Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Accepted Manuscript

Large-scale synthesis of water-soluble luminescent hydroxyapatite nanorods for


security printing

Xiaohu Chen, Xiaoying Jin, Junjun Tan, Wei Li, Minfang Chen, Lan Yao,
Haitao Yang

PII: S0021-9797(16)30077-7
DOI: http://dx.doi.org/10.1016/j.jcis.2016.01.078
Reference: YJCIS 21061

To appear in: Journal of Colloid and Interface Science

Received Date: 10 November 2015


Revised Date: 28 January 2016
Accepted Date: 30 January 2016

Please cite this article as: X. Chen, X. Jin, J. Tan, W. Li, M. Chen, L. Yao, H. Yang, Large-scale synthesis of water-
soluble luminescent hydroxyapatite nanorods for security printing, Journal of Colloid and Interface Science (2016),
doi: http://dx.doi.org/10.1016/j.jcis.2016.01.078

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Title:

Large-scale synthesis of water-soluble luminescent hydroxyapatite

nanorods for security printing

Xiaohu Chen c, Xiaoying Jin a, Junjun Tan a*, Wei Li b, Minfang Chen b, Lan Yao c,
Haitao Yang c

Corresponding author (*):

Junjun Tan;

Address:

a. School of Chemistry and Chemical Engineering, Hubei University of Technology,

Wuhan, 430068, Hubei, P. R. China;

b. School of Materials Science and Engineering, Tianjin University of Technology,

Tianjin, 300384, P. R. China;

c. School of pulp and paper engineering, Hubei University of Technology, Wuhan,

430068, Hubei, P. R. China;

Tel: 0086 27 59750460

Fax: 0086 27 59750482

Corresponding author e-mail address: tanjunjun_martyr@126.com

Xiaohu Chen and Xiaoying Jin contributed equally to this paper.

1
Abstract

Luminescent hydroxyapatite nanoparticles, which have excellent

biocompatibility, excellent photostability, and strong fluorescence, have received

increasing attention as bioprobes in cell imaging. However, they are also excellent

candidates for use in ink-jet security printing. for related

applications highly crystalline, mono-dispersible hydroxyapatite

nanorods with good colloidal stability and high fluorescence

hydroxyapatite

hydroxyapatite

hydroxyapatite with

a slightly milky color

anti-counterfeiting technologies.

Keywords: Hydroxyapatite nanorods; Europium doping; Colloidal stability;


luminescence;

2
1. Introduction

Because of its excellent biocompatibility, biological activity, and strong ion

exchange capacity, hydroxyapatite (Ca10(PO4)6(OH)2), the mineral component of bone

and teeth, is widely used in fields such as bone tissue engineering[1, 2], heavy metal

ions absorption[3, 4], protein separation[5, 6] and drug carriers[7, 8], and so on.

Because calcium ions and lanthanide ions have similar atomic radii, hydroxyapatite

nanocrystals are a good host material for lanthanide doping, which may endow the

material with fluorescent properties[9, 10]. Because of these advantages,

lanthanide-doped hydroxyapatite nanoparticles have attracted considerable attention

not only as bioprobes in cell imaging, but also as excellent candidates for ink-jet

security printing[11-20]. Many attempts have been made to develop synthetic

strategies that will provide materials suitable for that purpose, including

hydrothermal[21-26], chemical precipitate[27], microwave[28-30] and ultrasound[31]

methods.

In these synthetic strategies, the desired products are highly crystalline,

mono-dispersible hydroxyapatite nanorods with good colloidal stability and high

fluorescence in aqueous media. These materials would be useful in related

applications, especially for ink-jet printing. However, in most of the synthetic

strategies mentioned above, the focus was restricted to one aspect alone, such as

morphology, colloidal stability, fluorescence performance, or crystallinity of the

hydroxyapatite particles. Few studies can be found that simultaneously achieve all of

the desirable characteristics listed above.

3
It has recently been recognized that there is an abundance of citrate molecules in

bone tissue—about 5.5 wt% of the organic matter in bone—and that most of these

molecules are on the apatite surface[32]. A density of about one molecule per two

square nanometers of crystalline apatite surface has been determined using solid state

NMR analysis of natural bone[33]. These citrate molecules play very important roles

in facilitating the dispersion of hydroxyapatite nanorods in bone tissue and in

controlling the growth of apatite crystals. Our recent work shows that citrate

molecules, which have excellent biocompatibility, strong ability to bind with calcium

ions, and great dispersing ability, can be used as a rapid, scalable, and chemically

green processing aid in the synthesis of water soluble hydroxyapatite nanorods[34,

35]. The crystallinity, morphology, size and colloidal stability of hydroxyapatite

nanorods can be easily tuned by controlling the hydrothermal conditions. Moreover,

processing time can be greatly reduced by increasing the hydrothermal temperature.

In view of these points, we believe that these types of treatments could be introduced

for the large-scale synthesis of water-soluble hydroxyapatite nanorods exhibiting high

fluorescence. These nanorods could be applied to ink-based printing for security

information storage and anti-counterfeiting technologies. To the best of our

knowledge, the use of Eu3+-doped hydroxyapatite nanorods for security printing has

not previously been reported.

Inspired by the studies mentioned above and our previous work[36-38], a

systematic experimental evaluation of the effect of Eu3+ doping on particle

morphology, crystallinity, colloidal stability and fluorescence performance of

hydrothermally synthesized colloidal hydroxyapatite nanorods was conducted. The

results of this evaluation and a proposed explanation of these effects are discussed. A

4
preliminary evaluation of security printing using these materials was also conducted.

2. Materials and methods

2.1 Materials

Calcium chloride anhydrous (CaCl2, ≥ 96.0%), sodium phosphate tribasic

dodecahydrate (Na3PO4·12H2O, ≥ 98.0%), sodium citrate tribasic dihydrate

(C6H5Na3O7·2H2O, ≥ 99.0%), ethanol absolute (C2H6O, ≥ 99.7%), sodium hydroxide

(NaOH, ≥96.0%) and nitric acid (HNO3, 65.0-68.0%) were purchased from

Sinopharm Chemical Reagent Co. Ltd. Europium oxide (Eu2O3, 99.9% metals basis)

were purchased from Aladdin Industrial Corporation. All chemicals were used as

received, without further purification. Deionized water was used throughout. Stock

solutions of Eu(NO3)3 were prepared by dissolving the respective Eu2O3 compounds

in nitric acid with heating.

2.2 Methods
Preparation of colloidal hydroxyapatite nanorods

Colloidal hydroxyapatite nanorods were prepared using a hydrothermal method.

In a typical experiment, a sodium citrate (C6H5Na3O7·2H2O) solution of fixed

concentration (0.033 M, 10 mL) was added, with continuous stirring, to an aqueous

solution of CaCl2 (0.05 M, 10 mL) and Eu(NO3)3 (x M, 10 mL) over 10 min. The

doping level was calculated as x/0.05. Then, an aqueous solution of Na3PO4·12H2O

(0.033+x M, 10 mL) was added to the mixture with vigorous stirring over 15 min.

After that, the mixed solution was transferred, as obtained, to a Teflon-lined stainless

steel autoclave with a 50 mL capacity. The solution in the autoclave underwent

5
hydrothermal treatment at 150 °C for 12 h. After the hydrothermal treatment, the

autoclave was allowed to cool down naturally, and the resulting product was purified

using a three-cycle centrifugation-washing (3700 g) process with deionized water and

ethanol. Finally, the purified product was re-dispersed in deionized water to form an

aqueous dispersion. The pH was adjusted to pH 9 by the addition of 0.1 M NaOH A

portion of the sample was dried at 70 oC for 12 h to get powder for future

characterization.

Printing with photoluminescent hydroxyapatite ink

A commercial paper that proved a suitable substrate for the hydroxyapatite

nanorods and featured no background fluorescence under the UV lamp was chosen for

use as the test paper. The aqueous dispersion of hydroxyapatite nanorods (50 mg/mL),

acting as ink, was injected into a commercial ink-jet printer. The desired Chinese or

English words were printed onto a piece of the test paper. The fluorescence images

were obtained under 254 nm UV light.

2.3 Characterization

Wide-angle X-ray diffraction

The dried powders were characterized using X-ray diffraction with an X’Pert PRO

X-ray diffractometer (PANalytical B.V., Almelo, Netherlands). Analyses were

performed using a Cu Kα radiation (1.5406 A) source at 60 kV and 60 mA from 5° to

80° with a scan rate of 0.5°/min.

Transmission Electron Microscopy

6
The morphology of each product was inspected with a Tecnai G2/F20

Transmission Electron Microscope (TEM) (FEI, Hillsboro, OR, USA) using an

accelerating voltage of 200 kV. Each colloidal hydroxyapatite nanoparticle sample

was drawn by pipette and 1-2 droplets were placed onto the carbon side of a wholly

carbon-coated Cu TEM-grid. The samples were dried before being attached to the

sample holder on the microscope. Elemental analysis was conducted using TEM with

Zeta Potential and DLS Measurements

Zeta potential and particle size distribution measurements were performed using a

nano-ZS90 Zetasizer (Malvern instruments Ltd., Malvern, UK). Disposable

polystyrene cuvettes containing dispersions (0.01 g/mL) were allowed to equilibrate

for 2 min at 25 °C before the particle size distributions were measured. The data were

recorded and analyzed using Dispersion Technology Software v. 5.0 (Malvern

Instruments). For zeta potential measurement, samples (0.5 wt% particle dispersion)

were deposited into clear disposable zeta cells and the results were automatically

generated with the help of Dispersion Technology Software v. 5.0 (Malvern

Instruments). The sample pH, which was modified using NaOH or HCl (0.1 M), was

determined using a PB-10 pH meter (Sartorius AG, Goettingen, Germany) at 25 °C.

No additional electrolytes were added to each sample for the measurement of zeta

potential and particle size distribution.

Excitation and emission spectra

The excitation and emission spectra were measured by using a fluorescence

7
spectrophotometer (F-7000) (Hitachi High-Technologies Corporation Tokyo Japan).

The samples were prepared at 0.26% concentration. Spectra were recorded with a

photomultiplier voltage of 400 V, a scan speed of 2400 nm/min, and excitation and

emission slit widths of 5 nm. Excitation spectra were recorded under 618 nm emission

wavelength and emission spectra were recorded under 244 nm excitation wavelength.

3. Results and discussion


3.1 Effect of Eu3+ doping level on the crystallinity of hydroxyapatite nanorods

Figure 1. X-ray diffraction patterns of hydroxyapatite nanorods with different doping levels.

The fluorescence of nanoparticles is remarkably affected by the crystallinity of

the particles. Figure 1 shows the

pure hexagonal phase,

(JCPDS files, PDF No.86-740) and with the literature[34, 36, 39], suggesting that the

samples are highly crystalline and that doping does not interfere greatly with

crystal growth in hydroxyapatite nanorods during hydrothermal treatment. When the

doping level is below 5 %, the intensities and profiles of peaks are very similar,
8
indicating the crystalline structure of hydroxyapatite could be well preserved. When

doping level is increased to 8 %, the intensities of the peaks decrease and the

integrity of the peaks weakens, indicating obvious interference with crystal growth.

3.2 Effect of Eu3+ doping level on the morphology of hydroxyapatite

nanoparticles

Figure 2. TEM images of hydroxyapatite nanorods with different doping levels (mole ratio),

(a)0%, (b)1%, (c)3%, (d)5%, (e)8%.

The doping level affects not only the crystallinity of hydroxyapatite nanocrystals,

but also their morphology. TEM measurements were conducted on samples with

different doping levels to provide direct information about the size and shape of

the hydroxyapatite nanorods in each sample (Figure 2). When the doping level is

9
1%, the average particle length of is 74 nm and the average diameter is 22 nm (Figure

2b). As the doping level reaches 3%, the average length decreases to 69 nm and the

diameter decreases to 20 nm (Figure 2c). With further increases in the doping

level, to 5% and 8%, the average particle length decreases to 64 nm and 50 nm,

respectively, and the average particle diameter decreases to 17 nm and 12 nm,

respectively (Figure 2d, e). Overall, both the particle length and particle diameter

show a downward trend with increasing doping. It is also worth noting that a

higher doping level (>8%) not only causes the decrease of particle size but also makes

particles more irregular and rough.

The morphological evolution of hydroxyapatite nanorods at different doping

levels can be explained as follows. On one hand, as the is introduced into the

hydroxyapatite crystals, some of the calcium positions in the crystal lattice are

occupied by

[40]
in . On the other hand, the tiny difference in

radius of and Ca2+ may also lead to a change in the hydroxyapatite lattice

parameters.

10
3.3 Effect of pH on the colloidal stability of hydroxyapatite nanorods dispersions

Figure 3. Zeta potentials (a) and DLS (b) of hydroxyapatite nanorods with 5% doping at

different pHs.

To be considered a serious candidate for use in ink-based printing, the

-doped hydroxyapatite nanorods must be well-dispersed in aqueous media. This

would allow the dispersion of hydroxyapatite nanorods to withstand long-term storage

and to easily pass through the printer nozzle. Previous research in the absence of Eu3+

doping has shown that adjusting the mole ratio of calcium salts to sodium citrate and

controlling the hydrothermal reaction temperature allows substantial control of the

colloidal stability of the resulting hydroxyapatite nanoparticles in the hydrothermal

11
synthesis system[34]. In consideration of this, the effect of Eu3+ doping on the

colloidal stability of hydroxyapatite nanoparticles in aqueous media was investigated.

The effect of pH on the colloidal stability of hydroxyapatite with 5% Eu3+ doping

is shown in Figure 3. When the pH of dispersion is 5, the zeta potential value is -15

mV and hydrodynamic diameter is 2849 nm. These data indicate that the citrate ions

adsorbed on the surface of the hydroxyapatite nanorods do not fully dissociate, so

there is insufficient electrostatic repulsion force between hydroxyapatite nanorods[41].

Compared with the individual particle size observed using TEM, it is inferred that the

-doped hydroxyapatite nanorods form large aggregates in the aqueous phase.

When the pH increases to 9, the zeta potential decreases rapidly to -35 mV and the

hydrodynamic diameter decreases to 124 nm. This indicates that there is sufficient

electrostatic repulsion among the -doped hydroxyapatite nanorods owing to the

full dissociation of the citrate ions adsorbed on their surfaces. In this case, the

hydroxyapatite nanorods are dispersed in the water in the form of individual particles.

Further increase in pH to 11 would slightly reduce the zeta potential value to -31 mV

and the hydrodynamic diameter to 159.5 nm. A reasonable explanation corresponding

to the reduction is that competitive adsorption between hydroxide ions and citrate ions

occur at high pH, which makes desorption of a small portion of citrate ions from the

particle surface. Thus, the optimal pH for the formation of dispersions of

hydroxyapatite nanorods modified with sodium citrate is about 9. Interestingly, the

effect of pH on the colloidal stability of hydroxyapatite particles in the absence of

Eu3+ doping is similar to the results shown above, indicating that the Eu3+ doping

12
process has a negligible effect on the adsorption of citrate ions to hydroxyapatite

nanorods surfaces.

3.4 Effect of Eu3+ doping level on the colloidal stability of hydroxyapatite

nanorods dispersions

Figure 4. Zeta potential values (a) and DLS (b) of hydroxyapatite samples with different Eu3+

doping levels. The pH of these dispersions is 9.

In addition to the effect of pH, the effect of the Eu3+ doping level on the colloidal

stability of hydroxyapatite was also investigated. Figure 4 displays the change in size

distribution and zeta potential of hydroxyapatite nanorods with different doping

levels in aqueous dispersions at pH 9. As the doping level changes from 1% to 6%,

the zeta potential values fluctuates within a narrow range from -35 mV to -40 mV and

13
the hydrodynamic diameter ranges from 120 to 180 nm. These data indicate that

varying the doping level from 1% to 6% does not change the colloidal stability

much. But when the doping level ranges from 8% to 10%, the results are a little

different. Although there is little difference between the zeta potentials of these

hydroxyapatite dispersions, a small DLS peak on the micrometer scale appears at both

of these doping levels. A possible explanation is that an excess of

. Thus, high levels of doping should be avoided to ensure good

colloidal stability of the hydroxyapatite dispersion.

3.5 Effect of Eu3+ doping level on the fluorescence performance of

hydroxyapatite nanorod dispersions

Figure 5. Excitation (a) and emission (b) spectra of hydroxyapatite dispersions with different

doping levels, and a particle concentration of 0.26% (g/mL). A photograph (c) of an aqueous

dispersion of hydroxyapatite nanorods excited using 254 nm UV light; the doping level is

5%.

14
Fig. 5 depicts the photoluminescence emission spectrum obtained from

hydroxyapatite dispersions with different doping levels excited with 244 nm UV light.

All of the samples show two strong emission peaks at 592 nm and 618 nm under these

conditions, indicating a strong red fluorescence[30]. The peak at 592 nm is due to the

5
D0 → 7F1 electronic transition and the peak at 618 nm is due to 5D0 → 7F2 of

electronic transition. When a cuvette containing a 0.26 wt% hydroxyapatite dispersion

was illuminated under a 254 nm UV lamp, the bright red color of the sample could

easily be observed (Figure 5c). When the doping level increased from 1% to 5%, the

emission intensity increased. At higher doping levels, up to 10%, the intensity of the

emission light decreased. The optimal emission intensity was seen at 5% doping. A

reasonable explanation for this variation in emission intensity is that, at doping levels

from 1% to 5%, the density of light-emitting centers increases and the crystallinity of

the hydroxyapatite nanorods is essentially unaffected, which lead to the increase in

fluorescence intensity. The crystallinity of the hydroxyapatite nanorods decreases

markedly when the doping level is increased to 10%, and this change plays a

dominant role in the decrease in fluorescence intensity. It has been reported that the

luminescence property of apatites doped with lanthanide ions is due to the substitution

of Ca2+ ions in the corresponding crystallographic sites of the apatite lattice by the

lanthanide ions [42, 43]. Therefore an increase in the crystallinity of a sample favors

the incorporation of the fluorescent ions in the correct positions to act as fluorescence

centers and to reduce the possibility of quenching. As a result of heat treatment, the

thermal diffusion of the fluorescent ions to the crystallographic site of Ca2+ occurred

15
and consequently a strong emission was achieved. This is why fluorescent intensity of

samples by chemical precipitation with additional heat treatment is far stronger than

that without[44].

3.6 Ink-based printing of patterns using a dispersion of Eu3+-doped

hydroxyapatite nanorods as the ink

Figure 6. Colloidal dispersion (a) and powder form (b) of doped hydroxyapatite nanoparticles

illuminated with 254 nm UV light; Chinese characters obtained using ink containing a doped

hydroxyapatite dispersion under natural light (c) or 254 nm UV light (d).

Based on the results presented above, we have demonstrated that hydroxyapatite


16
nanorods displaying a strong red fluorescence under UV light could be easily and

scalable synthesized. These nanorods can easily be dispersed into aqueous media to

form dispersions with hydrodynamic diameters of 100-150 nm. In

addition, strong red fluorescence was observed whether in colloidal dispersion (Fig.

6a) or in powder form (Fig. 6b) of Eu3+ doped hydroxyapatite nanoparticles under 254

nm UV light.

These properties are desirable for ink-based security printing, which can be used

for storage of confidential information or anti-counterfeiting measures. To verify this

statement, inks made from dispersions of hydroxyapatite nanorods were used to

printing. A commercially available paper with no background UV fluorescence, upon

which the hydroxyapatite nanorods adhered well, was chosen as the printing paper.

Transparent aqueous dispersions of hydroxyapatite nanorods with particle

concentrations of 50 mg/cm3 were injected into a commercial ink-jet printer. When

Chinese or English characters (fig. 6d) were printed using the ink containing the

hydroxyapatite particle dispersion, red words could be easily observed under a

portable 254 nm UV lamp. More importantly, the information display under UV light

could not be identified by eye under natural light. We could not find any other clue

that would allow us to acquire the printed information upon inspection of the

appearance of the paper (Fig. 6c). These results imply that inks like these are very

suitable for the long-term storage of confidential information and for

anti-counterfeiting measures.

17
4. Conclusions
In summary, a systematic evaluation of the hydrothermal synthesis of

Eu3+-doped hydroxyapatite nanorods with aid of sodium citrate was conducted. The

evaluation focused on the crystallinity, morphology, colloidal stability and

fluorescence performance of the Eu3+-doped hydroxyapatite nanorods. The synthetic

procedure is simple and chemically green. Moreover, by adjusting the experimental

parameters, such as the mole ratio of calcium to citrate molecules, the Eu3+ doping

level, and the pH of the colloidal dispersion, good colloidal stability and high

fluorescence performance could be maintained simultaneously. These results

demonstrate that colloidal dispersions of Eu3+-doped hydroxyapatite nanorods can be

applied to the preparation of fluorescent inks suitable for ink-based security printing.

5. Acknowledgments
Financially support for this work was obtained from the National Natural

Science Foundation of China (21203059), the National Training Program of

Innovation and Entrepreneurship for Undergraduates (201410500001) and the Doctor

Foundation of Hubei University of Technology (BSQD12043). We also thank Dr. Tod

P. Holler of ScienceDocs Inc. (https://www.sciencedocs.com/chemistry-editor-holler/)

for language editing.

18
6. References
[1] H. Zhou, J. Lee, Nanoscale hydroxyapatite particles for bone tissue engineering, Acta biomaterialia,

7 (2011) 2769-2781.

[2] J.S. Cho, H.-S. Kim, S.-H. Um, S.-H. Rhee, Preparation of a novel anorganic bovine bone xenograft

with enhanced bioactivity and osteoconductivity, Journal of Biomedical Materials Research Part B:

Applied Biomaterials, 101B (2013) 855-869.

[3] S.R. Sousa, M.A. Barbosa, Effect of hydroxyapatite thickness on metal ion release from Ti6Al4V

substrates, Biomaterials, 17 (1996) 397-404.

[4] C. Piccirillo, S.I.A. Pereira, A.P.G.C. Marques, R.C. Pullar, D.M. Tobaldi, M.E. Pintado, P.M.L.

Castro, Bacteria immobilisation on hydroxyapatite surface for heavy metals removal, Journal of

Environmental Management, 121 (2013) 87-95.

[5] M.J. Gorbunoff, The interaction of proteins with hydroxyapatite: I. Role of protein charge and

structure, Analytical Biochemistry, 136 (1984) 425-432.

[6] S. Yao, Y. Huang, Y. Zhao, Y. Zhang, X. Zou, C. Song, Iminodiacetic acid functionalized porous

hydroxyapatite nanoparticles for capturing histidine-tagged proteins, Materials Science and

Engineering: C, 39 (2014) 1-5.

[7] C. Zhang, C. Li, S. Huang, Z. Hou, Z. Cheng, P. Yang, C. Peng, J. Lin, Self-activated luminescent

and mesoporous strontium hydroxyapatite nanorods for drug delivery, Biomaterials, 31 (2010)

3374-3383.

[8] J.S. Son, Y.-A. Choi, E.-K. Park, T.-Y. Kwon, K.-H. Kim, K.-B. Lee, Drug delivery from

hydroxyapatite-coated titanium surfaces using biodegradable particle carriers, Journal of Biomedical

Materials Research Part B: Applied Biomaterials, 101B (2013) 247-257.

[9] J. Pan, D. Wan, Y. Bian, H. Sun, C. Zhang, F. Jin, Z. Huang, J. Gong, Fluorescent

hydroxyapatite-loaded biodegradable polymer nanoparticles with folate decoration for targeted imaging,

AIChE Journal, 59 (2013) 4494-4501.

[10] R.L. Williams, M.J. Hadley, P.J. Jiang, N.A. Rowson, P.M. Mendes, J.Z. Rappoport, L.M. Grover,

Thiol modification of silicon-substituted hydroxyapatite nanocrystals facilitates fluorescent labelling

and visualisation of cellular internalisation, Journal of Materials Chemistry B, 1 (2013) 4370.

19
[11] M.L. You, J.J. Zhong, Y. Hong, Z.F. Duan, M. Lin, F. Xu, Inkjet printing of upconversion

nanoparticles for anti-counterfeit applications, Nanoscale, 7 (2015) 4423-4431.

[12] Z.Y. Cheng, R.B. Xing, Z.Y. Hou, S.S. Huang, J. Lin, Patterning of Light-Emitting YVO4:Eu3+

Thin Films via Inkjet Printing, Journal of Physical Chemistry C, 114 (2010) 9883-9888.

[13] B.K. Gupta, D. Haranath, S. Saini, V.N. Singh, V. Shanker, Synthesis and characterization of

ultra-fine Y2O3:Eu3+ nanophosphors for luminescent security ink applications, Nanotechnology, 21

(2010) 8.

[14] E. Binetti, C. Ingrosso, M. Striccoli, P. Cosma, A. Agostiano, K. Pataky, J. Brugger, M.L. Curri,

Nanocomposites based on highly luminescent nanocrystals and semiconducting conjugated polymer for

inkjet printing, Nanotechnology, 23 (2012) 9.

[15] J.M. Meruga, W.M. Cross, P.S. May, Q. Luu, G.A. Crawford, J.J. Kellar, Security printing of

covert quick response codes using upconverting nanoparticle inks, Nanotechnology, 23 (2012) 9.

[16] J.M. Meruga, A. Baride, W. Cross, J.J. Kellar, P.S. May, Red-green-blue printing using

luminescence-upconversion inks, Journal of Materials Chemistry C, 2 (2014) 2221-2227.

[17] A.K. Soni, V.K. Rai, BaZnLa2O5:Ho3+-Yb3+ phosphor for display and security ink application, J.

Opt. Soc. Am. B-Opt. Phys., 31 (2014) 2201-2207.

[18] M.H. Xu, G.L. He, Z.H. Li, F.J. He, F. Gao, Y.J. Su, L.Y. Zhang, Z. Yang, Y.F. Zhang, A green

heterogeneous synthesis of N-doped carbon dots and their photoluminescence applications in solid and

aqueous states, Nanoscale, 6 (2014) 10307-10315.

[19] L.H. Shi, Y.Y. Li, X.F. Li, X.P. Wen, G.M. Zhang, J. Yang, C. Dong, S.M. Shuang, Facile and

eco-friendly synthesis of green fluorescent carbon nanodots for applications in bioimaging, patterning

and staining, Nanoscale, 7 (2015) 7394-7401.

[20] L. Zhou, A.D. Zhao, Z.Z. Wang, Z.W. Chen, J.S. Ren, X.G. Qu, Ionic Liquid-Assisted Synthesis of

Multicolor Luminescent Silica Nanodots and Their Use as Anticounterfeiting Ink, Acs Applied

Materials & Interfaces, 7 (2015) 2905-2911.

[21] X. Li, J. Zhu, Z. Man, Y. Ao, H. Chen, Investigation on the structure and upconversion

fluorescence of Yb3+/Ho3+ co-doped fluorapatite crystals for potential biomedical applications,

Scientific Reports, 4 (2014) 4446.

[22] J. Hui, X. Zhang, Z. Zhang, S. Wang, L. Tao, Y. Wei, X. Wang, Fluoridated HAp:Ln3+ (Ln = Eu or

Tb) nanoparticles for cell-imaging, Nanoscale, 4 (2012) 6967-6970.


20
[23] X. Zhang, J. Hui, B. Yang, Y. Yang, D. Fan, M. Liu, L. Tao, Y. Wei, PEGylation of fluoridated

hydroxyapatite (FAp):Ln3+ nanorods for cell imaging, Polymer Chemistry, 4 (2013) 4120-4125.

[24] X. Zheng, M. Liu, J. Hui, D. Fan, H. Ma, X. Zhang, Y. Wang, Y. Wei, Ln3+-doped hydroxyapatite

nanocrystals: controllable synthesis and cell imaging, Physical Chemistry Chemical Physics, 17 (2015)

20301-20307.

[25] C. Yang, P. Yang, W. Wang, J. Wang, M. Zhang, J. Lin, Solvothermal synthesis and

characterization of Ln (Eu3+, Tb3+) doped hydroxyapatite, J Colloid Interface Sci, 328 (2008)

203-210.

[26] J.Y. Chane-Ching, A. Lebugle, I. Rousselot, A. Pourpoint, F. Pell, Colloidal synthesis and

characterization of monocrystalline apatite nanophosphors, Journal of Materials Chemistry, 17 (2007)

2904.

[27] M. Neumeier, L.A. Hails, S.A. Davis, S. Mann, M. Epple, Synthesis of fluorescent core-shell

hydroxyapatite nanoparticles, Journal of Materials Chemistry, 21 (2011) 1250-1254.

[28] A. Escudero, M.E. Calvo, S. Rivera-Fernandez, J.M. de la Fuente, M. Ocana, Microwave-Assisted

Synthesis of Biocompatible Europium-Doped Calcium Hydroxyapatite and Fluoroapatite Luminescent

Nanospindles Functionalized with Poly(acrylic acid), Langmuir, 29 (2013) 1985-1994.

[29] D.E. Wagner, K.M. Eisenmann, A.L. Nestor-Kalinoski, S.B. Bhaduri, A microwave-assisted

solution combustion synthesis to produce europium-doped calcium phosphate nanowhiskers for

bioimaging applications, Acta Biomaterialia, 9 (2013) 8422-8432.

[30] C. Yang, P. Yang, W. Wang, S. Gai, J. Wang, M. Zhang, J. Lin, Synthesis and characterization of

Eu-doped hydroxyapatite through a microwave assisted microemulsion process, Solid State Sciences,

11 (2009) 1923-1928.

[31] Y.C. Han, X.Y. Wang, H.L. Dai, S.P. Li, Synthesis and luminescence of Eu3+ doped

hydroxyapatite nanocrystallines: Effects of calcinations and Eu3+ content, J. Lumines., 135 (2013)

281-287.

[32] J.M. Delgado-López, R. Frison, A. Cervellino, J. Gómez-Morales, A. Guagliardi, N. Masciocchi,

Crystal Size, Morphology, and Growth Mechanism in Bio-Inspired Apatite Nanocrystals, Advanced

Functional Materials, 24 (2014) 1090-1099.

[33] Y.-Y. Hu, A. Rawal, K. Schmidt-Rohr, Strongly bound citrate stabilizes the apatite nanocrystals in

bone, Proceedings of the National Academy of Sciences, 107 (2010) 22425-22429.


21
[34] X. Jin, X. Chen, Y. Cheng, L. Wang, B. Hu, J. Tan, Effects of hydrothermal temperature and time

on hydrothermal synthesis of colloidal hydroxyapatite nanorods in the presence of sodium citrate, J.

Colloid Interface Sci., 450 (2015) 151-158.

[35] X. Jin, J. Zhuang, Z. Zhang, H. Guo, J. Tan, Hydrothermal synthesis of hydroxyapatite nanorods in

the presence of sodium citrate and its aqueous colloidal stability evaluation in neutral pH, J. Colloid

Interface Sci., 443 (2015) 125-130.

[36] M. Chen, J. Tan, Y. Lian, D. Liu, Preparation of Gelatin coated hydroxyapatite nanorods and the

stability of its aqueous colloidal, Applied Surface Science, 254 (2008) 2730-2735.

[37] J. Tan, M. Chen, J. Xia, Water-dispersible hydroxyapatite nanorods synthesized by a facile method,

Applied Surface Science, 255 (2009) 8774-8779.

[38] J.J. Tan, M. Zhang, J. Wang, J. Xu, D.J. Sun, Temperature induced formation of particle coated

non-spherical droplets, J. Colloid Interface Sci., 359 (2011) 171-178.

[39] B. Hu, C. Zhao, X. Jin, H. Wang, J. Xiong, J. Tan, Antagonistic effect in pickering emulsion

stabilized by mixtures of hydroxyapatite nanoparticles and sodium oleate, Colloids and Surfaces A:

Physicochemical and Engineering Aspects, 484 (2015) 278-287.

[40] M. Long, F. Hong, W. Li, F. Li, H. Zhao, Y. Lv, H. Li, F. Hu, L. Sun, C. Yan, Z. Wei,

Size-dependent microstructure and europium site preference influence fluorescent properties of

Eu3+-doped Ca10(PO4)6(OH)2 nanocrystal, Journal of Luminescence, 128 (2008) 428-436.

[41] A. López-Macipe, J. Gómez-Morales, R. Rodriguez-Clemente, The Role of pH in the Adsorption

of Citrate Ions on Hydroxyapatite, J. Colloid Interface Sci., 200 (1998) 114-120.

[42] I. Mayer, J.D. Layani, A. Givan, M. Gaft, P. Blanc, La ions in precipitated hydroxyapatites,

Journal of Inorganic Biochemistry, 73 (1999) 221-226.

[43] R. Ternane, M. Trabelsi-Ayedi, N. Kbir-Ariguib, B. Piriou, Luminescent properties of Eu3+ in

calcium hydroxyapatite, J. Lumines., 81 (1999) 165-170.

[44] S.P. Mondejar, A. Kovtun, M. Epple, Lanthanide-doped calcium phosphate nanoparticles with

high internal crystallinity and with a shell of DNA as fluorescent probes in cell experiments, Journal of

Materials Chemistry, 17 (2007) 4153-4159.

22
Graphical abstract

23

You might also like