Non-conventional hexagonal structure for boric acid

9 downloads 0 Views 2MB Size Report
Sep 21, 2014 - diffraction analysis showed a new compound, named HBA, belonging to the ... has a high cross-section for absorption of thermal neutrons, is contributing ..... The strong IR bands that appear at 1,406.97 cm-1 (for. TBA) and ... water to obtain solutions of different concentrations (3,. 4000.0. 3600. 3200. 2800.
J Therm Anal Calorim (2014) 118:1375–1384 DOI 10.1007/s10973-014-4169-5

Non-conventional hexagonal structure for boric acid Ana Harabor • P. Rotaru • R. I. Scorei N. A. Harabor



Received: 10 June 2014 / Accepted: 7 September 2014 / Published online: 21 September 2014 Ó Akade´miai Kiado´, Budapest, Hungary 2014

Abstract When a special preparation procedure has been applied, the crystallization system of boric acid has been changed from triclinic to hexagonal: at temperatures between 60 and 70 °C, under controlled pH conditions, the boric acid belonging to triclinic system was mixed with Dglucose, calcium carbonate, and calcium hydroxide. Thermal analysis evidenced a final compound with quite similar thermal behavior as that of initial triclinic boric acid but having some differences in decomposition kinetics. X-ray diffraction analysis showed a new compound, named HBA, belonging to the hexagonal crystallization system with the ˚ and following lattice parameters: a = b = 20.4869 A ˚ . This strong anisotropic structure was also c = 12.1506 A confirmed by the hexagonal form of the crystallites, grown from HBA and water solutions, which have been observed with a light polarized optical microscope. Exotic polycrystalline conglomerates grown from water solution of HBA have nice colours that are changing when they are set different angles between polarizer and analyser. FTIR measurements revealed the IR absorbance bands belonging to O–H, O–B, and H–O–B bonds of the trigonal planar boric acid, for both crystallographic systems, but some small differences between wave-numbers and peak

A. Harabor  P. Rotaru (&) Department of Physics, University of Craiova, 13 AI Cuza Street, 200585 Craiova, Romania e-mail: [email protected]; [email protected] R. I. Scorei Department of Biochemistry, University of Craiova, 13 AI Cuza Street, 200585 Craiova, Romania N. A. Harabor Department of Physics, Politehnica University of Bucharest, 313 Splaiul Independent¸ ei Blvd., 060042 Bucharest, Romania

intensities were encountered. Finally, the dielectric properties of the water solutions of HBA are analyzed by performing electric susceptibility measurements at different temperatures, from 25 to 50 °C. Keywords Boric acid  Thermal analysis  XRD  FTIR  Optical polarized microscopy  Electric susceptibility

Introduction The boric acid has multiple uses. Some of industrial applications of boric acid are: manufacture of textile fiberglass [1]; combination of boric acid with denaturated alcohol contributes to the reduction of the surface oxidation in jewelry industry; uses of B(OH)3 in production of the glass in LCD flat panel displays; in neutralization of the active hydrofluoric acid (HF). Since 1948, boric acid was firstly registered in the US as an insecticide, because it acts as a stomach poison affecting the insects’ metabolism [2]; the dry powder of B(OH)3 is abrasive to the insects’ exoskeletons; in combination with an ethylene glycol is used to treat external wood against fungal and insect attack [2]; concentrates of borate-based treatments can be used to prevent slime, mycelium, and algae growth, even in marine environments [2]. Boric acid buffers against rising pH in swimming pools [2]. We mention also the remarkable lubricant properties of a colloidal suspension of nanoparticles of boric acid dissolved in petroleum or vegetable oil [3]. Boric acid is added in the reactor coolant which circulates through the reactor and due to the fact that boron-10 has a high cross-section for absorption of thermal neutrons, is contributing to effectively regulate the rate of fission taking place in the reactor. The medical uses of boric acid are explained by its antiseptic and antibacterial action,

123

1376

Experimental HBA preparation The hexagonal structure of the boric acid (HBA) was obtained from a mixture of triclinic boric acid (TBA), Dglucose, calcium carbonate, and calcium hydroxide at temperatures between 60 and 70 °C, under controlled pH conditions. Methods and techniques XRD measurements were performed on powder samples, with a Shimadzu XRD-6000 X-ray Diffractometer, equipped with a vertical goniometry and a scintillation detector. The functioning parameters of the X-ray tube (A40-Cu type) were established at a voltage of 40 kV and a current of 30 mA. The wavelength of CuKa type X-ray radiation ˚ . As operation mode, has a mean value of k = 1.54189 A we have chosen a continuous scan measurement in a geometry (h/2h), setting a scan 2h rate of 1°min-1 and a scan 2h range from 10° to 65°. Divergence slit was of 1.0000°, scattering slit was of 1.0000°, and receiving slit of 0.1500 mm. Thermal analysis measurements (TG, DTG, DTA, and DSC) of both TBA and HBA compounds were carried out in dynamic air atmosphere (150 cm3 min-1), from room temperature (RT) to 1,000 °C, with a rate of 10 K min-1. A horizontal analyzer ‘‘DIAMOND’’ TG/DTA from PerkinElmer Instruments was used during the measurements. The thermogravimetric and enthalpic calculations were performed with the specialized software Pyris. The chemical bonds of samples were investigated by FTIR with a PerkinElmer ‘‘SPECTRUM100’’ spectrometer

123

O

H

100

H B

O

80

O

H

60 TBA

40 Intensity/a.u.

being very appreciated in acne, otitis externa treatments, in eye washing and others [2, 4]. But one of the most important researches for the medicine are the use of acid boric in combination with calcium carbonate and fructose or glucose to produce future cancer drugs. It is known that the cancer cells consume glucose in a proportion of 1,000 times more than the normal cells, and therefore, by some mechanism, boron is also transported into the cell in large amounts being toxic to cancer cells [5, 6]. In this paper, we report on the results of a study concerning the modification of the crystallization system of the boric acid (TBA) from triclinic to hexagonal, when a special preparation procedure has been applied. Some properties of the new obtained compound, named HBA, are analysed by thermal, XRD, optical, and FTIR investigations. In plus, the dielectric behavior of HBA compound in water solutions of different concentrations is also presented.

A. Harabor et al.

20 0 100

12

24

36

48

60

36 2θ /°

48

60

80 HBA

60 40 20 0 12

24

Fig. 1 XRD patterns of TBA and HBA in 10°–65°. 2h range. In the insets are given the trigonal H3BO3 molecular unit (the left) as well as the structure of one sheet consisting of hydrogen-bonded B(OH)3 molecules

in the wavenumber range of 600–4,000 cm-1. The spectrum was obtained using universal attenuated total reflectance (UATR) accessory, at a resolution of 4 cm-1, with ten scans, and CO2/H2O correction. Optical microstructure observations have been made with an optical polarizing microscope LEICA DM 2500 P equipped with video camera for microphotographs of sample and a hot thermostatic stage TMS 94 (Linkam Scientific Instruments Ltd.) connected to the temperature programmer. The HBA mono-crystals as well as polycrystalline conglomerates, synthesized by growth from water solutions on a plate silica support, have been observed between crossed analyser and polarizer under a polarizing microscope. Dielectric properties of HBA in water solutions and its temperature dependence were examined with a Digital Refractometer RA-520 N. We measured the refractive index of the HBA 1 H2O solutions by detecting the critical angle of the total reflection. The device was equipped with a temperature control system with the build in Peltier elements in order to maintain and measure the border temperature between sample and prism. Results and discussion XRD analysis Since 1954, Zachariasen [7] discovered, by X-ray diffraction methods, that B(OH)3 belongs to the triclinic system  consisting of stacked layered sheets: (space group P 1),

Hexagonal structure for boric acid

1377

Table 1 Main XRD peaks of TBA Relative Intensity

˚ dhkl/A

Table 2 Main XRD peaks of HBA 2h/deg

hkl

Intensity/a.u.

˚ dhkl/A

2h/deg

hkl

100

3.21868

27.693

002

1,578

3.1865997

28.0034

510

68

6.08715

14.57

010

117

3.2321643

27.59

421

67

5.97277

14.82

100

84.3

3.1271833

28.54

313

31

2.96772

30.0879

200

40.3

1.5932999

57.88

1020

26

2.85336

31.3241

221

29.5

6.07535

14.58

002

24

2.26908

39.6900

311

25.8

5.8711377

15.09

211

23

2.24480

40.1377

210

19.9

2.640982

33.95

611

19

2.93780

30.4016

021

17.0

3.0823622

28.97

511

14

3.17520

28.0800

002

16.5

2.9355689

30.45

422

13

2.66602

33.5881

112

16.4

5.9140788

14.98

300

11 10

2.29969 4.25865

39.1400 20.8419

121 111

15.3 12.55

4.2377529 2.9168057

20.97 30.65

302 430

11.6

2.7056657

33.11

610

10.2

4.0502333

21.97

003

10

4.0703482

21.84

320

9.6

4.5609814

19.46

311

9

3.037675

29.4

004

each sheet contains H3BO3 molecular units (possessing a nearly perfect C3h symmetry) that are connected between them through hydrogen bonds (as seen from the inserts of Fig. 1). Ab initio calculations show that the binding energy between stacked sheets is about one-third of the intralayer interaction energy [8]. This facilitates the mutual sliding between layers, thus explaining the lubricant properties of boric acid. The Bragg diffraction patterns, indicate a good crystallinity for both TBA and HBA compounds, but surprisingly they are belonging to different crystallographic systems (Fig. 1). By XRD, performed with a Shimadzu XRD 6000 Diffractometer using a CuKa source, we indexed the unit cell of TBA compound, as belonging to the triclinic system. For indexing diffraction patterns of this compound named TBA, we were guided by a similar structure given by the cards no 30-0199 [7], taken from 2001 JCPDS-international Centre for Diffraction Data (PCPDFWIN v.2.2). The main peaks (intensity, inter-plane distance, dhkl, diffraction angles, 2h, and (hkl) Miller indexes for diffraction planes) in the case of triclinic boric acid (TBA) are given in Table 1. In conformity with the card no. 30-0199 [7], the parameters for the triclinic lattice of TBA are the follow˚, ˚, ˚, ings: a = 7.039 A b = 7.053 A c = 6.578 A a = 92.58°, b = 101.17°, and c = 119.83°. The new prepared compound, named HBA, has a different structure with a better symmetry, the unit cell being indexed as belonging to the hexagonal system (P 6 space group), in conformity with the card no 36-1481, taken from 2001 JCPDS-international Centre for Diffraction Data (PCPDFWIN v.2.2). The calculated lattice constants (a, b, and c) of the hexagonal unit cell have been determined by means of the well known Eq. 1 giving the relationship

between the inter-plane distance, dhkl, Miller indexes (hkl), and lattice constants, respectively [9, 10]. 1 4 h2 þ hk þ k2 ‘2 ¼ þ 2: 2 a2 c dhkl 3

ð1Þ

Table 2 shows the main XRD peaks of HBA compound (intensity, inter-plane distance, dhkl, diffraction angles, 2h, and (hkl) Miller indexes for diffraction planes). The experimental lattice parameters were calculated using the least square refinement from a Shimadzu adequate program, obtaining the following cell parameters: ˚ and c = 12.1506 A ˚ . This proves a a = b = 20.4869 A structural organization degree at long range. There are some approximation methods to define the breadth B in the Scherrer formula to calculate the average crystallite sizes [11–14]. In the assumption of ellipsoidal particles, in Ref. [11] it is found that the Jones more purely empirical approach will provide the best means for making allowance to the dimensions of the samples. In this last case, using the integral breadths BI, when KI = 1.333, we obtained for HBA particle dimensions a mean value of about 19.77 nm. Thermal behavior Thermal analysis is a method that is frequently used for determining the temperature range of the thermal stability, as well as, the thermal effects characterizing the phase transitions and decomposition processes taking place either in organic or inorganic compounds [15–23].

123

1378

A. Harabor et al.

–30

95

Heat flow endo up/mW

–20

90

–10

10

–40

0.6

DTA

–30

0.5

DSC

–20 –10 0

80

DTG 75

20

30

65

40

60 56.44 10

10 20

70

30

50

0.7

85

Mass/%

0

–50

0.4

100

200

300

400

500

600

700

800

900

0.3 0.2 0.1 0.0 –0.1 –0.2

40

–0.3

50

–0.4

60

–0.5

70 1000

–0.6

TG

Derivative mass/mg min–1

100.04

Microvolt endo up/μV

–40

Temperature/°C Fig. 2 Thermoanalytical curves for TBA

100.02

–30

95 –30

0.5

–20

0.4

DTA 85

–10

Mass/%

–10 80

DSC

75

DTG

0

10 70

10 20

20

42

0.3 0.2 0.1 0.0

–0.1 –0.2

65 30

0.6

Microvolt endo up/μV

Heat flow endo up/mW

0.7

90

–20

0

–45 –40

Derivative mass/mg min–1

–40

30 –0.3

60 54.81 10

40

–0.4

50 1000

–0.5

TG 100

200

300

400

500

600

700

800

900

Temperature/°C Fig. 3 Thermoanalytical curves for HBA

In Fig. 2 are shown TG, DTG, DSC, and DTA curves obtained for TBA compound. The results of thermal measurements for HBA are presented in Fig. 3. Examination of the thermal results exhibits a good thermal stability till 98 °C for TBA and a slight increase of the thermal stability limit until 101 °C for HBA. Over those temperatures, the decomposition of both samples is taking place, the fall in mass being accompanied by endothermic processes. The endothermic process is finished at 203 °C for both samples. After 203 °C it is observed a very small rate of mass decreasing until

123

1,000 °C, without any resizable thermal effect. After the total decomposition, we recorded a residuum of 56.44 % for TBA specimen, and of 54.81 % for HBA sample, respectively. A simple calculation shows that after the total decomposition of the boric acid, we should obtain a residuum of B2O3 in a cuantum of 56.32 %. The most evident differences between TBA and HBA are seen on DTG curves. From DTG curve of Fig. 2, is observed that, in the case of TBA sample, the decomposition is occured in three steps: the three peaks of the decomposition process have the maximums at 117.56,

Hexagonal structure for boric acid

1379

–30 Peak = 117.56 °C

Heat flow endo up/mW

–25 –20 Peak = 160.53 °C

–15

Peak = 131.32 °C

–10 –5 Area = 6453.211 mJ Delta H = 1409.9911 J g–1

0 5 10 50 60

X1 = 98.00 °C Y1 = 2.2009 mW

80

100

X2 = 203.00 °C Y2 = 4.4572 mW 120

140

160

180

200

220

240

260

280

300

240

260

280

300

Temperature/°C Fig. 4 The total endothermic effect for the non-isothermal decomposition of TBA

–40 Peak = 130.75 °C –35

Heat flow endo up/mW

–30 Peak = 160.84 °C

–25 –20 –15

Area = 8295.523 mJ Delta H = 1512.6495 J g–1

–10 –5 0 5 10 50 60

X1 = 101.00 °C Y1 = –0.4028 mW

80

100

120

X2 = 203.00 °C Y2 = –0.4806 mW

140

160

180

200

220

Temperature/°C Fig. 5 The total endothermic effect for the non-isothermal decomposition of HBA

131.32, and 160.53 °C. Each step corresponds to the breaking of the one of the three O–H bonds. The first step of the decomposition is accompanied by a major endothermic effect having the peak at the same temperature of 117.56 °C (as seen from DSC curve, given in Fig. 4). The second (at 131.32 °C) and the third stages (at 160.53 °C) of the decomposition are also endothermic processes. For the total endothermic effect, we calculated, within the specialized software Pyris, a value of the enthalpy variation equal with DH = 1.41 kJ g-1.

Concerning the HBA decomposition, it is seen from DTG curve of Fig. 3 that this is taking place in three stages too, but with a different kinetics compared with TBA: the thermal effect of the first stage is camouflaged by the second endothermic effect having the maximum at 130.75 °C (Fig. 5); the third stage of the decomposition has the maximum of the endothermic effect at 160.84 °C. The enthalpy variation of the total endothermic effect corresponding to the decomposition of HBA is DH = 1.51 kJ g-1. The calculations give an energy of

123

1380

A. Harabor et al. 0.450 703.58 706.95

1191.08 1409.03 1190.52 1406.97

0.40

b

b

3191.20

0.35

3187.60

Absorbance/a.u.

0.30

a

a

b a

0.25

0.20

0.15

883.00 2259.62

0.10

2513.29 0.05

0.000 4000.0

3600

3200

2800

2400

2000

1800

1600

1400

1200

1000

800

650.0

Wavenumber/cm–1

Fig. 6 Absorbance FTIR spectra of TBA (a-curve) and HBA (b-curve)

Table 3 Absorbance and wavenumber for the principal IR absorption peaks characteristic to TBA and HBA Maximum of IR absorption

Wavenumber/cm-1

Absorbance/a.u.

TBA

HBA

TBA

HBA

1

3,187.60

3,191.20

0.28

0.30

2

1,406.97

1,409.03

0.33

0.36

3 4

1,190.52 706.95

1,191.08 703.58

0.35 0.34

0.38 0.38

4.40 eV for the decomposition of a TBA molecule, while for the decomposition of a HBA molecule is needed an energy of 4.68 eV. Those values proved that the bound forces are much stronger for HBA molecules than for TBA molecules. FTIR spectra of TBA and HBA The FTIR absorbance spectra of the boric acid in the cases of the two crystalization systems are presented in Fig. 6. The two spectra are overlapped in order to evidence the small differences between the positions and intensities of the IR absorbance for TBA and HBA, respectively. For a better comparison between the two compounds, in Table 3 are given the IR wavenumbers, and the intensities for the principal IR absorption peaks obtained in the cases of TBA and HBA, respectively. Using the cards no 511 and no 4609, from the IR spectra database of the FTIR spectrometer SPECTRUM100, we

123

identified the IR absorbance bands having the peaks at 3,187.6, and 706.95 cm-1 (for TBA) and at 3,191.20, and 703.58 cm-1 (for HBA) as belonging to the O–H bond stretching vibration of the trigonal boric acid molecule [24]. The strong IR bands that appear at 1,406.97 cm-1 (for TBA) and 1,409.03 cm-1 (for HBA) can be assigned to the asymmetric B–O bond stretching mode of the trigonal planar boric acid [24]. The B–O–H in plane bending [24] appears for TBA at 1,190.52 cm-1, while for HBA at 1,191.08 cm-1. Comparing the intensities of the principal IR absorbance peaks, given in the Table 3, we concluded that the IR absorption is a little much stronger in the case of HBA compared to TBA. We recorded some small displacements in the positions of the IR absorption peaks for the HBA compared to TBA. This is probably due to the differences between the crystalization system characteristic to each compound.

Temperature dependence of the refractive index of HBA ? H2O solutions Knowledge of the refractive index of aqueous solutions of salts and biological agents is of crucial importance in many applications. Different methods have been developed to measure the refractive index of liquids [25–29]. The solution was prepared by dissolving the adequate amounts of prepared HBA powder in 10 mL of distilled water to obtain solutions of different concentrations (3,

Hexagonal structure for boric acid

1381

Refractive index

1.336

1.334

1.332 0% 3% 5.12 %

1.330

294

300

306

312

318

324

Temperature/K Fig. 7 The refractive index over temperature for H2O and two HBA ? H2O solutions (3, and 5.12 %)

Susceptivity

0.785

1 : T2 1 v3%HBA ¼ 0:71569 þ 5842 2 : T 1 v5:12%HBA ¼ 0:71604 þ 6036 2 : T

0%

vH

2O

3% 5.12 %

0.780 0.775

0.770

0.765 0.0000096

HBA solutions of 3 and 5.12 %, respectively. The experimental data of refractive index over absolute temperature indicate polynomial fits. It is observed that we have as higher value as higher is the solution concentration (under the saturation state). The dielectric constant er = n2 has been extracted from the experimental refractive index data. Then, the electric susceptibilities ve = er-1 have been calculated and their temperature dependence have been plotted (Fig. 8). As it was explained in Ref. [29], the water’s strong polar character makes it very good at shielding charged atoms from one another. This fact contributes to the creation of hydrates shell of interactions around charged molecules or ions and thus increasing the dielectric constant of water solutions of those compounds. This behavior is also confirmed in the case of HBA water solutions as observed from Figs. 7 and 5. The following linear functions are the theoretical fits for T-2 dependence of the electric susceptivity in the case of 0, 3, and 5.12 % of aqueous HBA solutions:

0.0000104

0.0000112

0.0000120

(1/T 2)/K–2 Fig. 8 The electric susceptivity is plotted as a function of T-2 for all HBA ? H2O investigated solutions. The solid lines are the theoretical fits

5.12 and 7 %). For this purpose, an electronic Precisa XT 120A balance with a precision of 0.0001 g has been used. With a digital refractometer, we measured the refractive indices for the liquid samples at desired temperatures in the range from 20 to 50 °C. The light source was a LED type with the characteristic wavelength of 589.3 nm (the Na-D line). A CCD-optical sensor was used to receive the reflected light by a prism made from sapphire [30]. A precision of ±0.00002 for the refractive index in the range from 1.32 to 1.58, allows us to obtain highly reliable data from measurements results [30]. A decrease of refractive index with increasing temperature is observed in Fig. 7 for distilled water as well for the

¼ 0:71689 þ 5208

ð2Þ ð3Þ ð4Þ

As observed, the second term is the polar one (&T-2). For pure water (0 % of HBA) b0 % = 5,208 K2 is the contribution given by the dipole moments of water polar molecules. But the factor b takes as higher values as higher are the concentrations of HBA in water solution: b3 % = 5,842 K2 and b5.12 % = 6,036 K2. Because of insignificant value for the ionic polarizability of water dipoles at optical frequencies, we may assume that the electric susceptibility could be expressed by the Eq. 5 [25–30]:   p20 v ¼ er  1 ¼ N 1 a 1 þ ð5Þ þ N2 ðaþ þ a Þ: 3e0 kT Here N1 is the number of water molecules, which is temperature dependent and have the electronic polarizability, a1, N? 2 is the number of positive ions of HBA compound with the electronic polarizability, a? and N2 is the number of negative ions of HBA compound with the electronic polarizability, a [25–29]. There is an evidence that N? 2 = N 2 = N 2. By comparing the two last expressions, we conclude that the electronic contributions of positive ions and negative ions to the polarizability are lower than in the case of pure water. Instead, it is dominant the polar behavior being as higher as the solution concentration is increased [30].

123

1382

Fig. 9 Image of grown crystals observed with polarizing microscope, arranged for reflection view in cases of two angles between the polarizer and analyser: a 0°; b 45°

A. Harabor et al.

Fig. 10 Hexagonal shape of the base of a HBA mono-crystal, grown from water solution. The image was obtained through transparency view with an objective lens of 920

Optical properties

123

0.030

S/mm2

Crystallization from solution can be thought of as a two step process. The first step is the phase separation (or birth) of a new crystals. The second is the growth of these crystals to larger size. These two processes are known as nucleation, and crystal growth, respectively. The birth of a new crystals, which is called nucleation, refers to the beginning of the phase separation process. The solute molecules have formed the smallest sized particles possible under the conditions present. The next stage of the crystallization process is for these nuclei to grow larger by the addition of solute molecules from the supersaturated solution. This part of the crystallization process is known as crystal growth. Crystal growth, along with nucleation, controls the final particle size distribution obtained in the system. In addition, the conditions and rate of crystal growth have a significant impact on the product purity and the crystal habit. Due to the stacked layered sheets, each sheet consisting of hydrogen-bonded of trigonal B(OH)3 molecules, and taking into account the primitive cell of hexagonal type ˚ , higher than c = 12.1506 A ˚ , the with a = b = 20.4869 A crystal growth from water solutions begins in the nucleation stage with the smallest sized particles in the shape of hexagonal prisms having the hexagonal base on the support, to provide a stable equilibrium. Crystal growth, along with nucleation, the conditions and rate of crystal growth control the final particle size distribution and the product purity. In Figs. 9 and 10 are seen the images of hexagonal crystallites of different dimensions, grown from HBA water solution (7 %) at normal pressure and room temperature (22 °C), having the c-axes perpendicular to the

0.035

0.025 0.020 0.015 0.010 0

100

200

300

400

500

600

t /s Fig. 11 Hexagonal prism base area versus time in the case of HBA grown from water solution (7 %)

silica support. The photographs have been obtained with video camera for microphotographs, using the polarizing microscope arranged in reflection view (Fig. 9), and transparency view (Fig. 10) equipped. From Fig. 11, it is observed that we have a non-linear increase over time for the area of the base belonging to HBA hexagonal prism grown from 7 % of HBA in water solution. In Figs. 12 and 13, are given some examples of HBA exotic polycrystalline structures obtained for different angles between polarizer and analyser. As expected from the XRD data, the HBA compound belonging to the hexagonal system should crystallize from water solutions in mono-crystals or polycrystalline structures that exhibit the characteristic microscopic structure.

Hexagonal structure for boric acid

1383

Fig. 12 Exotic polycrystalline structures grown from HBA ? H2O solution (of 7 %) obtained through transparency view for 0° (left) and 45° (right) between the polarizer and analyser

(a)

(b) 20 μm

20 μm

Green

˚ of lattice parameters (a = b = 20.4869 A and ˚ ), and of the mean value of the particle c = 12.1506 A dimensions (of about 19.77 nm). By optical microscopy, we observed the evident hexagonal prism structure of monocrystals grown from water solution of HBA, as well as the strong anisotropy of the exotic HBA polycrystalline structures. The dielectric properties of HBA & H2O solutions proved polar character of the compound. Thermal analysis of HBA, performed under a heating rate of 10 K min-1, has shown a good stability till 101 °C, and then decomposition into three stages, all endothermic, but having a different kinetics compared to TBA. While for TBA, we have all the three endothermic processes very clearly separated between them, for HBA the first stage of decomposition is camouflaged by the second endothermic effect having its maximum at 130.75 °C; the third stage of the decomposition has the maximum of the endothermic effect at 160.84 °C. The calculations give an energy of 4.40 eV for the decomposition of a TBA molecule, while for the decomposition of a HBA molecule is needed an energy of 4.68 eV. Those values proved that the bound forces are stronger for HBA molecules than for TBA molecules. By FTIR analysis evidenced, the presence of the absorption IR bands assigned to O–H, B–O, and B–O–H bonds, respectively, characteristic to boric acid, but small displacements of the IR peak positions and higher intensity for absorbance have been observed in the case of HBA compared to TBA.

Yellow

References

Fig. 13 The colors of polycrystalline structures grown from HBA ? H2O solution (of 7 %) are changed when the angle from analyser and polarizer is changed from 0° (a) to 90° (b)

These microscopic sculptures belonging to an ‘‘unknown author’’ have nice combination of colors (Figs. 12 and 13), that are changing in complementary colors when the angle between polarizer and analyser takes four values: 0°, 45°, 90°, and 135°. As an illustration, in Fig. 13 is seen how green color is changed into yellow when the angle between polarizer and analyser is changes from 0° to 90°. This behavior proves a strong anisotropy of all the characteristic physical properties of the compound.

Conclusions A special preparation procedure has leaded to the change of the crystallization system of boric acid from the triclinic to hexagonal, as confirmed by XDR analysis. The XRD diffraction maxima of HBA powder permitted the calculation

1. Jolly WL. Modern inorganic chemistry. 2nd ed. New York: McGraw-Hill; 1991. 2. Iavazzo C, Gkegkes ID, Zarkada IM, Falagas ME. Boric acid for recurrent vulvovaginal candidiasis: the clinical evidence. J Womens Health. 2011;20:1245–55. 3. Housecroft CE, Sharpe AG. Inorganic chemistry. 2nd ed. Mu¨nchen: Pearson Prentice-Hall; 2005. 4. Perelygin YuP, Chistyakov DYu. Boric acid. Russ J Appl Chem. 2006;79:2041–2. 5. Rotaru P, Scorei R, Ha˘ra˘bor A, Dumitru MD. Thermal analysis of a calcium fructoborate sample. Thermochim Acta. 2010;506:8–13. 6. Scorei RI, Rotaru P. Calcium fructoborate—Potential antiinflammatory agent. Biol Trace Elem Res. 2011;143:1223–38. 7. Zachariasen WH. The precise structure of orthoboric acid. Acta Cryst. 1954;7:305–10. 8. Zapol P, Curtiss L, Erdemir A. Periodic ab initio calculations of orthoboric acid. J Chem Phys. 2000;113:3338–43. 9. Cullity B. Elements of X-ray diffraction. Reading: AddisonWesley; 1978. 10. Pop V, Chicinas I, Jumate N. Material physics. Experimental methods. Cluj-Napoca: Presa Universitara Clujeana House; 2001. 11. Patterson AI. The Scherrer formula for X-ray particle size determination. Phys Rev. 1939;56:978–82. 12. Moant¸ a˘ A, Ionescu C, Rotaru P, Socaciu M, Harabor A. Structural characterization, thermal investigation, and liquid crystalline behavior of 4-[(4-chlorobenzyl)oxy]-3,40 -dichloroazobenzene. J Therm Anal Calorim. 2010;102:1079–86.

123

1384 13. Harabor A, Rotaru P, Harabor NA. Thermal and spectral behavior of (Y, Eu)VO4 powder. J Therm Anal Calorim. 2013;111: 1211–9. 14. Pascu CI, Gingu O, Rotaru P, Vida-Simiti I, Harabor A, Lupu N. Bulk titanium for structural and biomedical applications obtaining by spark plasma sintering (SPS) from titanium hydride powder. J Therm Anal Calorim. 2013;113:849–57. 15. Rotaru A. Thermal analysis and kinetic study of Petrosani bituminous coal from Romania in comparison with a sample of Ural bituminous coal. J Therm Anal Calorim. 2012;110:1283–91. 16. Samide A, Tutunaru B, Negrila˘ C, Dobrit¸ escu A. Study of the corrosion products formed on carbon steel surface in hydrochloric acid solution. J Therm Anal Calorim. 2012;110:145–52. 17. Rotaru A, Moanta A, Sa˘la˘geanu I, Budrugeac P, Segal E. Thermal decomposition kinetics of some aromatic azomonoethers. Part I. Decomposition of 4-[(4-chlorobenzyl)oxy]-40 -nitro-azobenzene. J Therm Anal Calorim. 2007;87:395–400. 18. Donato DI, Lazzara G, Milioto S. Thermogravimetric analysis. A tool to evaluate the ability of mixtures in consolidating waterlogged archaeological woods. J Therm Anal Calorim. 2010;101: 1085–91. 19. Rotaru A, Gos¸ a M, Rotaru P. Computational thermal and kinetic analysis. Software for non-isothermal kinetics by standard procedure. J Therm Anal Calorim. 2008;94:367–71. 20. Rotaru A, Gos¸ a M. Computational thermal and kinetic analysis. Complete standard procedure to evaluate the kinetic triplet form non-isothermal data. J Therm Anal Calorim. 2009;97:421–6.

123

A. Harabor et al. 21. Badea M, Olar R, Marinescu D, Segal E, Rotaru A. Thermal stability of some new complexes bearing ligands with polymerizable groups. J Therm Anal Calorim. 2007;88:317–21. 22. Rotaru A, Bratulescu G, Rotaru P. Thermal analysis of azoic dyes: part I. Non-isothermal decomposition kinetics of [4-(4chlorobenzyloxy)-3-methylphenyl](p-tolyl)diazene in dynamic air atmosphere. Thermochim Acta. 2009;489:63–9. 23. Degeratu S, Rotaru P, Manolea Gh, Manolea HO, Rotaru A. Thermal characteristics of Ni–Ti SMA (shape memory alloy) actuators. J Therm Anal Calorim. 2009;97:695–700. 24. Peak D, Luther GW III, Sparks DL. ATR-FTIR spectroscopic studies of boric acid adsorption on hydrous ferric oxide. Geochim Cosmochim Acta. 2003;67:2551–60. 25. Feynman RP, Leighton RB, Sands M. The Feynman lectures on physics, vol. 2. Massachusetts: Addison-Wesley Reading; 1964. 26. Jackson JD. Classical Electrodynamics. New York-LondonSydney-Toronto: Wiley; 1975. 27. Kasap S. Dielectric materials: static relative permittivity. Saskatchewan: Mcgraw–Hill; 2006. 28. Spanulescu I. Electricity and magnetism. Bucharest: Victor Publisher; 2001. 29. Tanaka M, Sato M. Microwave heating of water, ice and saline solution: molecular dynamics study. J Chem Phys. 2007;126: 034509. 30. Harabor A. Temperature effects on the electric susceptibility for a solution made from 10 mg of NaCl and 1 ml of H2O. Phys AUC. 2006;16:68–73.