Next Article in Journal
Sliding Wear Behavior of High-Temperature Vacuum-Brazed WC-Co-NiP Functional Composite Coatings
Previous Article in Journal
How Can We Provide Additively Manufactured Parts with a Fingerprint? A Review of Tagging Strategies in Additive Manufacturing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Pyroelectric Performance of Lead-Free Zn-Doped Na1/2Bi1/2TiO3-BaTiO3 Ceramics

by
Satyanarayan Patel
1,*,†,
Kodumudi Venkataraman Lalitha
2,*,† and
Nishchay Saurabh
1
1
Department of Mechanical Engineering, Indian Institute of Technology Indore, Khandwa Road Simrol, Indore 453552, Madhya Pradesh, India
2
Institute of Materials Science, Technische Universität Darmstadt, 64287 Darmstadt, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2022, 15(1), 87; https://doi.org/10.3390/ma15010087
Submission received: 21 November 2021 / Revised: 13 December 2021 / Accepted: 15 December 2021 / Published: 23 December 2021
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)

Abstract

:
Lead-free Na1/2Bi1/2TiO3-BaTiO3 (NBT-BT) has gained revived interest due to its exceptionally good high power properties in comparison to commercial lead-based piezoelectrics. Recently, Zn-modified NBT-BT-based materials as solid solution and composites have been reported to exhibit enhanced depolarization temperatures and a high mechanical quality factor. In this work, the pyroelectric properties of Zn-doped NBT-6mole%BT and NBT-9mole%BT ceramics are investigated. The doped compositions of NBT-6BT and NBT-9BT feature a relatively stable pyroelectric property in a wide temperature range of ~37 K (300–330 K) and 80 K (300–380 K), respectively. A threefold increase in detector figure of merit is noted for 0.01 mole Zn-doped NBT-6mole% BT at room temperature in comparison to undoped NBT-6mole%BT and this increase is higher than those of major lead-free materials. A broad range of the temperature-independent behavior for the figures of merit was noted (303–380 K) for Zn-doped NBT-6mole% BT, which is 30 K higher than the undoped material. The large pyroelectric figures of merit and good temperature stability renders Zn-doped NBT-BT an ideal candidate for pyroelectric detector and energy harvesting applications.

1. Introduction

Pyroelectric materials have been widely investigated for various industrial applications, such as infrared detectors, thermal cameras, sensors, medical devices, etc. [1,2,3]. In addition to these applications, waste thermal energy conversion for low-power electronic devices and battery-less wireless sensor applications receive significant industrial and academic interest [4]. In this direction, mostly Pb[Zr(x)Ti(1-x)]O3 (PZT), Pb(Mg1/3Nb2/3)O3-PbTiO3 and their solid solutions are studied due to their impressive pyroelectric, piezoelectric and dielectric properties [5,6,7]. However, the toxicity of lead oxide limits their use in electronic devices (European Union and many other countries) and demands alternative lead-free materials [2,8]. Therefore, various lead-free ferroelectric ceramics have been developed and documented in the literature for the past few decades, that demonstrate excellent pyroelectric properties [9,10]. Among the investigated lead-free materials, Na1/2Bi1/2TiO3 and its solid solution are promising because of their ease of fabrication. K0.5Na0.5NbO3 (KNN) and Ba0.85Ca0.15Zr0.1Ti0.9O3 (BCZT) also exhibit promising piezoelectric and ferroelectric properties; however, fabrication of KNN is difficult compared to NBT-based compositions [1,11]. Na1/2Bi1/2TiO3-xBaTiO3 (NBT-BT) has been studied by various researchers, at 6–7 mole% BT, wherein it exhibits a morphotropic phase boundary (MPB) and features high remanent polarization, large piezoelectric constant (d33) and low coercive field [12,13]. Furthermore, at the MPB, poling is much easier due to the low coercive field, favorable for pyroelectric applications [14]. NBT-xBT phase diagram is complex and presents various phase states/transitions, such as ferroelectric (FE), non-polar (non-ergodic relaxor) and paraelectric (PE) rendering it beneficial, as one can utilize the large pyroelectric currents at the vicinity of these phase transitions [15,16].
0.94Na1/2Bi1/2TiO3-0.06BaTiO3 (NBT-6BT) has been extensively investigated from the ferroelectric, dielectric and piezoelectric application points of view; however, there are few publications on pyroelectric properties [17,18,19,20]. Takenaka et al. have reported a high pyroelectric coefficient (p) = 3.9 × 10−4 C m−2 K−1 for NBT-6BT [21]. In another study, Felix et al. reported pyroelectric behavior of NBT-xBT (4 ≤ x ≤ 6 mole%) and demonstrated that NBT-6BT exhibits higher p compared to other compositions (attributed to MPB) and features a similar p akin to PZT [22]. Guo et al. have studied Zr-doped NBT-xBT (0% ≤ x ≤ 12%) ceramics and found that p varies between 5.7 × 10−4 and 22.1 × 10−4 C m−2 K−1 from room temperature to depolarization temperature (Td) [10]. Abe et al. investigated MnO2-doped NBT-6BT and reported p = 3.5 × 10−4 C m−2 K−1 at room temperature [23]. Recently, Balakat et al. have been extensively investigating the pyroelectric properties of 0.94NaxBiyTiO3-0.06BazTiO3 [24], 0.94Na1/2Bi1/2TiO3-0.06Ba1+xTiO3 [25], La-doped NBT-6BT [26], Ta-doped NBT-6BT [27], and La, Ta-doped NBT-6BT [28]. The depolarization temperature (Td) is lowered using these doping strategies; however, only a small improvement was observed in the pyroelectric figures of merit (FOMs). Here, Td plays an important role from the practical application point of view, since the pyroelectric and piezoelectric properties vanish above this temperature. Lalitha et al. reported on Zn-doped NBT-6BT and NBT-9BT, that featured an increase in Td to 423 K and 469 K, respectively [29]. Saurabh N. et al. used nonstoichiometric 0.94NaxBiyTiO3-0.06BaTiO3 to analyze the effect on Td and reported an enhanced pyroelectric coefficient of 105 × 10−4 C m−2 K−1 for the nonstoichiometric sample at room temperature [30]. Further, Li et al. used the composite approach with nonstoichiometric (1-x)0.94Na0.48Bi0.44TiO3-0.06BaTiO3 and ZnO to enhance the pyroelectric properties and Td [31].
In order to assess the performance of pyroelectric materials, a variety of FOMs have been developed for different applications [32,33,34]. These include FOMs for current responsivity ( F i ), detectivity ( F d ), voltage responsivity ( F v ), the pyroelectric figure of merit ( F c ) and energy harvesting ( F e and F e * ). Therefore, to derive the best performance out of pyroelectric materials, not only a higher pyroelectric coefficient (p) but also small relative dielectric permittivity ( ε r ) and loss ( t a n δ ) are desirable [6]. Therefore, an optimum tradeoff between p, ε r , t a n δ and specific heat capacity (C) is required to enhance FOMs. In addition to this, materials should demonstrate temperature-independent pyroelectric performance. Hence, several methods are used to improve the FOMs, such as composite fabrication, porous ceramics and chemical modification [2,35,36,37]. However, researchers continue to explore new materials for efficient usage. In the present work, Zn-doped NBT-6BT and NBT-9BT ceramics are explored to characterize the pyroelectric performance. Hence, a comprehensive set of dielectric, pyroelectric and piezoelectric characterization is undertaken, along with the estimation of relevant FOMs. This work suggests an avenue to design pyroelectric devices with temperature-independent performance in a wide temperature range, possibly substituting lead-based ceramics.

2. Materials and Methods

0.94Na1/2Bi1/2TiO3-0.06BaTiO3 (NBT-6BT) and 0.91Na1/2Bi1/2TiO3-0.09BaTiO3 powders doped with xZnO (x = 0, 0.005, 0.01; x denotes number of moles) were prepared by a solid-state synthesis route. The raw powders of Bi2O3 (99.975%), BaCO3 (99.8%), Na2CO3 (99.5%), TiO2 (99.6%) and ZnO (99.99%) (all Alfa Aesar) were weighed according to their stoichiometric ratio. Using a planetary ball mill, the raw materials were ball-milled for 12 h in ethanol at 250 rpm (Fritsch Pulverisette 5). The resultant slurries were dried and calcined at 900 °C for 3 h in closed alumina crucibles with a heating rate of 5 K min−1. The calcined powders were remilled again for 6 h in ethanol at 250 rpm. Sintering was done at 1100 °C for 1 h at a heating rate of 5 K min−1. In addition, to avoid volatilization of Na+ and Bi3+ during the sintering process, samples were embedded in a powder bed of the same composition. The relative density of the samples was measured using the Archimedes method. A polished and thermally etched (100–150 °C below the sintering temperature for 15 min) sample was used for scanning electron microscopy (SEM) (Philips XL30 FEG). For the electrical characterization, samples were ground to a thickness of 0.6 mm and sputtered with silver electrodes. Samples were poled in silicone oil by applying a DC field 6 kV mm−1 at room temperature for 15 min and the piezoelectric, dielectric and pyroelectric measurements were performed after 24 h. The piezoelectric constant (d33) was measured using a d33 meter (PM 300; Piezo Test) under a load of 2 N at room temperature. An HP analyzer was used to measure the temperature and frequency-dependent permittivity using a programmable furnace (ramp rate of 2 K min−1). The pyroelectric current (I) was measured in a short circuit mode, utilizing a constant heating rate of 3 K min−1, while the discharge currents were monitored by an electrometer (Model 617; Keithley). The measurements were performed in two different methods. The samples were measured in one complete heating (above Td) and cooling cycle in one set of measurements. In another set of experiments, the samples were subjected to temperature cycling (heating at 3 K min−1 and natural cooling) up to ~20 K below the Td (10 cycles). The pyroelectric current was measured in the whole temperature range (above Td) after the thermal cycling. The measured I is used for the estimation of the pyroelectric coefficient (p) and the FOMs.

3. Results and Discussion

Phase purity analysis using XRD has been previously reported by Lalitha et al. [29] and samples from the same batch have been used in the present work. Hence, XRD data are not presented here. The X-ray diffraction pattern of NBT-yBT-xZn (y = 6, 9 and x = 0.005 and 0.01) ceramics correspond to the perovskite pattern. Secondary phases were not detected within the sensitivity of XRD, following previous reports [29]. This confirms the complete incorporation of ZnO into the NBT-6BT and NBT-9BT host lattice.
Figure 1 features the surface morphology of the NBT-yBT-xZn (y = 6, 9 and x = 0.005 and 0.01) ceramics and the insets present the corresponding grain size distribution. The micrographs feature a dense microstructure and the grain size increases with increasing ZnO content. Moreover, compositional segregation is not observed in the SEM images. Figure 2a shows the relative density as a function of ZnO content, which is in the range of 93–98% for NBT-yBT-xZn (y = 6, 9 and x = 0.005 and 0.01) ceramics. The relative density for 0.005 mole ZnO added NBT-6BT, i.e., 93% increases to 98% for 0.01 mole of ZnO. Further, the average grain size as a function of ZnO content was estimated and is shown in Figure 2b. It is observed that with increasing ZnO content, the average grain size increases from ~2–11.4 µm and 1.4–9.3 µm for NBT-6BT and NBT-9BT, respectively.
The depolarization temperature (Td) obtained from the depolarization current density (IA) for NBT-6BT and NBT-9BT is shown in Figure 3a,b. Td is the temperature at which the macroscopic polarization of the material disappears and can be observed with a sharp increment in the temperature-induced IA. NBT-6BT and NBT-9BT exhibit Td of 377 K and 442 K, respectively. With Zn-doping in NBT-6BT and NBT-9BT, Td increases to 415 K, 425 K and 472 K, 478 K, respectively. The graphical variation of Td with ZnO content is shown in Figure 3c.
This increase in the depolarization temperature of NBT-yBT-xZn (y = 6, 9 and x = 0, 0.005, 0.01) has been attributed to the large ionic polarizability of the Zn2+ [38] and onset of tetragonal distortion, stabilizing the ferroelectric domain structure [29]. Additionally, the increase in the grain size could also stabilize the ferroelectric state, leading to higher Td [39]. Piezoelectric constant (d33) for NBT-6BT is observed to decrease from 137 to 124 pC/N and 114 pC/N with increasing Zn concentration to 0.005 and 0.01, respectively, as seen in Figure 3c. A similar decrease in d33 is observed for NBT-9BT, where it decreased by 37% and 32%, with increasing ZnO content. The decrease in d33 may be due to the hardening caused by Zn2+ doping at the B site of NBT-yBT [29]. Figure 3d depicts the total polarization of NBT-6BT and NBT-9BT, which decreases with the increase in Zn concentration. Further, d33 and polarization also decreases with an increase in grain size, due to acceptor (Zn)-doping effects. Hence, the minimum d33 and polarization was observed for 0.01 mole ZnO doped NBT-6BT and NBT-9BT, featuring larger grain size. A comprehensive discussion on the effect of Zn-doping on the polarization and piezoelectric properties is discussed by Lalitha et al. [29].
The pyroelectric coefficient can be measured by various methods, including direct or indirect methods. However, in the present study, a direct measurement of a pyroelectric coefficient is used, owing to its higher accuracy. The pyroelectric coefficient is estimated by the Byer–Roundy technique based on the following equation [40]:
I p = A × p × d T d t
where Ip is the depolarization current, A is the area of the electrodes, p is the pyroelectric coefficient and d T / d t is the heating or the cooling rate. Figure 4a,b shows the depolarization current density (i.e., depolarization current per unit area of electrodes, IA) after the 1st cycle (i.e., below the Td) and after the 10th complete heating and cooling cycle (above Td), for NBT-6BT and NBT-9BT, respectively. This has been ascertained to minimize the effect of trapping and de-trapping charge arising from heating. These trapped charges may give an erroneous value to the depolarization current and hence, affecting the pyroelectric constant. Thus, the combined effect of the multiple heating and natural cooling cycles (i.e., ten in the present study) reduces any charge trapping effect. In addition, in practical application, multiple heating–cooling cycles are employed. It is important to note that, to avoid thermal depolarization, the samples were heated up to 20 K below Td during multiple heating cycles. Finally, after the tenth cycle, the samples were heated above the Td. The variations in IA for the first and after the tenth cycle for NBT-yBT-xZn (y = 6, 9 and x = 0, 0.5, 1) are shown in Figure 4a,b.
The pyroelectric coefficient p, calculated using Equation (1) for NBT-6BT-xZn and NBT-9BT-xZn, is shown in Figure 4c,d, respectively. p at room temperature for NBT-6BT and NBT-9BT is calculated to be 3.40 × 10−4 C m−2 K−1 and 1.72 × 10−4 C m−2 K−1, respectively. However, doping with ZnO in NBT-6BT and NBT-9BT did not significantly alter the p at room temperature. The maximum p is noted for NBT-6BT and NBT-9BT as 974.8 × 10−4 C m−2 K−1 (at 378 K) and 605 × 10−4 C m−2 K−1 (at 432 K), respectively. Moreover, the maximum p could not be increased by Zn-doping, but a temperature-independent behavior of p is observed. A temperature-independent range of 300–330 K and 300–380 K for p is noted for NBT-6BT and NBT-9BT, respectively. These values are increased to 370 K and 340 K for NBT-6BT and 380 K and 395 K for NBT-9BT with 0.005 and 0.01 moles of ZnO addition. Hence, NBT-yBT-xZn (y = 6, 9 and x = 0, 0.005, 0.01) with higher Td can be used in devices for electronic sensing or energy harvesting, as most of the devices heat up during their operation nearly in the range of 333 K–363 K. In addition, the increase in Td is highly relevant for energy harvesting applications, where a larger temperature difference is demanded. However, these compositions can also be useful to study the temperature-independent FOM. Thus, to understand the FOMs, it is essential to analyze the dielectric constant and loss, as the FOMs are largely dependent on the p and dielectric constant and loss.
The dielectric properties associated with the compositions can be observed by the variation in the dielectric permittivity ( ε r ) and loss ( t a n δ ) as a function of temperature as well as frequency is shown in Figure 5. The details of non-ergodic relaxor state (doped and undoped NBT-6BT), spontaneous ferroelectric order at room temperature (doped NBT-9BT) and ergodic state at high temperature have been discussed in detail in previous work [29]. Figure 5 shows that two dielectric anomalies are present in the poled state, owing to different transition processes. The first anomaly in the graph of dielectric constant is the ferroelectric to relaxor transition temperature (TF-R) and the second anomaly Tm shows the transition from relaxor to paraelectric. TF-R is comparable to Td with minor changes (~2–3 K) noted by some other research groups, in the absence of any other external stimuli (strong electric field or stress). Figure 5a shows the TF-R to be 376 K for NBT-6BT. An increased TF-R, i.e., 413 K and 423 K with 0.005 and 0.01 moles of ZnO, is obtained in NBT-6BT (Figure 5b,c). A similar trend is seen for NBT-9BT, where TF-R is observed to be 445 K, which increases to 470 K with Zn-doping, as seen in Figure 5d–f.
The increment in TF-R with increasing Zn concentration is attributed to stabilizing the ferroelectric order at a higher temperature [29]. ε r at room temperature for NBT-6BT and NBT-9BT is 678 and 912 at 1 kHz. However, with the addition of 0.005 and 0.01 mole of ZnO, ε r decreases to 561, 468 for NBT-6BT and 691 and 520 for NBT-9BT at 1 kHz, respectively. A decrease in t a n δ is observed with increasing Zn concentration at room temperature. This decrease in dielectric constant and loss is due to the acceptor-doping effect of Zn in NBT-yBT [29]. Furthermore, the dielectric constant and loss at room temperature decreases with increase in grain size. The lower dielectric constant can be a result of lower domain wall density (corresponding to larger grain size). The dielectric constant at TF-R is decreased by 23% (NBT-9BT-0.005Zn) and 45% (NBT-9BT-0.01Zn) compared to NBT-9BT; however, for NBT-6BT, it remains almost constant. This decrease in the dielectric constant at room temperature and increase in TF-R can be beneficial to enhance the FOM. However, ε r and t a n δ for Zn-doped materials increases rapidly after TF-R in the low frequency region. This rapid increase is due to enhanced conductivity in the doped compositions [29].
In order to further evaluate the pyroelectric performance, various pyroelectric FOMs are estimated. Therefore, p, ε r and t a n δ are used to determine the various FOMs presented in Figure 6 and Figure 7 for NBT-6BT-xZn and NBT-9BT-xZn (x = 0.005 and 0.01), respectively for the first cycle of heating. The corresponding FOMs after the tenth heating cycle are shown in Figure A1 and Figure A2 (Appendix A). For high current responsivity, the relevant FOM is F i = p c v , where c v is volume-specific heat [41]. Temperature-dependent specific heat capacity is taken from literature and c v is estimated as c v = C . ρ , where ρ is the density of the material [42]. The sample densities were measured using the Archimedes principle and the obtained values are used for calculations. The pyroelectric FOM is expressed as ( F C ) = p ϵ r . Energy harvesting FOMs are ( F e ) = p 2 ε r ε 0 and ( F e * ) = p 2 ε r ε 0 c v 2 [4]. High detectivity based FOM is ( F d ) = p c v ε r ε 0 t a n δ . Moreover, for the high voltage responsivity, the FOM is given as ( F v ) = p c v   ε r ε 0 , where ε 0 is the permittivity of free space [4]. Figure 6 and Figure 7 show the calculated FOMs in the temperature range of 300 K–400 K and 300 K–440 K, respectively. Figure 6a and Figure 7a show that, at room temperature, F i increases with increasing ZnO content, which is found as 1.07, 1.41, 1.44 × 10−10 C m J−1 and 0.81, 0.73 and 1.07 × 10−10 C m J−1 for NBT-6BT-xZn and NBT-9BT-xZn (x = 0, 0.005 and 0.01).
Similarly, at room temperature, F v increases with increasing ZnO content, which is found as 17.9, 28.20 and 34.5 × 10−3 m2 C−1 and 10, 13.6 and 23 × 10−3 m2 C−1 for NBT-6BT-xZn and NBT-9BT-xZn (x = 0, 0.005 and 0.01), respectively, as seen in Figure 6b and Figure 7b. A similar trend is observed for pyroelectric energy harvesting FOMs ( F e * ) and ( F e ) as seen in Figure 6c,d and Figure 7c,d.
The maximum value of FOMs ( F e * ) and ( F e ) obtained are 1033 × 10−12 m3 J−1 and 281 J m−3 K−2 for NBT-9BT-0.01Zn, 1363 × 10−12 m3 J−1 and 753 J m−3 K−2 for NBT-6BT-0.01Zn, which are much higher than that of undoped NBT-6BT and NBT-9BT, respectively. Similar behavior is observed for F d and F v , where the Zn-doping increases the FOM at room temperature, as seen in Figure 6e,f and Figure 7e,f. Zn-doping results in an increase in the temperature-independent behavior from 303–350 K for NBT-6BT to 303–380 K for NBT-6BT-0.001Zn. However, the FOMs after the tenth heating cycle showed a better temperature-independent behavior for Zn-doped NBT-6BT and NBT-9BT (Figure A1 and Figure A2). Table 1 presents the FOMs at Td for different NBT-yBT-xZn (y = 6, 9 and x = 0, 0.005, 0.01) after the tenth heating cycle. Most studies focus on enhancing the pyroelectric properties at Td, while lowering it to near room temperature. However, for practical applications, wide temperature-independent properties are needed. These results show that NBT-yBT-xZn can be a promising candidate for pyroelectric applications in a wide temperature range.

4. Conclusions

In the present work, pyroelectric properties and relevant FOMs for Na1/2Bi1/2TiO3-yBaTiO3-xZn ceramics (y = 6, 9 and x = 0, 0.005, 0.01) were investigated. Zn-doping increases the depolarization temperature by almost ~40 K compared to pure NBT-6BT and NBT-9BT. Both relative permittivity and dielectric loss decrease with increasing ZnO content which is beneficial for the pyroelectric applications. Zn-doping increases the range of temperature-independent pyroelectric coefficient (p) behavior compared to undoped Na1/2Bi1/2TiO3-yBaTiO3 (y = 6, 9) and features higher p in comparison to other lead-free and PZT ceramics. Pyroelectric detector FOMs (Fd) are enhanced with Zn doping to 8.16, 21.8 and 26 × 106 Pa−1/2 for x = 0, 0.005 and 0.01 in NBT-6BT-xZn, respectively, at room temperature. These results demonstrate that NBT-6BT-xZn (x = 0.01) is a promising material for infrared detectors and other pyroelectric applications in a wide operating temperature range of ~80 K (300–380 K).

Author Contributions

Conceptualization, investigation and methodology—K.V.L. and S.P.; Experimentation and data acquisition—K.V.L.; Analysis and interpretation—S.P.; writing-original draft preparation—S.P. and N.S.; writing-review and editing—K.V.L. All authors have read and agreed to the published version of the manuscript.

Funding

K.V.L. acknowledges support by the Deutsche Forschungsgemeinschaft (DFG—German Research Foundation) and the Open Access Publishing Fund of Technical University of Darmstadt. K.V.L. acknowledges and thanks the Alexander von Humboldt Foundation and the Deutsche Forschungsgemeinschaft under grant no. KO 5948/1-1 (No. 414311761) for financial support. S.P. thanks Science and Engineering Research Board (SERB) for financial support in the frame of the Start-up Research Grant no. SRG/2020/000188.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available within this article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Appendix A

Figure A1. Pyroelectric figures of merit versus temperature plots after tenth heating cycle for NBT-6BT-xZn (x = 0.5 and 1%): (a) current responsivity ( F i ), (b) the pyroelectric figure of merit ( F c ), energy harvesting (c) F e * , and (d) F e . (e) Detectivity ( F d ) and (f) the voltage responsivity ( F v ).
Figure A1. Pyroelectric figures of merit versus temperature plots after tenth heating cycle for NBT-6BT-xZn (x = 0.5 and 1%): (a) current responsivity ( F i ), (b) the pyroelectric figure of merit ( F c ), energy harvesting (c) F e * , and (d) F e . (e) Detectivity ( F d ) and (f) the voltage responsivity ( F v ).
Materials 15 00087 g0a1
Figure A2. Pyroelectric figures of merit versus temperature plots after tenth heating cycle of NBT-9BT-xZn (x = 0.5 and 1%): (a) current responsivity ( F i ), (b) the pyroelectric figure of merit ( F c ), energy harvesting (c) F e * , and (d) F e . (e) Detectivity ( F d ) and (f) the voltage responsivity ( F v ).
Figure A2. Pyroelectric figures of merit versus temperature plots after tenth heating cycle of NBT-9BT-xZn (x = 0.5 and 1%): (a) current responsivity ( F i ), (b) the pyroelectric figure of merit ( F c ), energy harvesting (c) F e * , and (d) F e . (e) Detectivity ( F d ) and (f) the voltage responsivity ( F v ).
Materials 15 00087 g0a2

References

  1. Zhang, D.; Wu, H.; Bowen, C.R.; Yang, Y. Recent Advances in Pyroelectric Materials and Applications. Small 2021, 2103960. [Google Scholar] [CrossRef] [PubMed]
  2. He, H.; Lu, X.; Hanc, E.; Chen, C.; Zhang, H.; Lu, L. Advances in lead-free pyroelectric materials: A comprehensive review. J. Mater. Chem. C 2020, 8, 1494–1516. [Google Scholar] [CrossRef]
  3. Moisello, E.; Malcovati, P.; Bonizzoni, E. Thermal Sensors for Contactless Temperature Measurements, Occupancy Detection, and Automatic Operation of Appliances during the COVID-19 Pandemic: A Review. Micromachines 2021, 12, 148. [Google Scholar] [CrossRef]
  4. Bowen, C.R.; Kim, H.A.; Weaver, P.M.; Dunn, S. Piezoelectric and ferroelectric materials and structures for energy harvesting applications. Energy Environ. Sci. 2014, 7, 25–44. [Google Scholar] [CrossRef] [Green Version]
  5. Chamankar, N.; Khajavi, R.; Yousefi, A.A.; Rashidi, A.S.; Golestanifard, F. Comparing the piezo, pyro and dielectric properties of PZT particles synthesized by sol–gel and electrospinning methods. J. Mater. Sci. Mater. Electron. 2019, 30, 8721–8735. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Xie, M.; Roscow, J.; Bao, Y.; Zhou, K.; Zhang, D.; Bowen, C.R. Enhanced pyroelectric and piezoelectric properties of PZT with aligned porosity for energy harvesting applications. J. Mater. Chem. A 2017, 5, 6569–6580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Kumar, P.; Sharma, S.; Thakur, O.; Prakash, C.; Goel, T. Dielectric, piezoelectric and pyroelectric properties of PMN–PT (68:32) system. Ceram. Int. 2004, 30, 585–589. [Google Scholar] [CrossRef]
  8. Lau, S.T.; Cheng, C.H.; Choy, S.H.; Lin, D.; Kwok, K.; Chan, H.L.W. Lead-free ceramics for pyroelectric applications. J. Appl. Phys. 2008, 103, 104105. [Google Scholar] [CrossRef] [Green Version]
  9. Liu, Z.; Lu, T.; Dong, X.; Wang, G.; Liu, Y. Ferroelectric ceramics for pyroelectric detection applications: A review. IEEE Trans Ultrason. Ferroelectr. Freq. Control 2020, 68, 242–252. [Google Scholar] [CrossRef]
  10. Guo, F.; Yang, B.; Zhang, S.; Wu, F.; Liu, D.; Hu, P.; Sun, Y.; Wang, D.; Cao, W. Enhanced pyroelectric property in (1−x)(Bi0.5Na0.5)TiO3-xBa(Zr0.055Ti0.945)O3: Role of morphotropic phase boundary and ferroelectric-antiferroelectric phase transition. Appl. Phys. Lett. 2013, 103, 182906. [Google Scholar] [CrossRef] [Green Version]
  11. Li, J.F.; Wang, K.; Zhu, F.Y.; Cheng, L.Q.; Yao, F.Z. (K, Na) NbO3-Based Lead-Free Piezoceramics: Fundamental Aspects, Processing Technologies, and Remaining Challenges. J. Am. Ceram. Soc. 2013, 96, 3677–3696. [Google Scholar] [CrossRef]
  12. Xu, X.; Jiang, J.; Ahmad, H.; Yuan, G.; Yin, J.; Liu, Z. Improved ferroelectricity of (1−x)Na0.5Bi0.5TiO3–xBaTiO3 ceramics rapidly sintered at low temperature. Ceram. Int. 2014, 40, 11819–11824. [Google Scholar] [CrossRef]
  13. Swain, S.; Kar, S.K.; Kumar, P. Dielectric, optical, piezoelectric and ferroelectric studies of NBT–BT ceramics near MPB. Ceram. Int. 2015, 41, 10710–10717. [Google Scholar] [CrossRef]
  14. Martin, A.; Khansur, N.H.; Webber, K.G. Electric field-induced changes in the ferroelastic behavior of (Na1/2Bi1/2)TiO3-BaTiO3. J. Eur. Ceram. Soc. 2018, 38, 4623–4630. [Google Scholar] [CrossRef]
  15. Guennou, M.; Savinov, M.; Drahokoupil, J.; Luo, H.; Hlinka, J. Piezoelectric properties of tetragonal single-domain Mn-doped NBT-6 %BT single crystals. Appl. Phys. A 2014, 116, 225–228. [Google Scholar] [CrossRef] [Green Version]
  16. Lalitha, K.V.; Koruza, J.; Rödel, J. Propensity for spontaneous relaxor-ferroelectric transition in quenched (Na1/2Bi1/2)TiO3-BaTiO3 compositions. Appl. Phys. Lett. 2018, 113, 252902. [Google Scholar]
  17. Samantaray, K.S.; Amin, R.; Rini, E.G.; Bhaumik, I.; Mekki, A.; Harrabi, K.; Sen, S. Defect Dipole Induced Improved Electrocaloric Effect in Modified NBT-6BT Lead-Free Ceramics. arXiv 2021, arXiv:2110.04262. [Google Scholar]
  18. Turki, O.; Slimani, A.; Seveyrat, L.; Sassi, Z.; Khemakhem, H.; Lebrun, L. Enhancement of dielectric, piezoelectric, ferroelectric, and electrocaloric properties in slightly doped (Na0.5Bi0.5)0.94Ba0.06TiO3 ceramic by samarium. J. Appl. Phys. 2019, 125, 174103. [Google Scholar] [CrossRef]
  19. Uddin, S.; Zheng, G.P.; Iqbal, Y.; Ubic, R.; Yang, J. Unification of the negative electrocaloric effect in Bi1/2Na1/2TiO3-BaTiO3 solid solutions by Ba1/2Sr1/2TiO3 doping. J. Appl. Phys. 2013, 114, 213519. [Google Scholar] [CrossRef] [Green Version]
  20. Cao, W.; Li, W.; Xu, D.; Hou, Y.; Wang, W.; Fei, W. Enhanced electrocaloric effect in lead-free NBT-based ceramics. Ceram. Int. 2014, 40, 9273–9278. [Google Scholar] [CrossRef]
  21. Takenaka, T.; Maruyama, K.-I.; Sakata, K. (Bi1/2Na1/2)TiO3-BaTiO3 system for lead-free piezoelectric ceramics. Jpn. J. Appl. Phys. 1991, 30, 2236. [Google Scholar] [CrossRef]
  22. Martinez, F.L.; Hinojosa, J.; Domenech, G.; Fernandez-Luque, F.J.; Zapata, J.; Ruiz, R.; Pardo, L. Dielectric constant tunability at microwave frequencies and pyroelectric behavior of lead-free submicrometer-structured (Bi0.5Na0.5)1-xBaTiO3 ferroelectric ceramics. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2013, 60, 1595–1602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Abe, J.; Kobune, M.; Nishimura, T.; Yazaw, T.; Nakai, Y. Effects of Manganese Addition on Pyroelectric Properties of (Bi0.5Na0.5TiO3)0.94(BaTiO3)0.06 Ceramics. Integr. Ferroelectr. 2006, 80, 87–95. [Google Scholar] [CrossRef]
  24. Balakt, A.; Shaw, C.; Zhang, Q. Large pyroelectric properties at reduced depolarization temperature in A-site nonstoichiometry composition of lead-free 0.94NaxBiyTiO3–0.06BazTiO3 ceramics. J. Mater. Sci. 2017, 52, 7382–7393. [Google Scholar] [CrossRef] [Green Version]
  25. Balakt, A.; Shaw, C.; Zhang, Q. The effects of Ba2+ content on depolarization temperature and pyroelectric properties of lead-free 0.94Na0.5Bi0.5TiO3–0.06Ba1+xTiO3 ceramics. J. Mater. Sci. Mater. Electron. 2016, 27, 12947–12954. [Google Scholar] [CrossRef] [Green Version]
  26. Balakt, A.; Shaw, C.; Zhang, Q. Enhancement of pyroelectric properties of lead-free 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics by La doping. J. Eur. Ceram. Soc. 2017, 37, 1459–1466. [Google Scholar] [CrossRef] [Green Version]
  27. Balakt, A.; Shaw, C.; Zhang, Q. The decrease of depolarization temperature and the improvement of pyroelectric properties by doping Ta in lead-free 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics. Ceram. Int. 2017, 43, 3726–3733. [Google Scholar] [CrossRef]
  28. Balakt, A.; Shaw, C.; Zhang, Q. Giant pyroelectric properties in La and Ta co-doped lead-free 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics. J. Alloys Compd. 2017, 709, 82–91. [Google Scholar] [CrossRef] [Green Version]
  29. Kodumudi Venkataraman, L.; Zhu, T.; Pinto Salazar, M.; Hofmann, K.; Iqbal Waidha, A.; Jaud, J.C.; Rödel, J. Thermal depolarization and electromechanical hardening in Zn2+-doped Na1/2Bi1/2TiO3-BaTiO3. J. Am. Ceram. Soc. 2021, 104, 2201–2212. [Google Scholar] [CrossRef]
  30. Saurabh, N.; Patel, S. Nonstoichiometric effect on electrocaloric, pyroelectric and energy storage properties of 0.94NaxBiyTiO3–0.06BaTiO3 bulk ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 26871–26893. [Google Scholar] [CrossRef]
  31. Li, L.; Wang, R.X.; Shen, M.; Zhang, H.; Gu, Z.B.; Zhang, S.T.; Wu, D. Phase/domain structure and enhanced thermal stable ferro-/pyroelectric properties of (1−x)0.94Na0.48Bi0.44TiO3-0.06BaTiO3: xZnO ceramics. J. Eur. Ceram. Soc. 2020, 40, 699–705. [Google Scholar]
  32. Bowen, C.; Taylor, J.; Le Boulbar, E.; Zabek, D.; Topolov, V. A modified figure of merit for pyroelectric energy harvesting. Mater. Lett. 2015, 138, 243–246. [Google Scholar] [CrossRef] [Green Version]
  33. Mangalam, R.V.K.; Agar, J.C.; Damodaran, A.R.; Karthik, J.; Martin, L.W.; Mangalam, V.K. Improved Pyroelectric Figures of Merit in Compositionally Graded PbZr1−xTixO3 Thin Films. ACS Appl. Mater. Interfaces 2013, 5, 13235–13241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Zhang, Y.; Roscow, J.; Xie, M.; Bowen, C. High piezoelectric sensitivity and hydrostatic figures of merit in unidirectional porous ferroelectric ceramics fabricated by freeze casting. J. Eur. Ceram. Soc. 2018, 38, 4203–4211. [Google Scholar] [CrossRef] [Green Version]
  35. Bowen, C.R.; Taylor, J.; LeBoulbar, E.; Zabek, D.; Chauhan, A.; Vaish, R. Pyroelectric materials and devices for energy harvesting applications. Energy Environ. Sci. 2014, 7, 3836–3856. [Google Scholar] [CrossRef] [Green Version]
  36. Sharma, M.; Singh, V.P.; Singh, S.; Azad, P.; Ilahi, B.; Madhar, N.A. Porous Ba0.85Ca0.15Zr0.1Ti0.9O3 Ceramics for Pyroelectric Applications. J. Electron. Mater. 2018, 47, 4882–4891. [Google Scholar] [CrossRef]
  37. Srikanth, K.; Singh, V.; Vaish, R. Enhanced pyroelectric figure of merits of porous BaSn0.05Ti0.95O3 ceramics. J. Eur. Ceram. Soc. 2017, 37, 3943–3950. [Google Scholar] [CrossRef]
  38. Li, L.; Zhu, M.; Zhou, K.; Wei, Q.; Zheng, M.; Hou, Y. Delayed thermal depolarization of Bi0.5Na0.5TiO3-BaTiO3 by doping acceptor Zn2+ with large ionic polarizability. J. Appl. Phys. 2017, 122, 204104. [Google Scholar] [CrossRef]
  39. Khatua, D.K.; Mishra, A.; Kumar, N.; Das Adhikary, G.; Shankar, U.; Majumdar, B.; Ranjan, R. A coupled microstructural-structural mechanism governing thermal depolarization delay in Na0.5Bi0.5TiO3-based piezoelectrics. Acta Mater. 2019, 179, 49–60. [Google Scholar] [CrossRef]
  40. Pintilie, L.; Iuga, A.; Stancu, V.; Botea, M. Estimation of the pyroelectric coefficient for ceramic Pb(Zr, Ti)O3 samples: Comparison of the results obtained by two easy to implement methods. Infrared Phys. Technol. 2020, 106, 103269. [Google Scholar] [CrossRef]
  41. Patel, S.; Chauhan, A.; Vaish, R. Large pyroelectric figure of merits for Sr-modified Ba0.85Ca0.15Zr0.1Ti0.9O3 ceramics. Solid State Sci. 2016, 52, 10–18. [Google Scholar] [CrossRef]
  42. Lewczuk-Jodłowiec, U.; Suchanicz, J.; Sitko, D.; Konieczny, K. Thermal properties of NBT-BT systems. Phase Transit. 2017, 90, 818–823. [Google Scholar] [CrossRef]
  43. Mahdi, R.; Al-Bahnam, N.; Abbo, A.; Hmood, J.; Majid, W.A. Optimization of sintering temperature for the enhancement of pyroelectric properties of lead-free 0.88(Na0.5Bi0.5)TiO3–0.084(K0.5Bi0.5)TiO3–0.036BaTiO3 piezoelectric ceramics. J. Alloys Compd. 2016, 688, 77–87. [Google Scholar] [CrossRef]
  44. Fang, B.; Wang, M.; Yuan, N.; Ding, J.; Zhao, X.; Xu, H.; Luo, H. Structural phase transition, optical and pyroelectric properties of lead-free single crystals. Chin. Sci. Bull. 2013, 58, 4064–4071. [Google Scholar] [CrossRef] [Green Version]
  45. Zhang, J.; Dong, X.; Cao, F.; Guo, S.; Wang, G. Enhanced pyroelectric properties of Cax(Sr0.5Ba0.5)1−xNb2O6 lead-free ceramics. Appl. Phys. Lett. 2013, 102, 102908. [Google Scholar] [CrossRef]
  46. Srikanth, K.S.; Singh, V.P.; Vaish, R. Pyroelectric performance of porous Ba0.85Sr0.15TiO3 ceramics. Int. J. Appl. Ceram. Technol. 2017, 15, 140–147. [Google Scholar] [CrossRef]
Figure 1. Scanning electron micrographs of thermally etched and polished surface of (a) NBT-6BT, (b) NBT-6BT-0.005Zn, (c) NBT-6BT-0.01Zn, (d) NBT-9BT, (e) NBT-9BT-0.005Zn and (f) NBT-9BT-0.01Zn. The inset shows the corresponding grain size distribution.
Figure 1. Scanning electron micrographs of thermally etched and polished surface of (a) NBT-6BT, (b) NBT-6BT-0.005Zn, (c) NBT-6BT-0.01Zn, (d) NBT-9BT, (e) NBT-9BT-0.005Zn and (f) NBT-9BT-0.01Zn. The inset shows the corresponding grain size distribution.
Materials 15 00087 g001
Figure 2. (a) Relative density and (b) average grain size as a function of Zn content for NBT-yBT-xZn (y = 6, 9 and x = 0, 0.005, 0.01) ceramics.
Figure 2. (a) Relative density and (b) average grain size as a function of Zn content for NBT-yBT-xZn (y = 6, 9 and x = 0, 0.005, 0.01) ceramics.
Materials 15 00087 g002
Figure 3. Depolarization current density after one complete cycle as a function of temperature for (a) NBT-6BT-xZn (x = 0, 0.005, 0.01) and (b) NBT-9BT-xZn (x = 0, 0.005, 0.01). The vertical green lines denote the Td. (c) Depolarization temperature (Td) and piezoelectric coefficient (d33) of NBT-yBT (y = 6, 9) as a function of ZnO content. (d) Total polarization of NBT-yBT (y = 6, 9) as a function of ZnO content.
Figure 3. Depolarization current density after one complete cycle as a function of temperature for (a) NBT-6BT-xZn (x = 0, 0.005, 0.01) and (b) NBT-9BT-xZn (x = 0, 0.005, 0.01). The vertical green lines denote the Td. (c) Depolarization temperature (Td) and piezoelectric coefficient (d33) of NBT-yBT (y = 6, 9) as a function of ZnO content. (d) Total polarization of NBT-yBT (y = 6, 9) as a function of ZnO content.
Materials 15 00087 g003
Figure 4. Depolarization current density and pyroelectric coefficient (p) for 1st and after 10th cycle as a function of temperature for (a), (c) NBT-6BT-xZn (x = 0, 0.005, 0.01) and (b), (d) NBT-9BT-xZn (x = 0, 0.005, 0.01), respectively.
Figure 4. Depolarization current density and pyroelectric coefficient (p) for 1st and after 10th cycle as a function of temperature for (a), (c) NBT-6BT-xZn (x = 0, 0.005, 0.01) and (b), (d) NBT-9BT-xZn (x = 0, 0.005, 0.01), respectively.
Materials 15 00087 g004
Figure 5. Dielectric permittivity ( ε r ) and loss ( t a n δ ) of poled samples as a function of temperature at various frequencies for (a) NBT-6BT, (b) NBT-6BT-0.005Zn, (c) NBT-6BT-0.01Zn, (d) NBT-9BT, (e) NBT-9BT-0.005Zn, and (f) NBT-9BT-0.01Zn.
Figure 5. Dielectric permittivity ( ε r ) and loss ( t a n δ ) of poled samples as a function of temperature at various frequencies for (a) NBT-6BT, (b) NBT-6BT-0.005Zn, (c) NBT-6BT-0.01Zn, (d) NBT-9BT, (e) NBT-9BT-0.005Zn, and (f) NBT-9BT-0.01Zn.
Materials 15 00087 g005
Figure 6. Pyroelectric figures of merit versus temperature plots of NBT-6BT-xZn (x = 0.005 and 0.01) for (a) current responsivity F i , (b) the pyroelectric figure of merit F c , energy harvesting (c) F e * and (d) F e . (e) detectivity F d and (f) voltage responsivity F v .
Figure 6. Pyroelectric figures of merit versus temperature plots of NBT-6BT-xZn (x = 0.005 and 0.01) for (a) current responsivity F i , (b) the pyroelectric figure of merit F c , energy harvesting (c) F e * and (d) F e . (e) detectivity F d and (f) voltage responsivity F v .
Materials 15 00087 g006
Figure 7. Pyroelectric figures of merit versus temperature plots of NBT-9BT-xZn (x = 0.005 and 0.01) for high (a) current responsivity F i , (b) the pyroelectric figure of merit F c , energy harvesting (c) F e * and (d) F e , (e) detectivity F d and (f) the voltage responsivity F v .
Figure 7. Pyroelectric figures of merit versus temperature plots of NBT-9BT-xZn (x = 0.005 and 0.01) for high (a) current responsivity F i , (b) the pyroelectric figure of merit F c , energy harvesting (c) F e * and (d) F e , (e) detectivity F d and (f) the voltage responsivity F v .
Materials 15 00087 g007
Table 1. Comparison of the pyroelectric coefficient and FOMs (after the tenth cycle of heating) at room temperature and depolarization temperature (Td) for some important bulk materials.
Table 1. Comparison of the pyroelectric coefficient and FOMs (after the tenth cycle of heating) at room temperature and depolarization temperature (Td) for some important bulk materials.
Compositionsp × 10−4 (C m−2 K−1)Fi × 10−10 (C m J−1)Fv × 10−3 (m2 C−1)Fd × 10−6 (Pa−1/2)Fc × 10−4 (C m−2K−1)Ref.
at room temperature
NBT-6BT3.401.0717.98.160.11Present work
NBT-6BT-0.005Zn2.181.4128.2021.80.16Present work
NBT-6BT-0.01Zn2.431.4434.526.30.18Present work
NBT-9BT1.720.81105.060.08Present work
NBT-9BT-0.005Zn1.710.8313.611.30.09Present work
NBT-9BT-0.01Zn1.511.072317.40.139Present work
NBT-6BT3.141.120.0219.081.58[24,26,27,28]
NBT-6Ba1.02T3.541.2440.958.3 [25]
N0.52B0.52T-6BT6.992.500.04716.632.86[24]
NBT-6BT-0.002Ta7.142.550.0331.292.42[27]
NBT-6BT-0.005La7.422.650.0481.42.98[26]
0.005La-NBT-6BT-0.002Ta12.924.610.0782.762.57[28]
BNT-BKT-BT3.662.150.02615.41 [43]
0.95NBT-0.05BT- Mn-doped *8.341.110.0168.640.29[44]
BNKBT3.251.950.02613.43 [8]
KNLNTS1.900.930.00711.51 [8]
Ca0.15(Sr0.5Ba0.5)0.85Nb2O63.611.710.02 [45]
NBT-0.07BZT5.712.030.02210.50 [8]
PZT4.141.420.0089.01 [8]
BaSn0.05Ti0.95O35.573.550.018 [37]
Ba0.85Sr0.15TiO37.23.140.009 [46]
at around depolarization temperature
NBT-6BT (352 K)9742.8172.178.250.25Present work
NBT-6BT-0.005Zn (395 K)1212.9458.946.010.22Present work
NBT-6BT-0.01Zn(400 K)96468.84112641566.03Present work
NBT-9BT(420 K)6052.7541.73.790.19Present work
NBT-9BT-0.005Zn(442 K)1874.0511610.250.32Present work
NBT-9BT-0.01Zn(442 K)1085.7731227.360.52Present work
N0.50B0.49T-6BT(393 K)47.815.1382.2450.771.03[30]
N0.5B0.5T-6BT(383 K)8728.1990.5266.441.46[30]
N0.5B0.51T-6BT(353 K)10328.44148.1397.492.02[30]
N0.51B0.5T-6BT (353 K)10534.37223.751202.41[30]
NBT-6BT (363 K)23.98.550.1965.5 [24,26,27,28]
NBT-6BT (363 K)53.319.70.1475.6 [25]
NBT-6Ba1.02T (358 K)7412601.64915 [25]
N0.52B0.52T-0.06BT (355.7 K)75.3326.920.39138.72 [24]
NBT-6BT-0.002Ta (351.8 K)146.152.20.4819.7 [27]
NBT-6BT-0.005La (346.5 K)86.130.80.5215.8 [26]
0.005La-NBT-6BT-0.002Ta (334.3 K)58.6220.940.187.60 [28]
* Single crystal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patel, S.; Lalitha, K.V.; Saurabh, N. Enhanced Pyroelectric Performance of Lead-Free Zn-Doped Na1/2Bi1/2TiO3-BaTiO3 Ceramics. Materials 2022, 15, 87. https://doi.org/10.3390/ma15010087

AMA Style

Patel S, Lalitha KV, Saurabh N. Enhanced Pyroelectric Performance of Lead-Free Zn-Doped Na1/2Bi1/2TiO3-BaTiO3 Ceramics. Materials. 2022; 15(1):87. https://doi.org/10.3390/ma15010087

Chicago/Turabian Style

Patel, Satyanarayan, Kodumudi Venkataraman Lalitha, and Nishchay Saurabh. 2022. "Enhanced Pyroelectric Performance of Lead-Free Zn-Doped Na1/2Bi1/2TiO3-BaTiO3 Ceramics" Materials 15, no. 1: 87. https://doi.org/10.3390/ma15010087

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop