Next Article in Journal
Brain Extraction Methods in Neonatal Brain MRI and Their Effects on Intracranial Volumes
Next Article in Special Issue
Comparison of Waste Biomass from Pine, Eucalyptus, and Acacia and the Biochar Elaborated Using Pyrolysis in a Simple Double Chamber Biomass Reactor
Previous Article in Journal
Personalization of Affective Models Using Classical Machine Learning: A Feasibility Study
Previous Article in Special Issue
Continuous Adsorption of Acid Wood Dyes onto an Activated Carbon Prepared from Pine Sawdust
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of High-Precision Sub-Micron CaCO3 Anticancer Drug Carriers from Coral Remains

1
Department of Resources Engineering, National Cheng Kung University, Tainan City 701401, Taiwan
2
Hierarchical Green-Energy Materials (Hi-GEM) Research Center, National Cheng Kung University, Tainan City 701401, Taiwan
*
Author to whom correspondence should be addressed.
Submission received: 23 December 2023 / Revised: 26 January 2024 / Accepted: 31 January 2024 / Published: 6 February 2024
(This article belongs to the Special Issue Bioenergy and Bioproducts from Biomass and Waste)

Abstract

:
Calcium carbonate (CaCO3) particles have attracted increasing attention as a promising material for drug delivery systems. In this study, coral remains were utilized as a raw material for a novel drug carrier. A series of pre-treatment and parameter experiments were conducted to synthesize sub-micron spherical CaCO3 particles. The CaCO3 particles exhibited uniform size distribution, with the minimum mean size being only 344 nm. The effects on the CaCO3 crystal phases and particle sizes were also discussed in this study. Drug loading experiments were also conducted to assess the feasibility of the CaCO3 drug carrier. We loaded TRITC-Dextran into CaCO3 particles for the simulation experiments. The loading capacity reached up to 9.6 wt.%, which was as high as common drug carriers such as liposomes. In this study, we aimed not only to tackle the local environmental issues caused by coral remains, but also to synthesize a suitable drug carrier for cancer therapy using the outstanding properties and low cost of CaCO3.

1. Introduction

Environmental degradation and disease are the two major challenges of this century. This study explores a novel solution to deal these issues by addressing the accumulated coral remains in Taiwan and their potential application in cancer drug development, one of the leading causes of death worldwide. Coral remains are composed of coral fragments resulting from weathering, waves, or trampling. It is a unique environmental issue in Taiwan. Despite coral remains being non-toxic and harmless, the accumulation of coral causes lots of problems along the coast in Penghu, one of the counties in Taiwan. Coral remains accumulating along the coast lead to a reduction in marine life living spaces and the obstruction of fishing routes, impacting fishermen’s livelihoods. The local government tried to dredge the coral remains and place them onto the beach. Unfortunately, the resulting odor and damage to the landscape continued to constitute serious problems. Coral remains consist of CaCO3, which is known as a safe, cheap, and clean industrial mineral. Compared to oyster shells, coral remains do not require a significant amount of time for washing and cleaning, making them an ideal material for this study.
Cancer is one of the most critical issues around the world. The International Agency for Research on Cancer (IARC) and the Union for International Cancer Control (UICC) reported nearly 19.3 million cancer cases, with the death toll reaching up to 10 million in 2020 [1]. Moreover, cancer has been the leading cause of death in Taiwan since 1982, accounting for 28.98% of all deaths in 2020, totaling 50,161 people [2]. Due to the toxicity of cancer drugs, the precision of drug delivery is a significant feature in curing cancers. Therefore, a growing amount of research is committed to loading drugs into unique containers to improve the precision of drug delivery, i.e., drug carriers [3,4]. There is currently a variety of drug carriers such as liposomes, solid lipid nanoparticles, gold nanoparticles, etc. However, they still have some critical problems that must be dealt with, e.g., bio-toxicity, aggregation, and high cost.
CaCO3 is a promising and novel material for drug carriers. It possesses several properties, including high biocompatibility, substantial drug loading capacity, pH sensitivity, low cost, and the capability to carry multiple drugs simultaneously [5,6,7,8,9]. It is widely used as an industrial mineral [10,11,12,13,14,15,16,17,18,19] and a raw material for pharmaceutical products [20,21]. CaCO3 has been utilized as a carrier to deliver proteins and genes [22]. Therefore, the public acceptability of CaCO3 is better than that of other materials. Moreover, according to the report from Vikulina et al. [23], CaCO3 costs only 0.2–0.4 USD/g, which is significantly lower than the 100 USD/g cost of liposomes.
CaCO3 is widely used in industry as a filler. It can not only lower the cost but also improve the properties of products [24,25]. CaCO3 has three major anhydrous phases: calcite, aragonite, and vaterite [26,27]. Calcite is the most stable phase among them, followed by aragonite and vaterite [28,29,30,31]. Vaterite is the least stable but the most promising phase for a drug carrier due to its high porosity, large surface, and higher solubility. Vaterite will gradually dissolve and recrystallize into calcite in water, which is an imperative mechanism for drugs to be released from vaterite (Figure 1). Owing to the instability of vaterite, the synthesizing process must be studied to control the morphology, particle size, and purity of it. To manipulate the morphologies and particle sizes of CaCO3, the addition of additives is the general approach for achieving this [32,33,34,35,36,37,38]. Parakhonskiy et al. reported that polyol can inhibit CaCO3 from growing into other phases [39]. Thus, we used ethylene glycol (EG) as an additive to maintain the vaterite phase [40,41,42]. EG can also control the particle size of vaterite, as studies have indicated that cancer cells take up molecules smaller than 600 nm selectively due to the enhanced permeability and retention effect (EPR effect) [43].

2. Materials and Methods

2.1. Materials

Coral remains from Penghu coastal accumulations and dredging operations were used as the raw material for this study. Coral remains were collected by scholars from Penghu University of Science and Technology. The dredging operations were carried out by the Penghu government. The coral remains were dredged and placed along the coast so that we could easily collect them with permission from the government.

2.2. Pretreatment and Characteristic Analysis of Coral Remains

Coral remains underwent washing, hydrogen peroxide immersing, drying, and grinding, followed by 60-mesh sieving. For elemental analysis, the sample was calcined at 1000 °C for 4 h, then dissolved in a mixture of aqua regia and hydrofluoric acid (HF) in the ratio of 9:1. Following these procedures, the solution was analyzed using Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES, Avio 200 Manx, PerkinElmer, Waltham, MA, USA).

2.3. Purification of Coral Remains and Preparation of CaCl2

Purification began with the calcination of coral remains at 1000 °C to decompose them into CaO, which was then dissolved in water to form Ca(OH)2. This step was followed by a reaction with CO2, resulting in the precipitation of CaCO3. The calcination was carried out in a tube furnace under different atmospheres (air, N2, and CO2) to study the decomposition rate of CaCO3 in varying conditions. The calcination temperatures in air and N2 were set between 700 °C and 1200 °C. In the case of CO2, temperatures ranged from 500 °C to 1000 °C due to no significant improvement observed beyond 700 °C, prompting the extension down to 500 °C. The decomposition rates of CaCO3 were calculated using the following equation (Equation (1)), where x represents the decomposition rate and 44 is the molecular weight of CO2. The purity of the resulting CaCO3 was confirmed using ICP-OES, and it was subsequently reacted with HCl to produce CaCl2, as shown in the following equations (Equations (2) and (3)):
x = ( mass   of   CaCO 3 mass   of   CaO ) / 44
CaCO 3 ( s ) + 2 HCl ( aq ) CaCl 2 ( aq ) + CO 2 ( g ) + H 2 O ( l )
CaCl 2 ( aq ) Δ CaCl 2 ( s )

2.4. Exploration of Optimal Parameters for the Synthesis of Vaterite

Solutions with different concentrations of CaCl2 and Na2CO3 were prepared and injected into EG in specific ratios. Further steps included centrifugation, washing, and drying at 60 °C for 2 h. Given the complex reaction mechanism involved in vaterite formation, a thorough investigation was conducted to gain a comprehensive understanding of its reaction mechanisms and characteristics. This involved exploring a range of parameters, such as the concentration of EG, stirring speed, total reaction time, initial salt concentration, reaction volume, and the adding order of CaCl2 and Na2CO3. The particle size and crystal phase variations of CaCO3 under these different conditions were carefully observed to identify the optimal synthesis parameters. For examining the vaterite crystal structure, Scanning Electron Microscopy (SEM, Hitachi SU-1510, Tokyo, Japan) was utilized, with the samples being coated with a palladium (Pd) layer. The particle size distribution was quantified using ImageJ Version 1.54h, with a sample size of N 150 particles per sample.

2.4.1. Adding Order of the Salts

CaCO3 was synthesized using the precipitation method, involving CaCl2 and Na2CO3. We found that the order in which these salts are added plays a critical role, a parameter not extensively discussed in the existing literature. Altering the sequence of adding CaCl2 and Na2CO3 resulted in significant variations in the crystal forms and particle sizes of the synthesized CaCO3. Therefore, the impact of the adding order of these salts was emphasized in this study.

2.4.2. Stirring Speed

This section examined the effect of various stirring speeds (400, 600, 800, 1000, and 1150 rpm) on the particle size of CaCO3.

2.4.3. EG Concentration

Ethylene glycol (EG) played a critical role in this study due to its high viscosity, which decelerates ion diffusion and inhibits the growth of CaCO3. We investigated various EG concentrations, including 0%, 25%, 50%, and 85%, to determine their impact on the synthesis of CaCO3.

2.4.4. Reaction Time

The presence of EG is known to mitigate the reaction speed. We explored how varying reaction times influence the diversity of CaCO3 formations. In this study, time intervals of 1 min, 30 min, 60 min, 180 min, and 1440 min (1 day) were examined to determine the optimal reaction duration.

2.4.5. Initial Salt Concentration

Nucleation theory suggests that the initial salt concentration can impact the nucleation rate, consequently affecting the particle size. This study varied the initial concentrations of CaCl2 and Na2CO3 to observe the resulting changes in CaCO3 and to validate the alignment of our findings with the nucleation theory.

2.4.6. Total Reaction Volume

This investigation explored the effects of different total reaction volumes, including the combined volumes of salt solutions and EG. Contrasting with the common practice in the literature, which typically synthesizes CaCO3 using drug carriers around 100 mL, we extended this range to include volumes of 4 mL, 40 mL, 120 mL, and 400 mL to investigate the formation and characteristic of CaCO3.

2.5. Drug Loading Simulation Experiment

The fluorescently labeled drug Tetramethylrhodamine Isothiocyanate–Dextran (TRITC-Dextran) was utilized for loading into CaCO3 carriers, and the loading capacity (LC) was subsequently calculated after drug loading. For LC calculation, an indirect method [44] was employed. This method involves analyzing the concentration of the drug in the supernatant collected after centrifugation the drug-loaded carrier. The LC is then estimated by deducting the drug concentration in the supernatant from the initial drug concentration. The concentration of TRITC-Dextran was maintained at 10 mg/mL in this study. We analyzed the collected supernatant using a Fluorescence Spectrophotometer (Fluorescence Spectrophotometer, Hitachi High-Tech F-7000, Hitachi High-Technologies Corporation, Tokyo, Japan). The drug-loaded carriers were examined under a confocal microscope (Confocal Microscope, Carl Zeiss LSM780, Carl Zeiss AG, Oberkochen, Germany) to assess the amount of TRITC-Dextran loaded into the CaCO3 particles. An excitation wavelength of λ = 530 nm was utilized in this study. The drug concentration in the supernatant was calculated using the following equation (Equation (4)):
F = P 0 ( 2.3 ε b c ) Q f k
where F represents the intensity of fluorescence, P 0 is the incident power, ε is the molar absorptivity, b is the path length, c is the concentration, Q f is the quantum efficiency, and k is the ratio of photons measured to photons emitted. Typically, P 0 , ε , b, Q f , and k are constant, making the concentration in the supernatant directly proportional to the fluorescence intensity F.

3. Result and Discussion

3.1. Characteristic Analysis of Coral Remains

Table 1 and Table 2 present the elemental composition of the raw and purified coral remains, as determined by ICP-OES. The raw coral remains primarily consisted of elements such as Ca, Mg, and Si, with no detectable levels of Pb and Cd. After purification, the purity of CaO significantly increased from 92.95% to 99.965%, indicating the effective removal of most impurities. This aligns with the 2023 Republic of China National Regulations Database standards for CaCO3 (§ 07014) [45], particularly for use in food and nutrition applications. These standards require a CaCO3 content of over 98%, less than 0.2% hydrochloric acid insoluble residue, and lead levels below 10 ppm. Consequently, the purified coral remains qualify as food-grade additive.

3.2. Purification of Coral Remains and Preparation of CaCl2

The calcination efficiency of coral remains under different atmospheres is shown in Figure 2a–c. The decomposition rate of CaCO3 to CaO in an N2 atmosphere showed a significant increase, achieving complete reaction in just 5 min at 800 °C, compared to the 15 min required at 900 °C in an air atmosphere. In a CO2 atmosphere, the decomposition rate experienced a slight increase and stabilized after reaching 700 °C. However, the decomposition rate plateaued at 13%, even after extending the duration to 30 min at 1000 °C. This could be attributed to the Le Châtelier’s principle, where the high partial pressure of CO2 generated during the calcination process inhibited further decomposition. Conversely, in the N2 atmosphere, the inert gas was continuously removed from the system with the simultaneous introduction of N2, facilitating a decomposition rate reaching 100% in just 5 min at 800 °C. These findings verified the theory that CO2 partial pressure significantly affects the efficiency of decomposition. Therefore, based on the results, an N2 atmosphere was determined to be the optimal condition for calcination.

3.3. Exploration of Optimal Parameters for the Synthesis of Vaterite

3.3.1. Adding Order of the Salts

This step explored the effects of the different adding orders of the salts on the synthesis of CaCO3, with operational parameters detailed in Table 3. The SEM results depicted in Figure 3 reveal that adding CaCl2 before Na2CO3 significantly affects the particle size, resulting in a more uniform distribution compared to adding Na2CO3 before CaCl2. The average particle size for CaCl2-first was found to be 423 nm, compared to 869 nm for Na2CO3-first. Consequently, CaCl2-first was chosen as the optimal parameter for subsequent reactions.

3.3.2. Stirring Speed

This step investigated the impact of different stirring speeds on the particle size of CaCO3, with the operational parameter outlined in Table 4. As shown in Figure 4a, an increase in stirring speed corresponded with a reduction in the average particle size of CaCO3. Notably, at 400 rpm, the formation of numerous calcite particles and some particle aggregation was observed (Figure 4b); as a result, the data obtained from this speed setting was deemed unsuitable in our discussion. At a speed of 1150 rpm, the particle size was reduced to an average of 364 nm, making it the optimal parameter.

3.3.3. EG Concentration

This step examined the effect of different EG concentrations on the particle size and crystal growth of CaCO3, with operational parameters detailed in Table 5. As demonstrated in Figure 5a,b, CaCO3 samples with an EG concentration of less than 50% tended to form calcite-like crystals, which was attributed to an overreaction. Conversely, increasing the EG concentration above 50% gradually suppressed the growth rate of CaCO3, favoring the formation of complete vaterite structures. This is evident in particle size and distribution observed at higher EG concentrations, which were more favorable compared to those observed at 50% or lower, as shown in Figure 5c,d. The mean particle sizes at 50% and 85% EG concentrations were 1038 nm and 475 nm, respectively. Therefore, the optimal EG concentration was determined to be 85%.

3.3.4. Reaction Time

This step explored the effect of different reaction times on the formation of vaterite crystals, with operational parameters outlined in Table 6. SEM analysis (Figure 6a–e) indicated that the optimal vaterite crystal phase and particle size distribution were achieved at a reaction time of 1 h (Figure 6c). It was observed that extending the reaction led to an increase in particle size. This increase might be attributed to the continuous formation of CaCO3 at nucleation sites, facilitated by the interaction of Ca2+ and CO32− ions. Notably, for reaction times shorter than one hour (Figure 6a,b), a clear presence of both calcite and larger vaterite particles was observed, with instances where calcite appeared embedded within vaterite. This suggests that the complete formation of vaterite might undergo a process of continuous decomposition and recrystallization, aligning with findings from Zhang et al. [46]. Therefore, a duration of 60 min was determined to be the optimal reaction time.

3.3.5. Initial Salt Concentration

Salt concentration affects the nucleation rate, which consequently influences the particle size. In this step, the effect of varying Na2CO3 concentrations on the particle size of CaCO3 was examined, with operational parameters detailed in Table 7. Results from Figure 7a–e indicate that the particle size of CaCO3 increases together with a rise in the concentration of Na2CO3, presenting an opposite trend from the classic nucleation theory [47]. Notably, the smallest particle size was 344 nm at a Na2CO3 concentration of 0.05 M, aligning with the hypothesis proposed by Trushina et al. [40].
However, this study yielded CaCO3 particles with an average size of 423 nm at 0.1 M Na2CO3, contrasting with findings from Trushina et al., who observed the smallest particles at 700 nm under similar conditions [40]. This phenomenon may be attributed to the interaction between the functional groups of EG and Ca2+ ions. The former, carrying opposite electrical charges, possibly facilitates the occupancy of Ca2+ ions at EG’s nucleation sites. Upon adding Na2CO3, the CO32− ions were prompted to react with Ca2+ ions at these nucleation sites, leading to the formation of CaCO3. Based on the experimental results, the optimal initial salt concentration was determined to be CaCl2:Na2CO3 = 1 M:0.05 M.

3.3.6. Total Reaction Volume

This step investigated the impact of total reaction volumes on the particle size of CaCO3, with operational parameters detailed in Table 8. Observations from Figure 8a–d reveal that the particle size of vaterite increased with the increment of reaction volume. Notably, the particle size of vaterite in a 400 mL solution reached 1002 nm. Moreover, the presence of calcite phase was observed in reactions conducted with 400 mL (Figure 8d). This phenomenon indicates that the total reaction volume has a noticeable impact not only on the particle size, but also on the crystal morphology of the products. Based on the experimental results, the optimal reaction volume was determined to be 4 mL for achieving the desired particle size and crystal structure.

3.4. Summary of Optimal Parameters for Drug Carrier Synthesis

The following table (Table 9) shows a series of optimal parameters identified from the experiments.

3.5. Drug Loading Simulation Experiment

The co-precipitation method was employed to synthesize calcium carbonate (CaCO3) drug carriers, incorporating the fluorescently labeled drug TRITC-Dextran directly into the CaCO3 synthesis process. The drug loading results of TRITC-Dextran (Figure 9b) shows that nearly all CaCO3 particles emitted fluorescence signals, indicating effective drug incorporation into the carrier matrix. The results of the loading capacity (LC), which is calculated by Equation (4), reveal that the LC for the CaCO3 drug carrier was 9.6 wt.%, compared with the LC of 10.3 wt.% achieved by Taman et al. [48] using liposomes for Gemcitabine (Gem) delivery. Typically, the LCs for liposome-based drug delivery systems range between 5 and 10 wt.%. In contrast, Donatan et al. [49], who utilized an adsorption method for loading TRITC-Dextran onto CaCO3, reported an LC of less than 1 wt.%; similarly, Parakhonskiy et al. [6], employing the adsorption method, achieved a maximum LC of only 0.01 wt.%. These comparisons highlight the efficiency of the co-precipitation method and its promising capability for drug carrier fabrication using CaCO3.

4. Conclusions

Compared to other calcium-rich biogenic wastes such as oyster and clam shells, coral remains at present to be a cleaner and more easily collected alternative. This study, initiated from a circular economy perspective, investigated the potential of repurposing coral remains, a topic typically associated with environmental challenges, into a valuable raw material for better use. The major composition of coral remains is CaCO3, a common raw material in industrial and pharmaceutical products. In this study, coral remains could meet the industrial standards for CaCO3 undergoing a series of purification and modification processes. Vaterite, one of the crystal phases of CaCO3, demonstrated excellent physicochemical properties, making it a promising material for anticancer drug carriers. The conclusions drawn from the experiments are as follows:
  • The composition analysis of coral remains showed the absence of heavy metals such as Pd and Cd, both of which are of most concern in food and pharmaceutical products, indicating their suitability as raw materials for these industries.
  • The efficient conversion of CaCO3 to CaO in an N2 atmosphere at 800 °C within just 5 min demonstrates a significant enhancement in process efficiency, potentially reducing both time and cost.
  • The whiteness level of purified CaCO3 reached 95%, surpassing many industrial standards (93%), making it suitable for industrial applications.
  • The synthesized vaterite possesses several desirable drug carrier properties such as spherical morphology, small particle size, high porosity, pH sensitivity, and biocompatibility. Given its widespread use in the food and pharmaceutical industries, CaCO3 has shown to be a superior option for anticancer drug delivery compared to other drug carriers such as mesoporous silica nanoparticles and gold nanoparticles.
  • The effective utilization of the EPR effect for targeting cancer cell necessitates particle sizes below 600 nm. Smaller particles are more efficient in penetrating cell membranes and also positively impact the drug loading capacity, a crucial factor of the efficiency of CaCO3 as a drug carrier. While typical reproducibility of vaterite particle size range from 4 to 6 µm [40,50,51], our study successfully synthesized drug carriers with an average size of 344 nm, and some even smaller than 200 nm.
  • During CaCO3 synthesis, the involvement of EG necessitated sufficient time to form a complete vaterite crystal. Insufficient time resulted in the formation of either amorphous phases or incomplete calcite crystals.
  • Given the high costs associated with cancer drug production and the expensive, intricately synthesized carriers typically used for their delivery, the overall price of anticancer drug carriers remains notably high. CaCO3, in contrast, is a widely available, cost-effective, and stable material. Its application as a drug carrier could significantly reduce costs while maintaining excellent biocompatibility and precision in drug delivery. This approach not only provides a solution to the high expenses involved in synthesizing drug carriers, but also tackles environmental concerns related to coral remains, offering a dual advantage.

Author Contributions

Conceptualization, S.-A.C. and W.-S.C.; methodology, W.-C.C.; software, S.-A.C.; validation, S.-A.C. and W.-C.C.; formal analysis, S.-A.C.; investigation, S.-A.C.; resources, S.-A.C.; data curation, S.-A.C.; writing—original draft preparation, P.-H.C.; writing—review and editing, P.-H.C.; visualization, P.-H.C.; supervision, W.-S.C.; project administration, S.-A.C.; funding acquisition, S.-A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science and Technology Council, Taiwan, under grant NSTC 111-2621-M-006-007.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Acknowledgments

The authors wish to extend their sincere gratitude to the members of the laboratory team for their invaluable administrative assistance throughout this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chhikara, B.S.; Parang, K. Global Cancer Statistics 2022: The trends projection analysis. Chem. Biol. Lett. 2023, 10, 45. [Google Scholar]
  2. Health Promotion Administration, Ministry of Health and Welfare. Cancer Registry Annual Report, 2020 Taiwan; 2020; pp. 1–669. Available online: https://www.hpa.gov.tw/Pages/Detail.aspx (accessed on 20 January 2024).
  3. Hami, Z. A brief review on advantages of nano-based drug delivery systems. Ann. Mil. Health Sci. Res. 2021, 19, e112274. [Google Scholar] [CrossRef]
  4. Singh, S. Nanomedicine–nanoscale drugs and delivery systems. J. Nanosci. Nanotechnol. 2010, 10, 7906–7918. [Google Scholar] [CrossRef] [PubMed]
  5. Zhao, Q.; Cheng, N.; Sun, X.; Yan, L.; Li, W. The application of nanomedicine in clinical settings. Front. Bioeng. Biotechnol. 2023, 11, 1219054. [Google Scholar] [CrossRef] [PubMed]
  6. Parakhonskiy, B.V.; Yashchenok, A.M.; Donatan, S.; Volodkin, D.V.; Tessarolo, F.; Antolini, R.; Möhwald, H.; Skirtach, A.G. Macromolecule Loading into Spherical, Elliptical, Star—Like and Cubic Calcium Carbonate Carriers. ChemPhysChem 2014, 15, 2817–2822. [Google Scholar] [CrossRef] [PubMed]
  7. Tannock, I.F.; Rotin, D. Acid pH in tumors and its potential for therapeutic exploitation. Cancer Res. 1989, 49, 4373–4384. [Google Scholar] [PubMed]
  8. Svenskaya, Y.; Parakhonskiy, B.; Haase, A.; Atkin, V.; Lukyanets, E.; Gorin, D.; Antolini, R. Anticancer drug delivery system based on calcium carbonate particles loaded with a photosensitizer. Biophys. Chem. 2013, 182, 11–15. [Google Scholar] [CrossRef] [PubMed]
  9. Wu, J.-L.; Wang, C.-Q.; Zhuo, R.-X.; Cheng, S.-X. Multi-drug delivery system based on alginate/calcium carbonate hybrid nanoparticles for combination chemotherapy. Colloids Surf. B Biointerfaces 2014, 123, 498–505. [Google Scholar] [CrossRef]
  10. Jimoh, O.A.; Ariffin, K.S.; Hussin, H.B.; Temitope, A.E. Synthesis of precipitated calcium carbonate: A review. Carbonates Evaporites 2018, 33, 331–346. [Google Scholar] [CrossRef]
  11. Wen, Y.; Xiang, L.; Jin, Y. Synthesis of plate-like calcium carbonate via carbonation route. Mater. Lett. 2003, 57, 2565–2571. [Google Scholar] [CrossRef]
  12. Doelle, K.; Amaya, J.J. Application of calcium carbonate for uncoated digital printing paper from 100% eucalyptus pulp. Tappi J. 2012, 11, 51–59. [Google Scholar] [CrossRef]
  13. Han, J.S.; Jung, S.Y.; Kang, D.S.; Seo, Y.B. Development of flexible calcium carbonate for papermaking filler. ACS Sustain. Chem. Eng. 2020, 8, 8994–9001. [Google Scholar] [CrossRef]
  14. Wang, J.; Wei, P.; Liu, P.; Sun, W. Identifying appropriate conditions for producing spindle-like causticizing precipitated calcium carbonate for paper filler applications. BioResources 2012, 7, 5894–5903. [Google Scholar] [CrossRef]
  15. Karakaş, F.; Hassas, B.V.; Celik, M.S. Effect of precipitated calcium carbonate additions on waterborne paints at different pigment volume concentrations. Prog. Org. Coat. 2015, 83, 64–70. [Google Scholar] [CrossRef]
  16. Mori, Y.; Isogai, A.; Enomae, T. Application of vaterite-type calcium carbonate prepared by ultrasound for ink jet paper. J. Imaging Sci. Technol. 2010, 54, 020504–020504-6. [Google Scholar] [CrossRef]
  17. Mohamed, M.; Yusup, S.; Maitra, S. Decomposition study of calcium carbonate in cockle shell. J. Eng. Sci. Technol. 2012, 7, 1–10. [Google Scholar]
  18. Wenyu, S.; Qingfeng, L.; Enguang, H.; Shoutian, C. Study on properties of polymers packed by aragonite whisker. In Proceedings of the 6th International Conference on Properties and Applications of Dielectric Materials (Cat. No. 00CH36347), Xi’an, China, 21–26 June 2000; IEEE: Piscataway, NJ, USA; pp. 431–434. [Google Scholar]
  19. Saulat, H.; Cao, M.; Khan, M.M.; Khan, M.; Khan, M.M.; Rehman, A. Preparation and applications of calcium carbonate whisker with a special focus on construction materials. Constr. Build. Mater. 2020, 236, 117613. [Google Scholar] [CrossRef]
  20. Ohgushi, H.; Okumura, M.; Yoshikawa, T.; Inboue, K.; Senpuku, N.; Tamai, S.; Shors, E.C. Bone formation processin porous calcium carbonate and hydroxyapatite. J. Biomed. Mater. Res. 1992, 26, 885–895. [Google Scholar] [CrossRef]
  21. Tolba, E.; Müller, W.E.; Abd El-Hady, B.M.; Neufurth, M.; Wurm, F.; Wang, S.; Schröder, H.C.; Wang, X. High biocompatibility and improved osteogenic potential of amorphous calcium carbonate/vaterite. J. Mater. Chem. B 2016, 4, 376–386. [Google Scholar] [CrossRef]
  22. Ramakrishna, C.; Thenepalli, T.; Ahn, J.W. Evaluation of various synthesis methods for calcite-precipitated calcium carbonate (PCC) formation. Korean Chem. Eng. Res. 2017, 55, 279–286. [Google Scholar]
  23. Vikulina, A.; Voronin, D.; Fakhrullin, R.; Vinokurov, V.; Volodkin, D. Naturally derived nano-and micro-drug delivery vehicles: Halloysite, vaterite and nanocellulose. New J. Chem. 2020, 44, 5638–5655. [Google Scholar] [CrossRef]
  24. Hubbe, M.A.; Gill, R.A. Fillers for papermaking: A review of their properties, usage practices, and their mechanistic role. BioResources 2016, 11, 2886–2963. [Google Scholar] [CrossRef]
  25. Tai, C.Y.; Chen, F.B. Polymorphism of CaCO3, precipitated in a constant—Composition environment. AIChE J. 1998, 44, 1790–1798. [Google Scholar] [CrossRef]
  26. Brečević, L.; Kralj, D. On calcium carbonates: From fundamental research to application. Croat. Chem. Acta 2007, 80, 467–484. [Google Scholar]
  27. Xie, A.-J.; Shen, Y.-H.; Zhang, C.-Y.; Yuan, Z.-W.; Zhu, X.-M.; Yang, Y.-M. Crystal growth of calcium carbonate with various morphologies in different amino acid systems. J. Cryst. Growth 2005, 285, 436–443. [Google Scholar] [CrossRef]
  28. Becker, A.; Ziegler, A.; Epple, M. The mineral phase in the cuticles of two species of Crustacea consists of magnesium calcite, amorphous calcium carbonate, and amorphous calcium phosphate. Dalton Trans. 2005, 2005, 1814–1820. [Google Scholar] [CrossRef]
  29. Toffolo, M.B. The significance of aragonite in the interpretation of the microscopic archaeological record. Geoarchaeology 2021, 36, 149–169. [Google Scholar] [CrossRef]
  30. Demichelis, R.; Raiteri, P.; Gale, J.D.; Dovesi, R. The multiple structures of vaterite. Cryst. Growth Des. 2013, 13, 2247–2251. [Google Scholar] [CrossRef]
  31. Friedman, G.; Schultz, D. Precipitation of vaterite (CaCO3) during oil field drilling. Mineral. Mag. 1994, 58, 401–408. [Google Scholar] [CrossRef]
  32. Park, W.K.; Ko, S.-J.; Lee, S.W.; Cho, K.-H.; Ahn, J.-W.; Han, C. Effects of magnesium chloride and organic additives on the synthesis of aragonite precipitated calcium carbonate. J. Cryst. Growth 2008, 310, 2593–2601. [Google Scholar] [CrossRef]
  33. Zhang, Z.; Xie, Y.; Xu, X.; Pan, H.; Tang, R. Transformation of amorphous calcium carbonate into aragonite. J. Cryst. Growth 2012, 343, 62–67. [Google Scholar] [CrossRef]
  34. Xiang, L.; Xiang, Y.; Wang, Z.; Jin, Y. Influence of chemical additives on the formation of super-fine calcium carbonate. Powder Technol. 2002, 126, 129–133. [Google Scholar] [CrossRef]
  35. Yu, J.; Lei, M.; Cheng, B.; Zhao, X. Effects of PAA additive and temperature on morphology of calcium carbonate particles. J. Solid State Chem. 2004, 177, 681–689. [Google Scholar] [CrossRef]
  36. Lam, R.S.; Charnock, J.M.; Lennie, A.; Meldrum, F.C. Synthesis-dependant structural variations in amorphous calcium carbonate. CrystEngComm 2007, 9, 1226–1236. [Google Scholar] [CrossRef]
  37. Isopescu, R.; Mateescu, C.; Mihai, M.; Dabija, G. The effects of organic additives on induction time and characteristics of precipitated calcium carbonate. Chem. Eng. Res. Des. 2010, 88, 1450–1454. [Google Scholar] [CrossRef]
  38. Meldrum, F. Calcium carbonate in biomineralisation and biomimetic chemistry. Int. Mater. Rev. 2003, 48, 187–224. [Google Scholar] [CrossRef]
  39. Parakhonskiy, B.V.; Haase, A.; Antolini, R. Sub-micrometer vaterite containers: Synthesis, substance loading, and release. Angew. Chem. 2012, 124, 1221–1223. [Google Scholar] [CrossRef]
  40. Trushina, D.B.; Bukreeva, T.V.; Antipina, M.N. Size-controlled synthesis of vaterite calcium carbonate by the mixing method: Aiming for nanosized particles. Cryst. Growth Des. 2016, 16, 1311–1319. [Google Scholar] [CrossRef]
  41. Flaten, E.M.; Seiersten, M.; Andreassen, J.-P. Polymorphism and morphology of calcium carbonate precipitated in mixed solvents of ethylene glycol and water. J. Cryst. Growth 2009, 311, 3533–3538. [Google Scholar] [CrossRef]
  42. Svenskaya, Y.I.; Fattah, H.; Inozemtseva, O.A.; Ivanova, A.G.; Shtykov, S.N.; Gorin, D.A.; Parakhonskiy, B.V. Key parameters for size-and shape-controlled synthesis of vaterite particles. Cryst. Growth Des. 2018, 18, 331–337. [Google Scholar] [CrossRef]
  43. Chawla, J.S.; Amiji, M.M. Biodegradable poly (ε-caprolactone) nanoparticles for tumor-targeted delivery of tamoxifen. Int. J. Pharm. 2002, 249, 127–138. [Google Scholar] [CrossRef] [PubMed]
  44. Trushina, D.B.; Borodina, T.N.; Belyakov, S.; Antipina, M.N. Calcium carbonate vaterite particles for drug delivery: Advances and challenges. Mater. Today Adv. 2022, 14, 100214. [Google Scholar] [CrossRef] [PubMed]
  45. 07014; Calcium Carbonate. Republic of China National Regulations Database Standards: Taipei, Taiwan, 2023.
  46. Zhang, Z.; Yang, B.; Tang, H.; Chen, X.; Wang, B. High-yield synthesis of vaterite CaCO3 microspheres in ethanol/water: Structural characterization and formation mechanisms. J. Mater. Sci. 2015, 50, 5540–5548. [Google Scholar] [CrossRef]
  47. Devos, C.; Van Gerven, T.; Kuhn, S. A review of experimental methods for nucleation rate determination in large-volume batch and microfluidic crystallization. Cryst. Growth Des. 2021, 21, 2541–2565. [Google Scholar] [CrossRef]
  48. Tamam, H.; Park, J.; Gadalla, H.H.; Masters, A.R.; Abdel-Aleem, J.A.; Abdelrahman, S.I.; Abdelrahman, A.A.; Lyle, L.T.; Yeo, Y. Development of liposomal gemcitabine with high drug loading capacity. Mol. Pharm. 2019, 16, 2858–2871. [Google Scholar] [CrossRef]
  49. Donatan, S.; Yashchenok, A.; Khan, N.; Parakhonskiy, B.; Cocquyt, M.; Pinchasik, B.-E.; Khalenkow, D.; Möhwald, H.; Konrad, M.; Skirtach, A. Loading capacity versus enzyme activity in anisotropic and spherical calcium carbonate microparticles. ACS Appl. Mater. Interfaces 2016, 8, 14284–14292. [Google Scholar] [CrossRef]
  50. Volodkin, D.V.; Petrov, A.I.; Prevot, M.; Sukhorukov, G.B. Matrix polyelectrolyte microcapsules: New system for macromolecule encapsulation. Langmuir 2004, 20, 3398–3406. [Google Scholar] [CrossRef] [PubMed]
  51. Flaten, E.M.; Seiersten, M.; Andreassen, J.-P. Growth of the calcium carbonate polymorph vaterite in mixtures of water and ethylene glycol at conditions of gas processing. J. Cryst. Growth 2010, 312, 953–960. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of drug release from a CaCO3 drug carrier.
Figure 1. Schematic illustration of drug release from a CaCO3 drug carrier.
Applsci 14 01336 g001
Figure 2. The decomposition rate under different atmospheres: (a) air; (b) N2; (c) CO2.
Figure 2. The decomposition rate under different atmospheres: (a) air; (b) N2; (c) CO2.
Applsci 14 01336 g002
Figure 3. Particle size distribution of CaCO3 synthesized with two different adding orders of the salts: (a) CaCl2-first (b) Na2CO3-first.
Figure 3. Particle size distribution of CaCO3 synthesized with two different adding orders of the salts: (a) CaCl2-first (b) Na2CO3-first.
Applsci 14 01336 g003
Figure 4. (a) Average particle size as a function of different stirring speeds. (b) SEM image of CaCO3 with 400 rpm stirring speed.
Figure 4. (a) Average particle size as a function of different stirring speeds. (b) SEM image of CaCO3 with 400 rpm stirring speed.
Applsci 14 01336 g004
Figure 5. SEM images of CaCO3 synthesized with insufficient concentrations of EG: (a) 0%; (b) 25%; (c) 50%; (d) 85%.
Figure 5. SEM images of CaCO3 synthesized with insufficient concentrations of EG: (a) 0%; (b) 25%; (c) 50%; (d) 85%.
Applsci 14 01336 g005
Figure 6. SEM images of CaCO3 synthesized with varying reaction times: (a) 5 min; (b) 30 min; (c) 60 min; (d) 180 min; (e) 1440 min.
Figure 6. SEM images of CaCO3 synthesized with varying reaction times: (a) 5 min; (b) 30 min; (c) 60 min; (d) 180 min; (e) 1440 min.
Applsci 14 01336 g006
Figure 7. SEM images of CaCO3 synthesized with varying initial salt concentrations of Na2CO3: (a) 0.05 M; (b) 0.1 M; (c) 0.33 M; (d) 0.5 M; (e) 1 M; (f) average particle size as a function of different initial salt concentrations.
Figure 7. SEM images of CaCO3 synthesized with varying initial salt concentrations of Na2CO3: (a) 0.05 M; (b) 0.1 M; (c) 0.33 M; (d) 0.5 M; (e) 1 M; (f) average particle size as a function of different initial salt concentrations.
Applsci 14 01336 g007
Figure 8. SEM images of CaCO3 synthesized with varying total reaction volumes: (a) 4 mL; (b) 40 mL; (c) 120 mL; (d) 400 mL; (e) average particle size as a function of different reaction volumes.
Figure 8. SEM images of CaCO3 synthesized with varying total reaction volumes: (a) 4 mL; (b) 40 mL; (c) 120 mL; (d) 400 mL; (e) average particle size as a function of different reaction volumes.
Applsci 14 01336 g008
Figure 9. Loading result of TRITC-Dextran: (a) synthesized calcium carbonate (CaCO3) only; (b) with the fluorescently labeled (red signals on the particles) drug TRITC-Dextran; (c) the fluorescence spectrum of the supernatant.
Figure 9. Loading result of TRITC-Dextran: (a) synthesized calcium carbonate (CaCO3) only; (b) with the fluorescently labeled (red signals on the particles) drug TRITC-Dextran; (c) the fluorescence spectrum of the supernatant.
Applsci 14 01336 g009
Table 1. Composition of raw coral remains.
Table 1. Composition of raw coral remains.
IngredientProportion
CaO92.95123%
MgO3.67358%
MnO20.00300%
Fe2O30.47249%
SiO22.13416%
Al2O30.77554%
PbON/A
CdON/A
Table 2. Composition of purified coral remains.
Table 2. Composition of purified coral remains.
IngredientProportion
CaO99.96482%
MgO0.00158%
MnO20.00006%
Fe2O30.00035%
SiO20.01910%
Al2O30.01409%
PbON/A
CdON/A
Table 3. Parameters for reaction base experiment.
Table 3. Parameters for reaction base experiment.
FactorsParameters
Adding OrderCaCl2-first, Na2CO3-first
Stirring Speed (rpm)1000 rpm
Concentration of EG85%
Reaction Time (min)60
CaCl2/Na2CO3 (M)1 M/0.1 M
Reaction Volume (mL)4
Table 4. Parameters for stirring speed experiment.
Table 4. Parameters for stirring speed experiment.
FactorsParameters
Adding OrderCaCl2-first
Stirring Speed (rpm)400, 600, 800, 1000, 1150
Concentration of EG85%
Reaction Time (min)60
CaCl2/Na2CO3 (M)1 M/0.1 M
Reaction Volume (mL)4
Table 5. Parameters for EG concentration experiment.
Table 5. Parameters for EG concentration experiment.
FactorsParameters
Adding OrderCaCl2-first
Stirring Speed (rpm)1000
Concentration of EG0%, 25%, 50%, 85%
Reaction Time (min)60
CaCl2/Na2CO3 (M)1 M/0.33 M
Reaction Volume (mL)4
Table 6. Parameters for reaction time experiment.
Table 6. Parameters for reaction time experiment.
FactorsParameters
Adding OrderCaCl2-first
Stirring Speed (rpm)1000
Concentration of EG85%
Reaction Time (min)5, 30, 60, 180, 1440
CaCl2/Na2CO3 (M)1 M/0.1 M
Reaction Volume (mL)4
Table 7. Parameters for initial salt concentration.
Table 7. Parameters for initial salt concentration.
FactorsParameters
Adding OrderCaCl2-first
Stirring Speed (rpm)1000
Concentration of EG85%
Reaction Time (min)60
CaCl2/Na2CO3 (M)1 M/0.05 M, 0.1 M, 0.33 M, 0.5 M, 1 M
Reaction Volume (mL)4
Table 8. Parameters for total reaction volume experiment.
Table 8. Parameters for total reaction volume experiment.
FactorsParameters
Adding OrderCaCl2-first
Stirring Speed (rpm)1000
Concentration of EG85%
Reaction Time (min)60
CaCl2/Na2CO3 (M)1 M/0.1 M
Reaction Volume (mL)4, 40, 120, 400
Table 9. Optimal parameters for CaCO3 synthesis.
Table 9. Optimal parameters for CaCO3 synthesis.
FactorsParameters
Adding OrderCaCl2-first
Stirring Speed (rpm)1150
Concentration of EG85%
Reaction Time (min)60
CaCl2/Na2CO3 (M)1 M/0.05 M
Reaction Volume (mL)4
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

Chen, P.-H.; Chen, W.-S.; Chen, S.-A.; Chen, W.-C. Synthesis of High-Precision Sub-Micron CaCO3 Anticancer Drug Carriers from Coral Remains. Appl. Sci. 2024, 14, 1336. https://0-doi-org.brum.beds.ac.uk/10.3390/app14041336

AMA Style

Chen P-H, Chen W-S, Chen S-A, Chen W-C. Synthesis of High-Precision Sub-Micron CaCO3 Anticancer Drug Carriers from Coral Remains. Applied Sciences. 2024; 14(4):1336. https://0-doi-org.brum.beds.ac.uk/10.3390/app14041336

Chicago/Turabian Style

Chen, Pin-Han, Wei-Sheng Chen, Szu-An Chen, and Wei-Chung Chen. 2024. "Synthesis of High-Precision Sub-Micron CaCO3 Anticancer Drug Carriers from Coral Remains" Applied Sciences 14, no. 4: 1336. https://0-doi-org.brum.beds.ac.uk/10.3390/app14041336

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