M.A.; Zonnenberg, B.A.; de Jong, H.W. 99mTc-MAA overestimates the absorbed dose to the lungs in radioembolization: A quantitative evaluation in patients treated with 166Ho-microspheres. Eur. J. Nucl. Med. Mol. Imaging 2014, 41, 1965–1975. [Google Scholar] [CrossRef]
  106. She, W.H.; Cheung, T.T.; Yau, T.C.; Chan, A.C.; Chok, K.S.; Chu, F.S.; Liu, R.K.; Poon, R.T.; Chan, S.C.; Fan, S.T.; et al. Survival analysis of transarterial radioembolization with yttrium-90 for hepatocellular carcinoma patients with HBV infection. Hepatobiliary Surg. Nutr. 2014, 3, 185–193. [Google Scholar] [CrossRef] [Green Version]
  107. Kwon, J.H.; Kim, G.M.; Han, K.; Won, J.Y.; Kim, M.D.; Lee, D.Y.; Lee, J.; Choi, W.; Kim, Y.S.; Kim, D.Y.; et al. Safety and efficacy of transarterial radioembolization combined with chemoembolization for bilobar hepatocellular carcinoma: A single-center retrospective study. Cardiovasc. Interv. Radiol. 2018, 41, 459–465. [Google Scholar] [CrossRef]
  108. Kim, D.Y.; Han, K.H. Transarterial chemoembolization versus transarterial radioembolization in hepatocellular carcinoma: Optimization of selecting treatment modality. Hepatol. Int. 2016, 10, 883–892. [Google Scholar] [CrossRef]
  109. Moreno-Luna, L.E.; Yang, J.D.; Sanchez, W.; Paz-Fumagalli, R.; Harnois, D.M.; Mettler, T.A.; Gansen, D.N.; de Groen, P.C.; Lazaridis, K.N.; Narayanan Menon, K.V.; et al. Efficacy and safety of transarterial radioembolization versus chemoembolization in patients with hepatocellular carcinoma. Cardiovasc. Interv. Radiol. 2013, 36, 714–723. [Google Scholar] [CrossRef] [Green Version]
  110. Lobo, L.; Yakoub, D.; Picado, O.; Ripat, C.; Pendola, F.; Sharma, R.; ElTawil, R.; Kwon, D.; Venkat, S.; Portelance, L.; et al. Unresectable hepatocellular carcinoma: Radioembolization versus chemoembolization: A systematic review and meta-analysis. Cardiovasc. Interv. Radiol. 2016, 39, 1580–1588. [Google Scholar] [CrossRef]
  111. Marcacuzco Quinto, A.; Nutu, O.A.; San Román Manso, R.; Justo Alonso, I.; Calvo Pulido, J.; Manrique Municio, A.; García-Sesma, Á.; Loinaz Segurola, C.; Martínez Caballero, J.; Jiménez Romero, L.C. Complications of transarterial chemoembolization (TACE) in the treatment of liver tumors. Cir. Esp. 2018, 96, 560–567. [Google Scholar] [CrossRef]
  112. Kim, H.C. Radioembolization for the treatment of hepatocellular carcinoma. Clin. Mol. Hepatol. 2017, 23, 109–114. [Google Scholar] [CrossRef] [Green Version]
  113. Kang, Y.J.; Lee, B.C.; Kim, J.K.; Yim, N.Y.; Kim, H.O.; Cho, S.B.; Jeong, Y.Y. Conventional versus small doxorubicin-eluting bead transcatheter arterial chemoembolization for treating barcelona clinic liver cancer stage 0/A hepatocellular carcinoma. Cardiovasc. Interv. Radiol. 2020, 43, 55–64. [Google Scholar] [CrossRef]
  114. Fidelman, N.; Kerlan, R.K., Jr. Transarterial chemoembolization and (90)Y radioembolization for hepatocellular carcinoma: Review of current applications beyond intermediate-stage disease. Am. J. Roentgenol. 2015, 205, 742–752. [Google Scholar] [CrossRef]
  115. Chauhan, N.; Bukovcan, J.; Boucher, E.; Cosgrove, D.; Edeline, J.; Hamilton, B.; Kulik, L.; Master, F.; Salem, R. Intra-arterial TheraSphere Yttrium-90 glass microspheres in the treatment of patients with unresectable hepatocellular carcinoma: Protocol for the STOP-HCC Phase 3 randomized controlled trial. JMIR Res. Protoc. 2018, 7, e11234. [Google Scholar] [CrossRef]
  116. Salem, R.; Padia, S.A.; Lam, M.; Chiesa, C.; Haste, P.; Sangro, B.; Toskich, B.; Fowers, K.; Herman, J.M.; Kappadath, S.C.; et al. Clinical, dosimetric, and reporting considerations for Y-90 glass microspheres in hepatocellular carcinoma: Updated 2022 recommendations from an international multidisciplinary working group. Eur. J. Nucl. Med. Mol. Imaging 2022, 50, 328–343. [Google Scholar] [CrossRef]
  117. Pham, T.M.; Duong, V.D.; Doan, V.-D.; Vo, V.T.; Le, V.T. Design synthesis of Y-90 glass microspheres and study of their therapeutic effects on mouse liver cancer cell line Hep3B. Chemosphere 2022, 299, 134431. [Google Scholar] [CrossRef]
  118. James, T.; Hill, J.; Fahrbach, T.; Collins, Z. Differences in radiation activity between glass and resin 90Y microspheres in treating unresectable hepatic cancer. Health Phys. 2017, 112, 300–304. [Google Scholar] [CrossRef]
  119. Schubiger, P.A.; Beer, H.F.; Geiger, L.; Rösler, H.; Zimmermann, A.; Triller, J.; Mettler, D.; Schilt, W. 90Y-resin particles—Animal experiments on pigs with regard to the introduction of superselective embolization therapy. Int. J. Radiat. Appl. Instrum. B. Nucl. Med. Biol. 1991, 18, 305–311. [Google Scholar] [CrossRef]
  120. Carmona, M.; Warchoł, J.; Lucas, A.D.; Rodriguez, J.F. Ion-Exchange equilibria of Pb2+, Ni2+, and Cr3+ ions for H+ on Amberlite IR-120 Resin. J. Chem. Engin. Data 2008, 53, 1325–1331. [Google Scholar] [CrossRef]
  121. Mondal, A.; De, S.; Maiti, S.; Sarkar, B.; Sk, A.K.; Jacob, R.; Moorthy, A.; Paira, P. Amberlite IR-120 (H) mediated “on water” synthesis of fluorescent Ruthenium(II)-arene 8-hydroxyquinoline complexes for cancer therapy and live cell imaging. J. Photochem. Photobiol. B 2018, 178, 380–394. [Google Scholar] [CrossRef]
  122. Švec, F.; Hradil, J.; Čoupek, J.; Kálal, J. Reactive polymers I. Macroporous methacrylate copolymers containing epoxy groups. Die Angew. Makromol. Chem. Appl. Macromol. Chem. Phys. 1975, 48, 135–143. [Google Scholar] [CrossRef]
  123. Saxena, S.; Gomber, C. Surmounting antimicrobial resistance in the millennium superbug: Staphylococcus aureus. Cent. Eur. J. Med. 2010, 5, 12–29. [Google Scholar] [CrossRef]
  124. Astete, C.E.; Sabliov, C.M. Synthesis and characterization of PLGA nanoparticles. J. Biomater. Sci. Polym. 2006, 17, 247–289. [Google Scholar] [CrossRef]
  125. Fredenberg, S.; Wahlgren, M.; Reslow, M.; Axelsson, A. The mechanisms of drug release in poly(lactic-co-glycolic acid)-based drug delivery systems—A review. Int. J. Pharm. 2011, 415, 34–52. [Google Scholar] [CrossRef]
  126. Fisher, D.R. Radiation safety for yttrium-90-polymer composites (radiogel™) in therapy of solid tumors. Health Phys. 2021, 120, 510–516. [Google Scholar] [CrossRef]
  127. Lambert, B.; Mertens, J.; Ravier, M.; Blanken, T.; Defreyne, L.; van Vlierberghe, H.; D’Asseler, Y.; Oltenfreiter, R. Urinary excretion of Yttrium-90 following intra-arterial microsphere treatment for liver tumours. J. Nucl. Med. 2011, 52, 1744. [Google Scholar]
  128. Bozkurt, Y.; Pazarlioglu, S.; Gokce, H.; Gurler, I.; Salman, S. Hydroxyapatite lanthanum oxide composites. Acta Phys. Pol. A 2015, 127, 1407–1409. [Google Scholar] [CrossRef]
  129. Wieszczycka, K.; Staszak, K.; Woźniak-Budych, M.J.; Jurga, S. Lanthanides and tissue engineering strategies for bone regeneration. Coord. Chem. Rev. 2019, 388, 248–267. [Google Scholar] [CrossRef]
  130. Kang, N.H.; Kim, S.J.; Song, S.H.; Choi, S.; Choi, S.Y.; Kim, Y.J. Hydroxyapatite synthesis using EDTA. J. Craniofac. Surg. 2013, 24, 1042–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Liu, L.; Liang, X.; Xu, X.; Zhang, X.; Wen, J.; Chen, K.; Su, X.; Teng, Z.; Lu, G.; Xu, J.; et al. Magnetic mesoporous embolic microspheres in transcatheter arterial chemoembolization for liver cancer. Acta Biomater. 2021, 130, 374–384. [Google Scholar] [CrossRef]
  132. Kubo, M.; Kuwayama, N.; Hirashima, Y.; Takaku, A.; Ogawa, T.; Endo, S. Hydroxyapatite ceramics as a particulate embolic material: Report of the physical properties of the hydroxyapatite particles and the animal study. Am. J. Neuroradiol. 2003, 24, 1540–1544. [Google Scholar] [PubMed]
  133. Smits, M.L.J.; Nijsen, J.F.W.; van den Bosch, M.A.A.J.; Lam, M.G.E.H.; Vente, M.A.D.; Huijbregts, J.E.; van Het Schip, A.D.; Elschot, M.; Bult, W.; de Jong, H.W.A.M.; et al. Holmium-166 radioembolization for the treatment of patients with liver metastases: Design of the phase I HEPAR trial. J. Exp. Clin. Cancer Res. 2010, 29, 70. [Google Scholar] [CrossRef] [Green Version]
  134. Grosser, O.S.; Ruf, J.; Kupitz, D.; Pethe, A.; Ulrich, G.; Genseke, P.; Mohnike, K.; Pech, M.; Richter, W.S.; Ricke, J.; et al. Pharmacokinetics of 99mtc-maa- and 99mtc-hsa-microspheres used in preradioembolization dosimetry: Influence on the liver-lung shunt. J. Nucl. Med. 2016, 57, 925–927. [Google Scholar] [CrossRef] [Green Version]
  135. Wunderlich, G.; Pinkert, J.; Andreeff, M.; Stintz, M.; Knapp, F.F., Jr.; Kropp, J.; Franke, W.G. Preparation and biodistribution of rhenium-188 labeled albumin microspheres B 20: A promising new agent for radiotherapy. Appl. Radiat. Isot. 2000, 52, 63–68. [Google Scholar] [CrossRef]
  136. Spa, S.J.; Welling, M.M.; van Oosterom, M.N.; Rietbergen, D.D.D.; Burgmans, M.C.; Verboom, W.; Huskens, J.; Buckle, T.; van Leeuwen, F.W.B. A supramolecular approach for liver radioembolization. Theranostics 2018, 8, 2377–2386. [Google Scholar] [CrossRef] [Green Version]
  137. Chung, A.S.; Niesen, M.C.; Graber, T.J.; Schwartz, A.J.; Beauchamp, C.P.; Clarke, H.D.; Spangehl, M.J. Two-stage debridement with prosthesis retention for acute periprosthetic joint infections. J. Arthroplast. 2019, 34, 1207–1213. [Google Scholar] [CrossRef]
  138. Welling, M.M.; Duszenko, N.; van Willigen, D.M.; Smits, W.K.; Buckle, T.; Roestenberg, M.; van Leeuwen, F.W.B. Cyclodextrin/Adamantane-mediated targeting of inoculated bacteria in mice. Bioconjug. Chem. 2021, 32, 607–614. [Google Scholar] [CrossRef]
  139. Yu, S.; Hao, S.; Sun, B.; Zhao, D.; Yan, X.; Jin, Z.; Zhao, K. Quaternized chitosan nanoparticles in vaccine applications. Curr. Med. Chem. 2020, 27, 4932–4944. [Google Scholar] [CrossRef]
  140. Wang, Y.; Xu, N.; Luo, Q.; Li, Y.; Sun, L.; Wang, H.; Xu, K.; Wang, B.; Zhen, Y. In vivo assessment of chitosan/β-glycerophosphate as a new liquid embolic agent. Interv. Neuroradiol. 2011, 17, 87–92. [Google Scholar] [CrossRef] [PubMed]
  141. Mourya, V.K.; Inamdar, N.N. Chitosan-modifications and applications: Opportunities galore. React. Funct. Polym. 2008, 68, 1013–1051. [Google Scholar] [CrossRef]
  142. Tan, M.L.; Choong, P.F.M.; Dass, C.R. Review: Doxorubicin delivery systems based on chitosan for cancer therapy. J. Pharm. Pharmacol. 2010, 61, 131–142. [Google Scholar] [CrossRef]
  143. Kim, J.S.; Kwak, B.K.; Shim, H.J.; Lee, Y.C.; Baik, H.W.; Lee, M.J.; Han, S.M.; Son, S.H.; Kim, Y.B.; Tokura, S.; et al. Preparation of doxorubicin-containing chitosan microspheres for transcatheter arterial chemoembolization of hepatocellular carcinoma. J. Microencapsul. 2007, 24, 408–419. [Google Scholar] [CrossRef]
  144. Park, K.B.; Kim, Y.M.; Shin, B.C.; Kim, J.R.; Ryu, J.M.; Lim, S.M. Therapeutic Application of New Holmium-166 Chitosan Complex in Malignant and Benign Diseases; International Atomic Energy Agency: Vienna, Austria, 1998; pp. 569–580. [Google Scholar]
  145. Lohar, S.; Jadhav, S.; Chakravarty, R.; Chakraborty, S.; Sarma, H.D.; Dash, A. A kit based methodology for convenient formulation of 166Ho-Chitosan complex for treatment of liver cancer. Appl. Radiat. Isot. 2020, 161, 109161. [Google Scholar] [CrossRef]
  146. Bruix, J.; Sala, M.; Llovet, J.M. Chemoembolization for hepatocellular carcinoma. Gastroenterology 2004, 127, S179–S188. [Google Scholar] [CrossRef]
  147. Lewis, A.L.; Holden, R.R. DC Bead embolic drug-eluting bead: Clinical application in the locoregional treatment of tumours. Expert Opin. Drug Deliv. 2011, 8, 153–169. [Google Scholar] [CrossRef]
  148. Gewirtz, D. A critical evaluation of the mechanisms of action proposed for the antitumor effects of the anthracycline antibiotics adriamycin and daunorubicin. Biochem. Pharm. 1999, 57, 727–741. [Google Scholar] [CrossRef]
  149. Hagan, A.; Phillips, G.J.; Macfarlane, W.M.; Lloyd, A.W.; Czuczman, P.; Lewis, A.L. Preparation and characterisation of vandetanib-eluting radiopaque beads for locoregional treatment of hepatic malignancies. Eur. J. Pharm. Sci. 2017, 101, 22–30. [Google Scholar] [CrossRef] [Green Version]
  150. Liang, B.; Zheng, C.-S.; Feng, G.-S.; Wu, H.-P.; Wang, Y.; Zhao, H.; Qian, J.; Liang, H.-M. Correlation of hypoxia-inducible factor 1α with angiogenesis in liver tumors after transcatheter arterial embolization in an animal model. Cardiovasc. Interv. Radiol. 2010, 33, 806–812. [Google Scholar] [CrossRef]
  151. Rhee, T.K.; Young, J.Y.; Larson, A.C.; Haines, G.K.; Sato, K.T.; Salem, R.; Mulcahy, M.F.; Kulik, L.M.; Paunesku, T.; Woloschak, G.E.; et al. Effect of transcatheter arterial embolization on levels of hypoxia-inducible factor-1alpha in rabbit VX2 liver tumors. J. Vasc. Interv. Radiol. 2007, 18, 639–645. [Google Scholar] [CrossRef]
  152. Gomes, J.; Gang, G.J.; Mathews, A.; Stayman, J.W. An investigation of low-dose 3D scout scans for computed tomography. Proc. Int. Soc. Opt. Eng. 2017, 10132, 677–682. [Google Scholar] [CrossRef] [Green Version]
  153. Welling, M.M.; Spa, S.J.; van Willigen, D.M.; Rietbergen, D.D.D.; Roestenberg, M.; Buckle, T.; van Leeuwen, F.W.B. In vivo stability of supramolecular host-guest complexes monitored by dual-isotope multiplexing in a pre-targeting model of experimental liver radioembolization. J. Control. Release 2019, 293, 126–134. [Google Scholar] [CrossRef]
  154. Duszenko, N.; van Willigen, D.M.; Welling, M.M.; de Korne, C.M.; van Schuijlenburg, R.; Winkel, B.M.F.; van Leeuwen, F.W.B.; Roestenberg, M. A supramolecular platform technology for bacterial cell surface modification. ACS Infect. Dis. 2020, 6, 1734–1744. [Google Scholar] [CrossRef] [PubMed]
  155. Welling, M.M.; Duszenko, N.; van Willigen, D.M.; Hensbergen, A.W.; Buckle, T.; Rietbergen, D.D.D.; Roestenberg, M.; van Leeuwen, F.W.B. Interventional nuclear medicine: “click” chemistry as an in vivo targeting strategy for imaging microspheres and bacteria. Biomater. Sci. 2021, 9, 1683–1690. [Google Scholar] [CrossRef] [PubMed]
  156. He, Q.; Zhang, J.; Liao, Y.; Alakpa, E.V.; Bunpetch, V.; Zhang, J.; Ouyang, H. Current advances in microsphere based cell culture and tissue engineering. Biotechnol. Adv. 2020, 39, 107459. [Google Scholar] [CrossRef]
  157. Guerra Liberal, F.D.C.; O’Sullivan, J.M.; McMahon, S.J.; Prise, K.M. Targeted alpha therapy: Current clinical applications. Cancer Biother. Radiopharm. 2020, 35, 404–417. [Google Scholar] [CrossRef]
  158. Vaidyanathan, G.; Zalutsky, M.R. Applications of 211At and 223Ra in targeted alpha-particle radiotherapy. Curr. Radiopharm. 2011, 4, 283–294. [Google Scholar] [CrossRef] [Green Version]
  159. Silindir-Gunay, M.; Karpuz, M.; Ozer, A.Y. Targeted alpha therapy and nanocarrier approach. Cancer Biother. Radiopharm. 2020, 35, 446–458. [Google Scholar] [CrossRef]
  160. Chow, R.; Simone, C.B., 2nd; Jairam, M.P.; Swaminath, A.; Boldt, G.; Lock, M. Radiofrequency ablation vs radiation therapy vs transarterial chemoembolization vs yttrium 90 for local treatment of liver cancer—A systematic review and network meta-analysis of survival data. Acta Oncol. 2022, 61, 484–494. [Google Scholar] [CrossRef]
Scheme 1. Theory of change for improving therapeutic embolization procedures using microspheres in HCC.
Scheme 1. Theory of change for improving therapeutic embolization procedures using microspheres in HCC.
Jcm 12 00918 sch001
Figure 1. Treatment of hepatocellular carcinoma (HCC) shows the advantages of chemo/radio-embolization drug targeting over systemic drug therapy. Systemic drug therapy (A) yields lower tumor targeting and more side effects than local delivery via trans-arterial chemo/radio embolic drug targeting (B). Yellow arrows = injection site. Created with BioRender.com.
Figure 1. Treatment of hepatocellular carcinoma (HCC) shows the advantages of chemo/radio-embolization drug targeting over systemic drug therapy. Systemic drug therapy (A) yields lower tumor targeting and more side effects than local delivery via trans-arterial chemo/radio embolic drug targeting (B). Yellow arrows = injection site. Created with BioRender.com.
Jcm 12 00918 g001
Figure 2. Schematic representations of building blocks and backbones for constructing radioactive microspheres. * = details binding unknown, insertion after mixing.
Figure 2. Schematic representations of building blocks and backbones for constructing radioactive microspheres. * = details binding unknown, insertion after mixing.
Jcm 12 00918 g002
Figure 3. Overview of microsphere properties and functionalizations. Details regarding the clinical features and pre-clinical set-ups are described in Section 2, Section 3 and Section 4. Created with BioRender.com.
Figure 3. Overview of microsphere properties and functionalizations. Details regarding the clinical features and pre-clinical set-ups are described in Section 2, Section 3 and Section 4. Created with BioRender.com.
Jcm 12 00918 g003
Table 1. Properties of microspheres: list of requirements for embolization (selected from: https://www.microspheres.us/properties-of-microspheres/ accessed on 1 November 2022).
Table 1. Properties of microspheres: list of requirements for embolization (selected from: https://www.microspheres.us/properties-of-microspheres/ accessed on 1 November 2022).
PropertyImportance
Specific gravity (particle density)Dispersion in other media or occlusion of the micro-vasculature
SizeParticle size (diameter = 50–750 mm) that allows occlusion of the microvasculature
DurabilityStrength during production, solvent resistance, sterilization, chemical stability, or biodegradation, the release of the therapeutical payload
BiocompatibilitySafety, toxicity, stability, suitable for intra-arterial delivery
PharmacologyControlled dosimetry and dosing, full control over release profile by diffusion, zero-order kinetics
Surface propertiesHydrophobic vs. hydrophilic surface, surface area, and porosity, ability to coat or functionalize the spheres
Table 2. Overview of non-functionalized microspheres for hepatic trans-arterial embolization (TAE). PC = pre-clinical use, H = Human use, FDA = FDA approved as an active implantable medical device. See Section 4 for a detailed evaluation of the microspheres.
Table 2. Overview of non-functionalized microspheres for hepatic trans-arterial embolization (TAE). PC = pre-clinical use, H = Human use, FDA = FDA approved as an active implantable medical device. See Section 4 for a detailed evaluation of the microspheres.
Microspheres CompositionProduct NameParticle Size Range (⌀ μm)PC/H (FDA Clearance)BiodegradableReferences
Tris acryl gelatin microspheres (TAGM)Embosphere® (Merit Medical Systems, South Jordan, UT, USA)100–300, 300–500H (FDA)No[24]
Polyvinyl alcohol (PVA) sodium acrylate co-polymer microspheresContour SE® (Boston Scientific, Marlborough, MA, USA),
Bead Block® (Boston Scientific, Marlborough, MA, USA)
Embozene® (Varian Medical Systems, Palo Alto, CA, USA)
Irregular:
150–250, 250–355, 355–500, 500–710
H (FDA)No[25]
Polyvinyl alcohol-based hydrogel microspheres with sulphonate groupsLC Bead® (Boston Scientific, Marlborough, MA, USA)75–150, 100–300, 300–500, 500–700H (FDA)No[26,27,28,29]
Co-polymer of PEG and diacrylamideHydropearl® (Terumo Medical Co., Somerset, NJ, USA)75–1100H (FDA)No[1]
Starch microspheresEmbocept® (Pharmacept, Berlin, Germany),
Spherex® (Magle Life Sciences, Lund, Sweden)
50PCYes[1]
Gelatin microspheresGel-Bead (Teleflex, Morrisville, NC, USA)100–300, 300–500, 500–700, 7000–1000H (FDA)Yes[1]
Collagen-coated poly-(DL-lactic acid-co-glycolic acid (PLGA) microspheresOcclusin500® (IMBiotechnologies, Edmonton, AL, Canada)150–210HYes[1]
Table 3. Drug-eluting microspheres for functionalized hepatic TACE. PC = pre-clinical use, H = human use, FDA = FDA approved as an active implantable medical device. See Section 4 for a detailed evaluation of the microspheres.
Table 3. Drug-eluting microspheres for functionalized hepatic TACE. PC = pre-clinical use, H = human use, FDA = FDA approved as an active implantable medical device. See Section 4 for a detailed evaluation of the microspheres.
Microspheres
Composition
Product NameParticle Size Range (⌀ µm)Drug LoadPC/H (FDA)BiodegradableReferences
Polyvinyl alcohol (PVA) sodium acrylate co-polymer microspheresQuadraSphere® and HepaSphere™ (Merit Medical Systems, Inc., South Jordan, UT, USA)
DC Bead® (Boston Scientific, Marlborough, MA, USA),
LC Bead®, and Bead Block® (Boston Scientific, Marlborough, MA, USA)
50–100, 100–300, 200–400Doxorubicin, irinotecan, epirubicin, oxaliplatinPC/H (FDA)No[14,26,32,33,34,35,36]
Ion-exchange microspheresCalliSpheres® Beads (Jiangsu Hengrui Medicine Co. Ltd. Jiangsu, China)100–300Doxorubicin, pirarubicin, oxaliplatinPC/HNo[37,38,39,40,41]
Tris acryl gelatin microspheres (TAGM)Embosphere (Merit Medical Systems, South Jor-dan, UT, USA),
Embozene®, and Oncozene™ (Varian Medical Systems, Palo Alto, CA, USA)
40–120, 100–300Doxorubicin and IrinotecanH (FDA)No[13,36,42,43]
Poly-lactide-co-glycolide (PLGA)Dexon®, Vicryl®, PerserisTM, Indivior (Indivior Inc. North Chesterfiled, VI, USA), Risperdal Consta®20–100Mitomycin, doxorubicin, irinotecan, sunitinib, cisplatinPC/H (FDA)Yes[44,45,46,47]
Albumin microspheresNab-paclitaxel10–220Mitomycin C, doxorubicin, paclitaxelPC/H (FDA)Yes[48,49]
Table 4. Microspheres used for trans-arterial radioembolization (TARE). PC = pre-clinical use, H = human use, FDA = FDA approved as an active implantable medical device. See Section 4 for a detailed evaluation of the microspheres.
Table 4. Microspheres used for trans-arterial radioembolization (TARE). PC = pre-clinical use, H = human use, FDA = FDA approved as an active implantable medical device. See Section 4 for a detailed evaluation of the microspheres.
Microspheres
Composition
Product NameParticle Size Range (⌀ µm)Radioisotope LoadPre-Clinical/
Human Use (FDA Clearance)
BiodegradableReferences
GlassLipiocis, TheraSphere® (Boston Scientific, Marlborough, MA, USA)50–150, 20–30, 25–3232P, 90Y, 177lu, 186Re, 188RePC/H (FDA for 90Y, 186Re, and 188Re)No[60,61,62,63,64,65]
ResinSIR-Spheres®
(Sirtex Medical Inc. Woburn, MA, USA), Amberlite IR-120 (Thermo Fisher Scientific, Landau, Germany)
20–6090Y, 153SmPC/H (FDA for 90Y)No[66,67,68,69]
Polyhydroxyamic acid polyacrylamide (PHA)Experimental54177lu, 131IPCNo[70,71,72]
Styrene divinylbenzeneAmberlite IR-120 (Thermo Fisher Scientific, Landau, Germany)20–40152SmPCNo[69,73]
Poly- DL-lactic acid-co-glycolic acid (PLGA)YPO4 crystalline particles Radiogel® (Vivos Inc., Richland, WA, USA)0.5–290YPC (FDA-approved as a medical device)Yes[74]
Poly (L-lactic acid) PLLAResomer® L104 (Merck, Darmstad, Germany)10–45, 20–40188Re/166Ho/175YbPC/CYes[75,76,77,78,79,80,81]
Poly (glycidyl methacrylate-co-ethylene dimethacrylate & Quinoline-8-olG-Gel (Merck, Darmstad, Germany)20–40131I, 177luPCNo[66,70,71,72,82]
HydroxyapatiteQuiremSpheres
(Quirem Medical, Deventer, The Netherlands)
20–60166HoPCNo[78,83,84]
AlbuminHSA-B20 (Rotop Pharmaka, Dresden, Germany)
Vasculosis® (Global Medical Solutions, Auckland, New Zealand)
MAA (DRAXIMAGE®, Kirkland, QC, Canada), Pulmocis® (Curium, London, UK)
25–35, 15–3790Y, 186Re, 188RePC/CYes[63,64,65,85,86,87]
ChitosanMillican (Dong Wha Pharmaceutical Co., Soeul, South Korea)5–20166HoPC/CYes[75,76,77,78,79,88,89,90]
Starch-based microparticles (SBMP)Experimental Kit18–42188RePCNo[91,92,93]
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

Welling, M.M.; Duszenko, N.; van Meerbeek, M.P.; Molenaar, T.J.M.; Buckle, T.; van Leeuwen, F.W.B.; Rietbergen, D.D.D. Microspheres as a Carrier System for Therapeutic Embolization Procedures: Achievements and Advances. J. Clin. Med. 2023, 12, 918. https://doi.org/10.3390/jcm12030918

AMA Style

Welling MM, Duszenko N, van Meerbeek MP, Molenaar TJM, Buckle T, van Leeuwen FWB, Rietbergen DDD. Microspheres as a Carrier System for Therapeutic Embolization Procedures: Achievements and Advances. Journal of Clinical Medicine. 2023; 12(3):918. https://doi.org/10.3390/jcm12030918

Chicago/Turabian Style

Welling, Mick. M., Nikolas Duszenko, Maarten P. van Meerbeek, Tom J. M. Molenaar, Tessa Buckle, Fijs W. B. van Leeuwen, and Daphne D. D. Rietbergen. 2023. "Microspheres as a Carrier System for Therapeutic Embolization Procedures: Achievements and Advances" Journal of Clinical Medicine 12, no. 3: 918. https://doi.org/10.3390/jcm12030918

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