Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

| |

Received: 1 March 2021    Revised: 7 April 2021    Accepted: 7 April 2021

DOI: 10.1111/jace.17853

ORIGINAL ARTICLE

Development of microfibers for bone regeneration based on


alkali-­free bioactive glasses doped with boron oxide

Anuraag Gaddam1   | Przemysław Gołębiewski2,3   | Hugo R. Fernandes1   |


Dariusz Pysz2  | Ana S. Neto1   | Ryszard Diduszko2   | Agnieszka Malinowska2   |
Ryszard Stępień2   | Jarosław Cimek2   | Ryszard Buczyński2,3   | José M. F. Ferreira1

1
CICECO –­Aveiro Institute of Materials,
Department of Materials and Ceramic Abstract
Engineering, University of Aveiro, In this paper, we present a new series of alkali-­free bioactive glasses (BG) based
Aveiro, Portugal
2
on FastOs® composition (38.49 SiO2 –­36.07 CaO –­19.24 MgO –­5.61 P2O5 –­0.59
Łukasiewicz Research Network
–­Institute of Microelectronics and
CaF2, expressed in mol %), which was modified by partially replacing silicon diox-
Photonics, Warsaw, Poland ide network-­former with boron trioxide network-­former, utilizing calcium oxide as a
3
Faculty of Physics, University of charge compensator. The main objective of this study was to obtain a new family of
Warsaw, Warsaw, Poland
bioactive glasses suitable for the fabrication of glass fibers. The BGs were prepared
Correspondence by melt quenching technique and their structural and thermal properties were deter-
Anuraag Gaddam and José M. F. Ferreira, mined. Glass rods were used to obtain fibers by the classic drawing technique. The
CICECO –­Aveiro Institute of Materials,
bioactivity of the fibers was subsequently assessed through immersion tests in simu-
Department of Materials and Ceramic
Engineering, University of Aveiro, lated body fluid (SBF) to establish their ability to form hydroxyl carbonated (HCA)
Santiago University Campus, 3810-­193 apatite onto their surfaces. Glasses with moderate substitution of SiO2 with B2O3
Aveiro, Portugal.
Email: anuraagg@ua.pt (A. G.); jmf@
exhibited enhanced thermal properties, allowing to significantly suppress the crystal-
ua.pt (J. M. F. F.) lization trend, and favoring to draw the fibers. The structure of the studied glasses was
obtained by NMR spectroscopy. The structure-­property correlations were established
Funding information
Fundação para a Ciência e a by their relationship to the configurational entropy. Smaller amounts of substitution
Tecnologia, Grant/Award Number: resulted in larger entropy of the glasses. Moreover the SBF tests revealed an extensive
6818 -­311, UIDB/50011/2020
formation of HCA, comparable to the parent FastOs®BG composition, which assures
and UIDP/50011/2020; Centro de
Investigação em Materiais Cerâmicos fast bonding to the bone. Thus, presented glass fibers may be considered as promis-
e Compósitos, Grant/Award Number: ing materials for wool-­like bone implants or as reinforcing constituent of biopolymer
022161; Narodowa Agencja Wymiany
matrix composites.
Akademickiej, Grant/Award Number: PN/
BIL/2018/1/00253
KEYWORDS
bioactive glass, fibers, structure, structure-­properties

1  |   IN T RO D U C T IO N soft tissues.1,2 For over 50 years, researchers explored variety


of glass compositions, not only silicate3,4 but also borate-­5,6
Since 1969, when Larry L. Hench discovered 45S5 glass and phosphate7-­based glasses. However, only two composi-
(also known as Bioglass®) with nominal composition of: 46.1 tions: 45S5 (NovaMin®, Biogran®, Novabone®, Perioglas®)8
SiO2, 2.6 P2O5, 24.4 Na2O, 26.9 CaO (in mol %), glass mate- and S53P4 (53% SiO2, 4% P2O5, 23% Na2O, 20% CaO (in
rials have been considered very promising for biomedical ap- wt. %), also named BonAlive®)9,10 succeeded commercially.
plications with their outstanding ability to bond to bone and Nevertheless, both above-­mentioned materials exhibit several

|
4492    ©
wileyonlinelibrary.com/journal/jace
2021 The American Ceramic Society J Am Ceram Soc. 2021;104:4492–4504.
GADDAM et al.   
|
   4493

drawbacks. 45S5 glass may have a negative effect on remod- connectivity, partially compromising their bioactivity.24 One
eling of bone as its dissolution and resorption rates are rea- of the first bioactive glasses successfully used for fiber draw-
sonably fast, thus leading to the formation of a gap between ing was the composition labeled as 13–­93 (53 SiO2, 6 Na2O,
the implant and the bone.11 Moreover its high sodium oxide 12 K2O, 5 MgO, 20 CaO, 4 P2O5, in wt. %).24,25 The partial
content and the easiness of sodium ions’ leaching leads to sig- replacement of SiO2 in 13–­93 bioactive glass with B2O3 en-
nificant pH increase, causing cytotoxicity.12,13 Finally, 45S5 hanced the degradation rate and the surface mineralization of
exhibits poor sintering ability due to its very small difference trabecular scaffolds immersed in simulated body fluid (SBF).
between Tg and the onset of crystallization temperature (Tc), However, high B2O3 contents of the glass severely reduced
also called thermal stability (ΔT = Tc − Tg), 132°C.14 As a re- the ability of the scaffolds to support proliferation and func-
sult, thermal processing of Bioglass® is difficult and usually tion of osteogenic MLO-­A5 cells during conventional in vitro
leads to poor densification of, for example, 3D printed porous cell culture studies.26
scaffold causing poor mechanical properties. These scaffolds In this study, we investigated the possibility of drawing
are also partially crystallized which tends to hinder glass FastOs®BG-­based bioactive glass fibers from glass rods by
bioactivity.15 In turn, S53P4 exhibits slower dissolution rate the same classic method used for drawing glass fibers for
and better thermal stability as a result of its higher network photonic applications. After setting proper parameters, this
connectivity (NC). Nevertheless, clinical test proved that its process allows us to draw a large quantity of fibers with
dissolution rate is too slow in the case of large defect sizes, very precise control of their diameters. However, this pro-
and remnants were observed even 14  years post-­surgery.16 cess demands high thermal stability of the glasses to avoid
Attempting to overcome these drawbacks, a different con- crystallization. As mentioned above, early crystallization
cept was recently proposed by Ferreira et al.,17 aiming at leads to poor mechanical properties of obtained fibers, and
developing glasses with network connectivity similar to that in case of bioactive glasses it also results in decrease of bio-
of 45S5, while avoiding alkali oxides in glass structure that activity.15 Aiming at further enhancing ΔT, a new series of
are responsible for cytotoxicity. Accordingly, a new series of alkali-­free bioactive glasses based on the FastOs®BG com-
alkali-­free bioactive glasses based on ternary diopside (Di) position was prepared, in which silica was partially replaced
–­fluorapatite (FA) –­tricalcium phosphate (TCP) system by boron oxide and calcium oxide. Additional calcium oxide
were developed.17 In this series, TCP-­20 glass (70 Di –­10 is needed to maintain NC at the same level, as boron oxide
FA –­20 TCP, also known as FastOs®BG) showed outstand- will react with alkaline earth oxides to form fourfold coor-
ing properties. Its thermal stability allowed in obtaining fully dinated (BIV) units. Boron oxide will increase the entropy of
dense and amorphous compacts after sintering at 800°C. the investigated bioactive glasses and is expected to improve
Furthermore, the formation of hydroxyl carbonated apatite their thermal properties.27,28 Furthermore, boron oxide con-
(HCA) was detected just after 1 h of immersion in simulated centration may also help adjusting the glass dissolution
body fluid (SBF), anticipating a fast-­bonding ability to hard rate and tune the glass bioactivity.29,30 Addition of boron to
and soft tissues in comparison to other previously reported the glass may also induce positive effects on angiogenesis
bioactive glasses. Moreover amorphous sintered compacts and osteogenesis as suggested elsewhere.31–­33 The struc-
of FastOs®BG powder provided proper conditions for adhe- ture of glasses was assessed by Fourier Transform Infrared
sion and proliferation of mesenchymal stem cells (MSCs).17 Spectroscopy (FTIR) and Nuclear Magnetic Resonance (11B
Further studies revealed superior stimulation for differentia- NMR, 29Si NMR, and 31P NMR). The thermal properties
tion of human MSCs when compared to Bioglass®.18 Finally, of obtained glasses were assessed by differential scanning
in vivo tests in a sheep animal model showed slower resorp- calorimetry (DSC) and dilatometry (DIL) measurements.
tion and higher osteointegration and osteoconduction than The ability of BGs fibers to form hydroxyl carbonated apa-
observed for 45S5 used as control.19 In conclusion, it can be tite on their surfaces in vitro was investigated by immersing
stated that in the field of implants for large bone defects, the them in SBF. The surface changes occurred upon immersion
potential application of FastOs®BG is even greater than that of the fibers in SBF were assessed using X-­ray diffraction
of 45S5 Bioglass®. (XRD), FTIR spectra, and Scanning Electron Microscopy
Bioactive glasses may be incorporated into human skel- (SEM).
eton in different ways. Bone defects can be filled with glass
graft (single piece of glass), powder as a part of cement,20
cotton-­wool-­like form21 or as composites with bioresorbable 2  |  EXPERIM ENTAL
polymers.22,23 The composites may incorporate glasses either
as powders or in the fibers’ form. The fibers’ drawing re- A series of alkali-­free, silica-­based bioactive glasses with a
quires glasses with an extended working range to prevent the general formula: (38.49 − 2x) SiO2 · (x) B2O3 · (36.07 + x)
occurrence of crystallization. Achieving this target implies CaO · 19.24 MgO · 5.61 P2O5 · 0.59 CaF2 (where x = 0.25,
compositional changes towards increasing glass network 0.50, 1.00, 3.00, and 5.00 mol.%) was prepared by the melt
|
4494       GADDAM et al.

quenching technique. Glass samples were accordingly named intact fibers were analyzed by XRD. The surface morphology
B2x, where x corresponds to the amount of B2O3 in the glass of fibers before and after in vitro testing was observed using
composition. FastOs®BG glass was the base glass compo- SEM Hitachi SU-­70. Before observation, all samples were
sition with label B0.0. All glass compositions are presented covered with carbon using sputter coater.
in Table 1. Starting powder reagents of high purity were The FTIR measurements were performed using Vertex
used, which are: SiO2 (>99.5%), CaCO3 (>99.5%), MgCO3 80v FTIR spectrometer (Bruker). Glass samples, as-­drawn
(BDH Chemicals Ltd, >99.0%), NH4H2PO4 (Sigma–­Aldrich, glass fibers, and fibers after immersion in SBF were ground
>99.0%), and CaF2 (Sigma Aldrich, 325 mesh, >99.9%). in order to obtain fine powders. The as obtained powders were
Weighed powders were mixed carefully using mortar and then mixed with KBr in 1:30 weight ratio. All FTIR measure-
pestle and placed in Pt-­Rh crucibles and calcined at 900°C ments were carried out at room temperature within the range
for 1  h (heating rate 3°C  min−1) in the air. The calcined of 360–­4500 cm−1 with a resolution of 2 cm−1. Obtained data
batches were reground and melted at 1570°C for 1 h (heating were examined using OPUS software (Bruker).
rate 10°C min−1). Finally, the melts were cast into a copper The 29Si and 31P MAS-­NMR spectra were collected on a
mold to form bars and transferred to an annealing furnace Bruker Avance III HD 400 MHz (9.4 T) spectrometer, while
11
that was preheated to 800°C, followed by 1 h of dwelling at B MAS-­ NMR spectroscopy was performed on Bruker
this temperature, and then cooled down slowly. Avance III HD 700  MHz (16.4  T) narrow-­bore spectrome-
DIL measurements were performed using bulk cuboid-­ ter. For 29Si MAS-­NMR spectra, the Larmor frequency of
shaped glass samples (10 mm in length, cross section of c. 79.5 MHz was used while samples were spun at 12 kHz in
3  ×  3  mm), using a DIL 801 Bähr Thermoanalyse GmbH 7  mm ZrO2 rotors and the nuclei were excited with a 10°
apparatus. All the measurements were carried out in air from pulse at relaxation delay times of 60 s. For 31P MAS-­NMR
room temperature to 800°C with a heating rate of 5°C min−1. spectra, the Larmor frequency of 161.9 MHz was used while
DSC measurements were performed from 25°C to 1300°C samples were spun at 5 kHz in 4 mm ZrO2 rotors and the nu-
at a heating rate of 10°C  min−1 using Netzsch STA 449F1 clei were excited with a 90° pulse at relaxation delay times of
calorimeter. All measurements were carried out in pure argon 60 s. For 11B MAS-­NMR Hahn echo spectra, the Larmor fre-
flow utilizing alumina crucibles. quency of 224.6 MHz was used while samples were spun at
Glass rods were drawn into fibers’ form at fiber-­drawing 15 kHz in 4 mm ZrO2 rotors and the nuclei were excited with
tower. Temperature, feeding speed, and pulling speed were a 90° soft pulse at relaxation delay times of 2 s. The samples,
adjusted for each glass sample based on the data obtained Si(CH3)4 (0  ppm), H3PO3 (0  ppm) and H3BO3 (19.6  ppm),
from DIL and DSC measurements. Using this method glass were used as chemical shift references for 29Si, 31P and 11B,
fibers within 150−600 μm range could be obtained for each respectively. The NMR spectra were decomposed using the
glass composition. DMFit software.35
One meter of each fiber type was cut into small pieces and The presence of crystalline phases in bulk glasses was
immersed in SBF solution (prepared according to ISO 23317 evaluated by XRD. Measurements were performed using a
standard) up to 14  days in an incubator with shaking plat- Rigaku SmartLab 3  kW X-­ray diffractometer with Cu Kα
form at 37°C. Based on the diameter of each fiber, measured (λ = 1.5418 Å) radiation and silicon semiconductor strip (D/
using micrometer screw gauge, the amount of SBF was cal- teX Ultra 250) detector. Diffraction patterns of glass samples
culated, considering the surface area to volume of SBF ratio were recorded in the Bragg-­Brentano reflection geometry
of 0.5  cm2/ml as suggested by Popa et al.34 Aged samples (θ/2θ scan) and in continuous scanning mode. Qualitative
were filtrated using 63-­micron polymer sieve, washed with phase analysis of crystalline phases (if observed) was per-
deionized water, and slowly dried on a Petri dish in a closed formed using PDF4+ 2019 database and PDXL2 Software
desiccator filled with silica gel at the bottom. Afterward, the supplied by Rigaku.

T A B L E 1   Compositions of the glasses employed in the current


study
3  |  RESULTS
mol.% B0.0 B0.5 B1.0 B2.0 B6.0 B10.0 The experimental conditions used in the preparation of
SiO2 38.5 38.0 37.5 36.5 32.5 28.5 glasses resulted in amorphous samples except for B10.0,
CaO 36.1 36.3 36.6 37.1 39.1 41.1 which showed a small amount of devitrification due to sur-
MgO 19.2 19.2 19.2 19.2 19.2 19.2
face crystallization (Figure 1). The crystalline phase was
identified as hydroxyapatite substituted by silicate units.36
B2O3 0.0 0.3 0.5 1.0 3.0 5.0
But the samples with less than 10.0 mol% B are thermally
P2O5 5.6 5.6 5.6 5.6 5.6 5.6
more stable. Notwithstanding, the B10.0 sample was still
CaF2 0.6 0.6 0.6 0.6 0.6 0.6
considered in the rest of the study even with the small
GADDAM et al.   
|
   4495

amount of surface crystallization, which is only confined to 3.1  |  Thermal characterization


the surface of bulk cast glasses and could be mechanically
removed. Figure 2A presents DSC curves of the glasses and it can
be noticed that all glasses show two glass transition tem-
peratures, viz., Tg1 and Tg2. The presence of the dual glass
transition temperatures is due to the occurrence of liquid-­
liquid phase separation (LLPS) in the current glasses. The
LLPS was confirmed by SEM observation of a polished
and acid etched (with 2 vol% HF solution) B0.0 glass sam-
ple (Figure S1). The DSC characteristic temperatures, Tg1,
Tg2, Tc (crystallization onset temperature), Tp (peak crys-
tallization temperature), and Tm (melting point) were ob-
tained according to the procedure presented in the Figure
S2a; which was repeated 10 times to obtain averages and
standard errors. From which, the parameters, glass stability
(ΔT1, ΔT2) and Hrubý (KH1, KH2)37,38 were calculated ac-
cording to the formulae,

ΔT1 = Tc − Tg1 ; ΔT2 = Tc − Tg2 (1)

( ) ( )
Tc − Tg1 Tc − Tg2
KH1 = ( ) ; KH2 = ( ). (2)
Tm − T c Tm − T c

All the thermal parameters obtained from DSC are pre-


sented in Table 2. The DIL curves of the glasses are pre-
sented in Figure 2B. As glasses are isotropic materials, the
linear (αL) and volumetric (αV) thermal expansion coeffi-
cients are related to each other by, αV = 3αL. The αV values,
F I G U R E 1   X-­ray diffraction patterns of the glasses. Glass B10.0 dilatometric glass transition temperatures (Tg–­DL) and glass
shows small crystallization, which is identified as hydroxyapatite softening points (Ts) were obtained by DIL. The αL values
substituted with silicate group (Ca10(PO4)4.92(SiO4)1.08(OH)O0.664, were determined by a best-­ fit line between the tempera-
ICDD ref. code 01-­080-­6260) tures 100−650°C, and 500−650°C. The Tg–­DL values were

F I G U R E 2   Stacked plots of (A) DSC


and (B) DIL curves of the glasses. (ΔL is the
change in length from the original length,
Lo)
|
4496       GADDAM et al.

T A B L E 2   Thermal and physical properties of glasses

B0.0 B0.5 B1.0 B2.0 B6.0 B10.0


Tg1 (°C) 716.8 ± 0.3 715.8 ± 0.3 711.6 ± 0.5 698.9 ± 0.2 680.0 ± 0.5 672.6 ± 0.2
Tg2 (°C) 821.8 ± 0.2 825.8 ± 0.2 819.0 ± 0.2 812.3 ± 0.3 794.0 ± 0.3 777.8 ± 0.2
Tc (°C) 900.8 ± 0.3 906.7 ± 0.2 900.2 ± 0.3 886.6 ± 0.3 862.4 ± 0.3 830.8 ± 0.1
Tp (°C) 927 932 928 916 887 859
Tm (°C) 1274.1 ± 0.2 1255.4 ± 1.5 1247.6 ± 0.1 1257.7 ± 0.3 1208.0 ± 0.6 1138.1 ± 0.6
ΔT1 (°C) 184.0 ± 0.6 190.9 ± 0.5 188.6 ± 0.8 187.7 ± 0.5 182.4 ± 0.8 158.2 ± 0.4
ΔT2 (°C) 79.0 ± 0.4 80.9 ± 0.4 81.2 ± 0.4 74.3 ± 0.6 68.4 ± 0.5 53.0 ± 0.3
KH1 0.493 ± 0.002 0.547 ± 0.004 0.543 ± 0.003 0.506 ± 0.002 0.528 ± 0.003 0.515 ± 0.003
KH2 0.212 ± 0.001 0.232 ± 0.002 0.234 ± 0.001 0.200 ± 0.002 0.198 ± 0.002 0.172 ± 0.002
–­1
ΔH (J g ) 214.2 251.1 239.8 255.3 279.6 227.6
–­1 −6 −6 −6 −6 −6
αV 100-­650 (K ) 30.5 × 10 29.6 × 10 31.3 × 10 29.1 × 10 30.5 × 10 34.2 × 10−6
–­1 −6 −6 −6 −6 −6
αV 500-­650 (K ) 33.3 × 10 32.7 × 10 32.8 × 10 31.9 × 10 33.7 × 10 37.8 × 10−6
Tg–­DL (°C) 728.0 ± 0.4 727.2 ± 0.3 721.8 ± 0.8 718.8 ± 0.4 697.6 ± 0.4 688.4 ± 0.9
Ts (°C) 766 765 755 750 730 719

The Ts were obtained taking the peak maximum of the curve.


The values of αV, Tg–­DL, and Ts are presented in Table 2.
Using the DSC curves, the configurational entropy
(ΔSconf.) of the samples as a function of temperature (T or T ′)
relative to Tg1, was calculated using the formula:

T liquid ( �) glass ( � )
cp T − cp T

dT� (3)
( )
ΔSconf. T, Tg1 =
T�
Tg1

The parameters cp and cp are the specific heat capac-


liquid glass

ities for liquid and glass phases, respectively. At glass tran-


sition temperature, the sample changes from glass to liquid.
From DSC, the specific heat capacities for glass, cp , could
glass

be obtained at temperatures below Tg1. Extrapolating cp for


glass

temperatures above Tg1 and until Tc, the ΔSconf. can be calcu-
lated using Equation (3) in the range Tg1 to Tc. The calculated
values are presented in Figure 3. It can be seen that sam-
ples B0.0 and B1.0 exhibit larger values of ΔSconf. compared to
other samples, with sample B1.0 showing the largest values in
the entire range.

3.2  |  NMR and FTIR spectroscopy


F I G U R E 3   Configuration entropy of the samples calculated from
DSC: (A) normal scale, (B) reduced scale [Color figure can be viewed Figure 4 presents NMR spectra of the glasses for the nuclei
29
at wileyonlinelibrary.com] Si, 31P, and 11B. The 29Si NMR spectra can give informa-
tion about the distribution of QnSi (mT). According to the
QnSi (mT) notation, which is a standard notation for silicate
determined as intersection of two best fitting lines before units, QSi represents central Si tetrahedron, and n represents
and after glass transition, an example of the procedure is pre- the number of bridging oxygens (BOs) and m represents
sented in Figure S1b; this procedure was repeated 10 times the number of other tetrahedral or triangular units of type T
and the averages, and the standard errors were calculated. (P or B) attached to the central Si tetrahedron through the
GADDAM et al.   
|
   4497

T A B L E 3   Decomposition parameters of 31P NMR spectra and


their assignment

δiso width Amount


Type (ppm) (ppm) (%) Assignment
B0.5 Line 1 1.90 7.72 94.2 Q0P
Line 2 −5.87 7.00 5.8 Q1P
B1.0 Line 1 1.93 7.66 94.7 Q0P
Line 2 −6.00 7.00 5.3 Q1P
B2.0 Line 1 1.92 7.67 94.9 Q0P
Line 2 −5.83 7.00 5.1 Q1P
B6.0 Line 1 2.00 7.51 95.8 Q0P
Line 2 −5.31 7.00 4.2 Q1P

through charge dispersal effect40; or form BPO4 units similar


to AlPO4 units that mutually charge compensate.41 In both
scenarios, QnP (mB) type phosphate units are created. The 11B
nuclei are quadrupolar with spin 3/2 and thus experience sec-
ond order quadrupolar effects resulting in line broadening.
However, using the high magnetic field, such as in the cur-
rent study (16.4 T), the quadrupolar effects would be reduced
and useful structural information could be obtained. The 11B
MAS NMR could provide structural information about dif-
ferent three-­and four-­coordinated borate units given by the
notation, BIII and BIV, respectively.
In case of 29Si NMR spectra (Figure 4A), one can ob-
serve broad signals with maximum at c. 80 ppm, which cor-
responds to Q2Si (0T) units. The broad peak is skewed to the
right suggesting the presence of other QnSi (mT) units. With
increasing boron oxide and calcium oxide substitution, the
29
Si nuclei experience a deshielding effect causing an overall
shift of the spectrum towards left. This effect could be caused
due to the depolymerization of the silicate network or some
other complex phenomenon. For example, the formation of
every Si−O−BIV on a given Si tetrahedron, causes a desh-
ielding of( ~5 ppm of( 29Si )chemical(shift.)27,42 Therefore, the
units Q3Si 3BIV , Q2Si 2BIV and Q1Si 1BIV would experience
)

a deshielding of ~15, ~10, and ~5 ppm, respectively.


The results of 31P NMR (Figure 4B) show that in the in-
vestigated glass system, compositional changes have very
F I G U R E 4   NMR spectra of (A) 29Si, (B) 31P, and (C) 11B nuclei
small influence on the phosphate structure as the spectra
[Color figure can be viewed at wileyonlinelibrary.com]
look nearly identical with only small differences. All the
glasses show a slightly asymmetric spectra that are peaked
BOs. The 31P NMR spectra can give information about the at ~1 ppm. This suggests that majority of the phosphate net-
distribution of QnP (mT). The QnP (mT) notation for phosphate work is composed of Q0P (0T) units. This is due to the high
units is slightly different from QnSi (mT) (for silicate units).39 affinity of phosphorus oxide (which is an acid according to
Here, QnP (mT) represents a central phosphate tetrahedron that Lewis theory) towards alkaline earth oxides. Where, P2O5 re-
is connected to n other phosphate units through BOs and m acts with CaO and/or MgO to form orthophosphates.43 The
represents the number of T (B or Si) tetrahedral or triangu- slightly asymmetrical 31P NMR spectra was decomposed into
lar units that are connected through the BOs. Moreover it is two gaussian lines (Lines 1 and 2) (Figure S3) and the results
known that phosphate units could charge compensate tet- of the parameters are presented in Table 3. It can be noticed
rahedral borate units (BIV) by forming P−O−BIV linkages that with the increasing substitution of B units, the amount
|
4498       GADDAM et al.

of Line 2 decreases with a corresponding increase of Line expected (~13 ppm). This could possibly be due to the forma-
1. The possible assignment of Lines 1 and 2 are presented in tion of BIII−O−P bonds.45,46 The larger electron cloud on P
Table 3 and is discussed in greater detail in the subsequent units would cause greater chemical shielding on these units.
sections. Overall, with increasing boron content, the amount BIII units
Figure 4C presents results of 11B NMR, exhibiting two dis- increase relative to BIV units.
tinct broad peaks. These peaks at ~15 and ~2 ppm correspond Figure 5A presents results of FTIR spectra of glasses.
to BIII and BIV units, respectively. To quantify the amounts of All the glass compositions exhibit a broad band within wav-
different units, the spectra were decomposed using minimum enumber region from 850 to 1010  cm−1, due to the over-
number of lines that give the best possible fit. For BIII spe- lap of the vibrational bands of Q0Si (855  cm−1), Q1Si and Q2Si
cies, two lines with second order quadrupolar effects were (921 cm−1), Q2Si and Q3Si (1003 cm−1), and a band positioned
used. For BIV units, one or two mixed Gaussian/Lorentzian at ~1030–­1050  cm−1), attributed (Si–­O–­Si) in all QnSi sili-
line shapes were used. The results of the decomposition: iso- cate species.34 Figure 5B shows spectra of boron containing
tropic chemical shift (δiso), Quadrupolar Coupling Constants glasses with the spectra of B0.0 being subtracted from them.
(CQ) and the chemical shift anisotropy (η) are presented in It can be seen from Figure 5B that absorption bands grad-
Table 4 and Figure S2. From this, the total proportions of ually increase with increasing boron oxide contents. The
BIII and BIV were calculated. Going from B0.5 to B6.0, the BIV absorption band with maximum at 1438 cm−1 corresponds
units experience a slight deshielding effect resulting in the to B−O symmetric stretching vibrations of different borate
peak shifting to left. Consequently, the decomposition of 11B groups in BIII units. The intensity of the band at 1274 cm−1
NMR spectra show two peaks separated by ~0.5  ppm. The increases, and this may be related to the asymmetric stretch-
variation in the 11B chemical shift in this region could be ing vibrations of B−O bonds present in BIII units. Both above
caused by different BIV−O−P, BIV−O−Si, BIV−O−BIV, and mentioned bands suggest increasing contents of threefold
BIV−O−BIII bonds. The part of the 11B NMR spectrum cor- coordinated boron, which agrees with the 11B NMR spec-
responding to the BIII units was decomposed into two peaks tra. Furthermore, the absorption spectra exhibit a band with
viz. BIII‒­a, and BIII‒­b. The BIII‒­a peak is the major peak with maximum at ~724 cm−1, correlated with B−O−B bending
δiso of ~19.3  ppm, according to the NMR parameters, this vibration. And the bands at 807 and 585 cm−1, correspond
peak could be assigned to T2 species, which are BIII units to symmetrical vibrations in BIV units and bending vibration
with one NBO.44 The second peak for BIII units, BIII‒­b, has a O−B−O bridges, respectively. However, the last two bands
lower δiso value compared to BIII‒­a units and shifts to lower overlap with symmetrical SiO4 and O−Si−O bands respec-
values with increasing boron content. Moreover the chemi- tively, thus their intensity may be reduced.47 As a result of
cal shift anisotropy of BIII‒­b peak decreases with increasing decrease in silica molar fraction, one can also observe de-
boron content. This BIII‒­b peak could be assigned to T3 spe- creasing intensity of bands at 1172 and 454  cm−1, which
cies, which are BIII units with no NBO.44 However, the δiso correspond to Si−O−Si stretching vibration and Si−O rock-
value of these units is much lower (~9  ppm) than what is ing vibration, respectively.

T A B L E 4   Decomposition parameters of 11B NMR spectra

Amount
Sample Type δiso (ppm) CQ (GHz) η Amount Assignment Type (%)
B0.5 BIII‒­a 19.21 2.82 0.64 74 T2 BIII 84
III‒­b 3
B 10.14 2.69 0.10 10 T BIV 16
IV‒­a
B 5.49 —­ —­ 16
B1.0 BIII‒­a 19.1 2.65 0.71 85 T2 BIII 90
III‒­b 3 IV
B 10.0 2.86 0.05 5 T B 10
IV‒­a
B 3.7 —­ —­ 6
BIV‒­b 2.4 —­ —­ 4
III‒­a
B2.0 B 19.2 2.61 0.77 89 T2 BIII 92
III‒­b 3
B 8.9 2.52 0.09 2 T BIV 8
IV‒­a
B 2.8 —­ —­ 5
IV‒­b
B 2.4 —­ —­ 4
B6.0 BIII‒­a 19.8 2.75 0.65 92 T2 BIII 94
III‒­b 3 IV
B 8.7 2.23 0.05 1 T B 6
BIV‒­a 3.7 —­ —­ 6
GADDAM et al.   
|
   4499

F I G U R E 5   FTIR spectra: absorption spectra of obtained glass


F I G U R E 6   XRD patterns of as-­drawn fibers
series (A), absorption spectra with B0.0 spectra subtracted from other
glasses spectra (B) [Color figure can be viewed at wileyonlinelibrary.
com]
exhibited further crystallization towards silicate substituted
apatite phase.
3.3  |  Characterization of the fibers Figure 7 shows SEM micrographs of fibers before and
after immersion in SBF. Fibers immersed in SBF became
Of all the samples drawn to fibers on the fiber drawing tower, covered uniformly with a cracked layer resembling “tree
samples B0.0, B0.5 and B1.0 performed well. Other glass sam- bark.” Such appearance is very common in bioactive glasses
ples, undergone partial non-­uniform surface crystallization after in vitro testing and results from the surface dissolution
during the fiber drawing hampering the drawing process of the glass substrate and the concomitant precipitation of
(Figure S4). Thus, only short time interval was available for carbonated apatite. Under higher magnitude, the morphology
these glasses to be completely drawn. Nevertheless, sample of obtained layers can be easily observed. It consists of small
B1.0 performed the best with minimum crystallization until (c. 150  nm) particles connected together, forming porous
the end of the process. Similarly, B0.0 sample performed well, structure.
however, there were signs of partial crystallization on this The FTIR spectra of the SBF-­treated fibers are shown in
sample. Sample B0.5 showed small amount of crystallization Figure 8A. As a result of the fibers ground to powders, the
and its performance was in between B0.0 and B1.0. Therefore, surface composition became too diluted to be clearly detected
the performance of the samples during the fiber drawing pro- by FTIR. Therefore, the glass-­component of the spectrum
cess showed a non-­monotonic trend. was removed by subtracting the baseline spectrum in each
XRD measurements were performed on powders received case. The results are presented in Figure 8B. From Figure 8B,
by grinding glass fibers made by drawing process. XRD pat- the presence of the phosphate groups, namely, the asymmet-
terns are presented in Figure 6. In case of samples B0.0, B0.5, ric ν4 bending mode of (PO4)3− groups (~561 and 602 cm−1),
and B1.0, crystalline portion did not match any phase in data- symmetric ν1 stretching of (PO4)3− groups (~963 cm−1), and
base. In case of B2.0 and B6.0, the amounts of crystalline phases the bands arising from ν3 modes of PO4 groups (~1090 and
were slightly noticeable. However, further analysis resulted 1040  cm−1).48 Considering that fibers of similar diameters
in finding residual, insignificant amount of hydroxyapatite (200−250 μm) were chosen for the in vitro tests, these results
substituted with silicate group (Ca10(PO4)4.92(SiO4)1.08(OH) may be representative of amount of HA formed on the sur-
O0.664, ICDD ref. code 01-­080-­6260). Finally, B10.0 sample face of the fibers.
|
4500       GADDAM et al.

F I G U R E 7   SEM micrographs of fibers


before and after immersion in SBF solution
for 14 days

4  |   D IS C U SS ION of the BIV units that are present in the current glasses could
have been generated by the charge compensation by P tetrahe-
As mentioned in the introduction, in the current study, the dra instead of R. This would further enhance the silicate and
glasses were designed by replacing equimolar amounts SiO2 phosphate network with increased modifiers. The quantifica-
units with RBO2 units, where R ≡ Ca0.5 or Mg0.5. In this case tion of these bonds is difficult based on the current 1D NMR
if each borate unit, that is, BIV or BIII, has one R associated techniques. However, the presence of two peaks for BIV units
with it, either forming NBOs on BIII units or charge compen- in 11B NMR results suggest the formation of these units. It is
sating BIV units, the silicate and phosphate network connectiv- worth noting that sample B0.5 contains a greater amount of T3
ity should not be affected. Thus, preserving the chemistry and units compared to any other sample. As a result, the 29Si NMR
bioactivity of B0.0, which are the much desirable properties of spectra of B0.5 show a slightly greater deshielding effect that
B0.0 composition. The addition of RBO2 units would rise the B1.0 (Figure 4A).
overall configurational entropy of the glass system, which, in The configurational entropy (ΔSconf.) values from the
turn, lowers its viscosity and possibly increases its stability. DSC show a non-­monotonic trend with the increasing boron
Nevertheless, according to the 11B NMR results (Table 4), oxide addition. The structural interpretation of this behavior
there is a formation of T3 species of BIII units without any R is essential, which would provide deeper insights into the
associated, likely enhancing the depolymerization of the sili- materials’ behavior. Therefore, in this paper, we propose a
cate and phosphate networks. Accordingly, 29Si NMR spectra method to extract configurational entropy from the NMR
show increasing deshielding effect with the addition of boron. spectra. An NMR spectrum essentially contains short-­range
This deshielding effect is caused by the direct depolymeriza- structural information, that is, types of atoms and their po-
tion of the silicate network as mentioned above. However, it sitions around a given nucleus. For one atomic nucleus, this
is to be noted that part of this deshielding effect could also be information could be represented on multidimensional phase
caused by the formation of Si‒­O‒­BIV bonds. The presence of space.49 For a large number of atoms, there will be a distri-
a single BIV unit in the next nearest neighborhood of Si causes bution of information. However, in a 1D MAS-­NMR, this in-
an increase of the 29Si-­δiso values by ~5 ppm.42 Moreover, part formation is resolved onto a 1D line and is given by chemical
GADDAM et al.   
|
   4501

F I G U R E 9   Shannon entropy calculated from NMR spectra

this is in accordance with the configuration entropy calcu-


lated from the DSC measurements (Figure 3B). The HNMR
only accounts for the short-­range structure. Thus, the agree-
ment between the trends in HNMR and ΔSconf. suggest that the
NMR short range structure should be reliable for correlating
the trends in the relaxation dynamics of the current glasses.
The relaxation dynamics in glasses depends on the
available thermal energy and the volume of the configura-
tional space available to the system to explore.50 Therefore,
F I G U R E 8   (A) FTIR spectra of the glasses immersed in SBF
the configurational entropy, which is the measure of the
for 14 days and (B) FTIR spectra with the subtracted baseline
corresponding to the base glass spectra shown in Figure 5A [Color size of the configuration space, obtained from the DSC and
figure can be viewed at wileyonlinelibrary.com] NMR can provide insights into the dynamics/viscosity of
the glasses. That means, larger the entropy, the lower the
shift or frequency (ω). As a result, part of the information is viscosity. This was also used in phenomenological mod-
lost, nevertheless, NMR still provides sufficient useful struc- elling of viscosity of the glasses.51 According to Figure
tural information to probe the short-­range structure of the 3A, the viscosity of the glasses should steadily decrease
glasses. From this information, we propose a method for the with increasing boron content with exception of B1.0. This
calculation of the Shannon entropy, HNMR, corresponding to would facilitate the fiber drawing by lowering the draw-
the short-­range structure given by Equation (4) ing temperature. However, the fiber drawing also depends

on the glass stability. The glass stability values (ΔT and
KH) calculated from DSC measurements (Table 2; Figure

(4)
∑ ( )
HNMR = − nT fT (𝜔)log2 fT (𝜔) d𝜔
S5) show that for lower additions of boron, the stability
T = Si,B,P
−∞ increases; and at higher additions, it starts to decrease. The

combined effects of glass stability and viscosity dictate the
performance of the fiber drawing. Even though the sample

I(𝜔)
f(𝜔) = ; A= I(𝜔)d𝜔 (5)
A B0.5  has highest glass stability, it also has larger viscosity
−∞ (based on the entropy argument) at higher temperatures.
Thus, fibers for this sample can only be drawn at higher
The nT is the atomic fraction of network formers of type T. temperatures. Therefore, this glass did not perform as good
The function fT (𝜔) is the distribution function that is obtained as B0.0 or B1.0. The thermal stability increases slightly in
from the NMR spectra by normalizing the NMR intensity case of B0.5, B1.0, B2.0, a trend that is then reversed with
(I(𝜔)) as shown in Equation (5). Here, the entropy is calcu- further substitution of silica by B2O3. The most plausible
lated using frequency instead of chemical shift. However, as explanation for this reversion is the enhanced diffusivity
the function fT (𝜔) is normalized, this would not affect the of the species within the silicate network, as the silicate
final result. This procedure was applied to all NMR spectra network gets depolymerized. That results in continued
and the information entropy was calculated and is presented decreasing viscosity of the melts with further increasing
in Figure 9. These results also show a non-­monotonic trend B2O3 content, facilitating atomic arrangements and crys-
with B1.0 having the largest amount of entropy. Interestingly, tallization. Despite the variation of the variation in glass
|
4502       GADDAM et al.

composition and structure, all the glasses performed excep- Agnieszka Malinowska  https://orcid.
tionally well in the SBF solution resulting in the formation org/0000-0003-4835-6180
of hydroxy apatite layer (Figures 7 and 8). Deeper studies Ryszard Stępień  https://orcid.org/0000-0003-2352-6382
on the performance of the fibers in SBF and cell cultures Jarosław Cimek  https://orcid.org/0000-0002-1886-7986
will be reported in a future paper devoted to the biological Ryszard Buczyński  https://orcid.org/0000-0003-2863-725X
aspects of the performance of these fibers. José M. F. Ferreira  https://orcid.org/0000-0002-7520-2809

R E F E R E NC E S
5  |   CO NC LUS ION S 1. Hench LL. The story of Bioglass®. J Mater Sci Mater Med.
2006;17(11):967–­78.
The current study establishes the structure-­property rela- 2. Hench LL, Splinter RJ, Allen WC, Greenlee TK. Bonding mech-
anisms at the interface of ceramic prosthetic materials. J Biomed
tionships and their influence on fiber drawing ability of the
Mater Res Symp. 1971;5(6):117–­41.
boron containing bioactive glasses. Small additions of boron
3. Vale AC, Carvalho AL, Barbosa AM, Torrado E, Mano JF,
oxide improve the fiber drawing ability. Samples with lower Alves NM. Novel antibacterial and bioactive silicate glass
contents of boron (B0.5, B1.0, B2.0) exhibit improved ther- nanoparticles for biomedical applications. Adv Eng Mater.
mal stability in comparison with the original FastOs®BG. 2018;20(5):1700855.
However, with further increase of boron oxide content, the 4. Fagerlund S, Hupa L. Chapter 1: Melt-­derived bioactive silicate
lowering melt viscosity facilitates diffusion and atomic re- glasses. In: Boccaccini AR, Brauer DS, Hupa L, editors. Bioactive
arrangement, therefore decreasing the thermal stability of glasses: fundamentals, technology and applications. Smart mate-
rials series. Cambridge: Royal Society of Chemistry; 2016. p. 1–­
the glasses. This explains the presence of hydroxyapatite
26. https://pubs.rsc.org/en/conte​nt/chapt​erhtm​l/2016/bk978​17826​
substituted with silicate group identified in the as-­cast B10.0
29764​-­00001​?isbn=978-­1-­78262​-­976-­4&serco​de=bk. Accessed
glass. Fiber drawing led to slight crystallization trends in 24 Apr 2021.
some glass samples. Nevertheless, B2.0 and B6.0 were drawn 5. Cole KA, Funk GA, Rahaman MN, McIff TE. Characterization
with hardly detectable traces of crystallization as confirmed of the conversion of bone cement and borate bioactive glass
by XRD. These results show great potential of this glass composites. J Biomed Mater Res Part B Appl Biomater.
series to be applied in production of bioactive glass fibers. 2020;108(4):1580–­91.
Further optimization of technological process may result in 6. Rahimnejad Yazdi A, Torkan L, Stone W, Towler MR. The impact
of gallium content on degradation, bioactivity, and antibacterial
avoiding crystallization during fibers fabrication process
potency of zinc borate bioactive glass. J Biomed Mater Res Part B
entirely. Appl Biomater. 2018;106(1):367–­76.
7. Sanz CK, dos Santos AR, da Silva MHP, Marçal R, Tute EM,
ACKNOWLEDGMENTS Meza EL, et al. Niobo-­phosphate bioactive glass films produced
This work was developed in the framework of Portugal-­ by pulsed laser deposition on titanium surfaces for improved cell
Poland (FCT-­ NAWA) Scientific and Technological adhesion. Ceram Int. 2019;45(14):18052–­8.
Cooperation, 2019/2020 –­Project no. 6818 -­311 and Polish 8. Jones JR. Porous bioactive ceramic and glass scaffolds for bone re-
generation. In: Hench LL, editor. An introduction to bioceramics.
National Agency for Academic Exchange, grant number
2nd ed. London: Imperial College Press; 2013. p. 463–­85. https://
PN/BIL/2018/1/00253. The work was developed within the
www.world​scien​tific.com/doi/10.1142/97819​08977​168_0032.
scope of the project CICECO-­Aveiro Institute of Materials, Accessed 26 Apr 2021.
UIDB/50011/2020 & UIDP/50011/2020, financed by na- 9. Leppäranta O, Vaahtio M, Peltola T, Zhang D, Hupa L, Hupa
tional funds through the FCT/MEC and when appropri- M, et al. Antibacterial effect of bioactive glasses on clinically
ate co-­financed by FEDER under the PT2020 Partnership important anaerobic bacteria in vitro. J Mater Sci Mater Med.
Agreement. The NMR spectrometers are part of the National 2008;19(2):547–­51.
NMR Network (PTNMR) and are partially supported by 10. Fagerlund S, Massera J, Moritz N, Hupa L, Hupa M. Phase compo-
sition and in vitro bioactivity of porous implants made of bioactive
Infrastructure Project No 022161 (co-­financed by FEDER
glass S53P4. Acta Biomater. 2012;8(6):2331–­9.
through COMPETE 2020, POCI and PORL and FCT through
11. Sepulveda P, Jones JR, Hench LL. In vitro dissolution of melt-­
PIDDAC). derived 45S5 and sol-­gel derived 58S bioactive glasses. J Biomed
Mater Res. 2002;61(2):301–­11.
ORCID 12. Cannillo V, Chiellini F, Fabbri P, Sola A. Production of Bioglass®
Anuraag Gaddam  https://orcid.org/0000-0002-4266-6092 45S5 -­Polycaprolactone composite scaffolds via salt-­leaching.
Przemysław Gołębiewski  https://orcid. Compos Struct. 2010;92(8):1823–­32.
org/0000-0002-3892-2188 13. Rottensteiner U, Sarker B, Heusinger D, Dafinova D, Rath S,
Beier J, et al. In vitro and in vivo biocompatibility of alginate
Hugo R. Fernandes  https://orcid.org/0000-0002-9689-3127
dialdehyde/gelatin hydrogels with and without nanoscaled bio-
Ana S. Neto  https://orcid.org/0000-0002-8083-5758 active glass for bone tissue engineering applications. Materials.
Ryszard Diduszko  https://orcid.org/0000-0003-0590-6824 2014;7(3):1957–­74.
GADDAM et al.   
|
   4503

14. Groh D, Döhler F, Brauer DS. Bioactive glasses with improved 32. Gorustovich AA, Steimetz T, Nielsen FH, Guglielmotti MB.

processing. Part 1. Thermal properties, ion release and apatite for- Histomorphometric study of alveolar bone healing in rats fed a
mation. Acta Biomater. 2014;10(10):4465–­73. boron-­deficient diet. Anat Rec. 2008;291(4):441–­7.
15. Brauer DS. Bioactive glasses-­
structure and properties. Angew 33. Haro Durand LA, Góngora A, Porto López JM, Boccaccini

Chemie Int Ed. 2015;54(14):4160–­81. AR, Zago MP, Baldi A, et al. In vitro endothelial cell response
16. Lindfors NC, Koski I, Heikkilä JT, Mattila K, Aho AJ. A prospec- to ionic dissolution products from boron-­ doped bioactive
tive randomized 14-­year follow-­up study of bioactive glass and glass in the SiO2-­CaO-­P2O5-­Na2O system. J Mater Chem B.
autogenous bone as bone graft substitutes in benign bone tumors. J 2014;2(43):7620–­30.
Biomed Mater Res -­Part B Appl Biomater. 2010;94(1):157–­64. 34. Popa AC, Stan GE, Husanu MA, Mercioniu I, Santos LF,

17. Goel A, Kapoor S, Rajagopal RR, Pascual MJ, Kim H-­W, Ferreira Fernandes HR, et al. Bioglass implant-­coating interactions in syn-
JMF. Alkali-­free bioactive glasses for bone tissue engineering: a thetic physiological fluids with varying degrees of biomimicry. Int
preliminary investigation. Acta Biomater. 2012;8(1):361–­72. J Nanomedicine. 2017;12:683–­707.
18. Brito AF, Antunes B, dos Santos F, Fernandes HR, Ferreira JMF. 35. Massiot D, Fayon F, Capron M, King I, Le Calvé S, Alonso B,
Osteogenic capacity of alkali-­free bioactive glasses. In vitro studies. et al. Modelling one-­and two-­dimensional solid-­state NMR spec-
J Biomed Mater Res -­Part B Appl Biomater. 2017;105(8):2360–­5. tra. Magn Reson Chem. 2002;40:70–­6.
19. Cortez PP, Brito AF, Kapoor S, Correia AF, Atayde LM, Dias-­ 36. Gomes S, Nedelec JM, Jallot E, Sheptyakov D, Renaudin G.

Pereira P, et al. The in vivo performance of an alkali-­free bioac- Silicon location in silicate-­ substituted calcium phosphate ce-
tive glass for bone grafting, FastOs®BG, assessed with an ovine. J ramics determined by neutron diffraction. Cryst. Growth Des.
Biomed Mater Res Part B Appl Biomater. 2017;105(1):30–­8. 2011;11(9):4017–­26.
20. Shi Q, Wang J, Zhang J, Fan J, Stucky GD. Rapid-­setting, mes- 37. Hrubý A. Evaluation of glass-­forming tendency by means of DTA.
oporous, bioactive glass cements that induce accelerated in vitro Czechoslov J Phys. 1972;22(11):1187–­93.
apatite formation. Adv Mater. 2006;18(8):1038–­42. 38. Cabral A Jr, Fredericci C, Zanotto E. A test of the Hrubÿ pa-
21. Poologasundarampillai G, Wang D, Li S, Nakamura J, Bradley R, rameter to estimate glass-­ forming ability. J Non Cryst Solids.
Lee PD, et al. Cotton-­wool-­like bioactive glasses for bone regener- 1997;219:182–­6.
ation. Acta Biomater. 2014;10(8):3733–­46. 39. Lang DP, Alam TM, Bencoe DN. Solid-­
state 31P/27AI and
31 23
22. Zhang K, Wang Y, Hillmyer MA, Francis LF. Processing and P/ Na TRAPDOR NMR investigations of the phosphorus en-
properties of porous poly(L-­lactide)/bioactive glass composites. vironments in sodium aluminophosphate glasses. Chem Mater.
Biomaterials. 2004;25(13):2489–­500. 2001;13(2):420–­8.
23. Kolan KCR, Semon JA, Bindbeutel AT, Day DE, Leu MC.
40. Funke LM, Eckert H. Charge compensation in sodium boro-

Bioprinting with bioactive glass loaded polylactic acid composite phosphate glasses studied by 11B{23Na} and 31P{23Na} rota-
and human adipose stem cells. Bioprinting. 2020;18:e00075. tional echo double resonance spectroscopy. J Phys Chem C.
24. Brink M, Turunen T, Happonen R-­P, Yli-­Urpo A. Compositional 2016;120(6):3196–­205.
dependence of bioactivity of glasses in the system Na2O-­K2O-­MgO-­ 41. Yu Y, Stevensson B, Edén M. Medium-­range structural organization
CaO-­B2O3-­P2O5-­SiO2. J Biomed Mater Res. 1997;37(1):114–­21. of phosphorus-­bearing borosilicate glasses revealed by advanced
25. Fu Q, Rahaman MN, Bal BS, Huang W, Day DE. Preparation and solid-­state NMR experiments and MD simulations: consequences
bioactive characteristics of a porous 13–­93 glass, and fabrication of B/Si substitutions. J Phys Chem B. 2017;121(41):9737–­52.
into the articulating surface of a proximal tibia. J Biomed Mater 42. Nanba T, Nishimura M, Miura Y. A theoretical interpretation of
Res Part A. 2007;82A(1):222–­9. the chemical shift of 29Si NMR peaks in alkali borosilicate glasses.
26. Fu Q, Rahaman MN, Bal BS, Bonewald LF, Kuroki K, Brown Geochim Cosmochim Acta. 2004;68(24):5103–­11.
RF. Silicate, borosilicate, and borate bioactive glass scaffolds with 43. Yu Y, Edén M. Structure-­composition relationships of bioactive
controllable degradation rate for bone tissue engineering applica- borophosphosilicate glasses probed by multinuclear 11B, 29Si, and
31
tions. II. In vitro and in vivo biological evaluation. J Biomed Mater P solid state NMR. RSC Adv. 2016;6(103):101288–­303.
Res Part A. 2010;95A(1):172–­9. 44. Kroeker S, Stebbins JF. Three-­
coordinated boron-­ 11 chemical
27. Gaddam A, Fernandes HR, Ferreira JMF. Glass structure and crys- shifts in borates. Inorg Chem. 2001;40(24):6239–­46.
tallization of Al and B containing glasses belonging to the Li2O–­ 45. Tricot G, Raguenet B, Silly G, Ribes M, Pradel A, Eckert H. P-­
SiO2 system. RSC Adv. 2015;5(51):41066–­78. O-­B3 linkages in borophosphate glasses evidenced by high field
11 31
28. Gaddam A, Fernandes HR, Pascual MJ, Ferreira JMF. Influence of B/ P correlation NMR. Chem Commun. 2015;51(45):9284–­6.
Al2O3 and B2O3 on sintering and crystallization of lithium silicate 46. Muñoz-­
Senovilla L, Tricot G, Muñoz F. Kinetic fragility and
glass system. J Am Ceram Soc. 2016;99(3):833–­40. structure of lithium borophosphate glasses analysed by 1D/2D
29. Huang W, Day DE, Kittiratanapiboon K, Rahaman MN. Kinetics NMR. Phys Chem Chem Phys. 2017;19(34):22777–­84.
and mechanisms of the conversion of silicate (45S5), borate, and 47. Lai Y, Zeng Y, Tang X, Zhang H, Han J, Su H. Structural investi-
borosilicate glasses to hydroxyapatite in dilute phosphate solu- gation of calcium borosilicate glasses with varying Si/Ca ratios by
tions. J Mater Sci Mater Med. 2006;17(7):583–­96. infrared and Raman spectroscopy. RSC Adv. 2016;6(96):93722–­8.
30. Yao A, Wang D, Huang W, Fu Q, Rahaman MN, Day DE. In vitro 48. Markovic M, Fowler BO, Tung MS. Preparation and comprehen-
bioactive characteristics of borate-­based glasses with controllable sive characterization of a calcium hydroxyapatite reference mate-
degradation behavior. J Am Ceram Soc. 2007;90(1):303–­6. rial. J Res Natl Inst Stand Technol. 2004;109 [6]:553.
31. Gorustovich AA, López JMP, Guglielmotti MB, Cabrini RL.
49. Gaddam A, Montagne L, Ferreira JMF. Statistics of silicate units
Biological performance of boron-­ modified bioactive glass in binary glasses. J Chem Phys. 2016;145(12):124505.
particles implanted in rat tibia bone marrow. Biomed Mater. 50. Mauro JC, Smedskjaer MM. Statistical mechanics of glass. J Non
2006;1(3):100–­5. Cryst Solids. 2014;396–­397:41–­53.
|
4504       GADDAM et al.

51. Adam G, Gibbs JH. On the temperature dependence of coopera-


tive relaxation properties in glass-­forming liquids. J Chem Phys. How to cite this article: Gaddam A, Gołębiewski P,
1965;43(1):139–­46. Fernandes HR, et al. Development of microfibers for
bone regeneration based on alkali-­free bioactive glasses
doped with boron oxide. J Am Ceram Soc.
SUPPORTING INFORMATION
2021;104:4492–4504. https://doi.org/10.1111/
Additional supporting information may be found online in
jace.17853
the Supporting Information section.

You might also like