International Materials Reviews ISSN: 0950-6608 (Print) 1743-2804 (Online) Journal homepage: http://www.tandfonline.com/loi/yimr20 Low-loss dielectric ceramic materials and their properties M. T. Sebastian, R. Ubic & H. Jantunen To cite this article: M. T. Sebastian, R. Ubic & H. Jantunen (2015) Low-loss dielectric ceramic materials and their properties, International Materials Reviews, 60:7, 392-412 To link to this article: http://dx.doi.org/10.1179/1743280415Y.0000000007 Published online: 13 Nov 2015. Submit your article to this journal Article views: 4 View related articles Full Terms & Conditions of access and use can be found at http://www.tandfonline.com/action/journalInformation?journalCode=yimr20 Download by: [Penn State University] Date: 26 November 2015, At: 11:01 FULL CRITICAL REVIEW Low-loss dielectric ceramic materials and their properties M. T. Sebastian*1, R. Ubic2 and H. Jantunen1 Downloaded by [Penn State University] at 11:01 26 November 2015 In addition to the constant demand of low-loss dielectric materials for wireless telecommunication, the recent progress in the Internet of Things (IoT), the Tactile Internet (fifth generation wireless systems), the Industrial Internet, satellite broadcasting and intelligent transport systems (ITS) has put more pressure on their development with modern component fabrication techniques. Oxide ceramics are critical for these applications, and a full understanding of their crystal chemistry is fundamental for future development. Properties of microwave ceramics depend on several parameters including their composition, the purity of starting materials, processing conditions and their ultimate densification/porosity. In this review the data for all reported low-loss microwave dielectric ceramic materials are collected and tabulated. The table of these materials gives the relative permittivity, quality factor, temperature variation of the resonant frequency, crystal structure, sintering temperature, measurement frequency and references. In addition, the methods commonly employed for measuring the microwave dielectric properties, important from the applications point of view, factors affecting the dielectric loss, methods to tailor the dielectric properties and materials for future applications, are briefly described. The data will be very useful for scientists, industrialists, engineers and students working on current and emerging applications of wireless communications. Keywords: Microwave dielectrics, Dielectric resonators, LTCC, ULTCC, Microwave applications, Microwave ceramics Introduction Microwave dielectric materials play a key role in global society, with a wide range of applications from terrestrial and satellite communications, including Internet of Things (IoT), software radio, GPS and DBS TV, to environmental monitoring via satellite, etc. Today low-loss dielectric materials are all-pervasive. The mobile phone is one of the most widely spread technologies on the planet. In many countries, the number of mobile subscriptions exceeds the population. The IoT is posed to make an explosive growth in the near future. In this paradigm, many every-day objects will be networked via radio-frequency identification (RFID), printed electronics and sensor network technologies. According to GSMA intelligence, the revenue from interconnected devices for mobile network operators alone in the segments of automotive, health, utilities and consumer electronics will be $1.3 trillion by 2020. In order to meet the specifications of future systems, new designs and improved or new microwave dielectric components are required. The recent progress in the IoT, microwave 1 Microelectronics and Materials Physics Laboratory, Department of Electrical Engineering, University of Oulu, Oulu90014, Finland Department of Materials Science & Engineering, Boise State University, Boise, ID, USA 2 *Corresponding author, email mailadils@yahoo.com 392 Ñ 2015 Institute of Materials, Minerals and Mining and ASM International Published by Maney for the Institute and ASM International Received 26 January 2015; accepted 15 June 2015 DOI 10.1179/1743280415Y.0000000007 telecommunications, satellite broadcasting and intelligent transport systems (ITS) has resulted in an increasing demand for low-loss dielectric materials. Indeed, low-loss dielectric oxide ceramics have revolutionised the microwave wireless communication industry by reducing the size and cost of filter, oscillator and antenna components in applications ranging from cellular phones to IoT. Wireless communication technology demands materials with highly specialised properties. The importance of miniaturisation cannot be overemphasised in any handheld communication application, as can be seen in the dramatic decrease in the size and weight of devices in recent years. This constant need for miniaturisation provides a continuing driving force for the discovery and development of ever smaller/lighter dielectrics which can outperform existing materials. Recently the demand for materials with low sintering temperature has increased not only to lower the energy cost of devices but also to integrate with polymers and silver-based electrodes. Several polymer-based (polymer–ceramic) composites have also recently been developed for wireless communication technology. In the present paper, we restrict our discussions to ceramic materials. For polymer-based composite dielectric materials, the reader is referred to the recent review by Sebastian and Jantunen.1 The number of papers published on low-loss microwave materials and related devices has increased considerably over the years as shown in Fig. 1. International Materials Reviews 2015 VOL 60 NO 7 Sebastian et al. Downloaded by [Penn State University] at 11:01 26 November 2015 1 Number of papers published on dielectric resonators (DRs) and devices versus year A dielectric resonator (DR) is an electromagnetic component that exhibits resonance for a narrow range of frequencies. The resonance is similar to that of a circular, hollow metallic waveguide except that the boundary is defined by a large change in permittivity rather than conduction. Dielectric resonators generally consist of a ceramic puck and require high values of relative permittivity (er) and quality factor (Q) and near-zero temperature coefficients of resonant frequency (tf). The quality factor, which is a function of resonant frequency, is sometimes expressed as Q f, the product of Q and the resonant frequency (in GHz). While Q f is not technically a dimensionless figure of merit, the units (GHz) are almost invariably dropped. The resonant frequency is determined by the overall physical dimensions of the puck and the permittivity of the material and its immediate surroundings. Optimising these three properties simultaneously is difficult. Oxide ceramics are critical elements in these microwave devices, and a full understanding of their crystal chemistry is fundamental to future development. Properties of microwave ceramics depend on several parameters including the processing conditions and the purity of starting materials. Design of the heating/cooling schedule requires knowledge of the formation mechanisms of various phases in multicomponent systems, and the starting powders must sinter to high density to obtain optimum electrical properties. Low-permittivity ceramics are used for millimetrewave communication and also as substrates for microwave integrated circuits. Medium-1r ceramics with 1r in the range of 25–50 are used for satellite communications and in mobile phone base stations. High-1r materials are used in mobile phone handsets where miniaturisation is very important. For millimetre-wave and substrate applications, temperature-stable, low-permittivity and high-Q are required for high-speed signal transmission with minimum attenuation. The signal transmission speed increases as the relative permittivity decreases. High-Q dielectrics minimise circuit insertion losses and can be used to create highly selective filters. In addition, a high-Q suppresses the electrical noise in oscillator devices. Although several manufacturers may produce similar components for the same application, there are subtle differences in circuit design, construction and packaging. Since frequency Low-loss dielectric ceramic materials drift of a device is a consequence of the overall thermal expansion drift of its unique combination of components, each design requires a slightly different tf for temperature compensation. Typically, ceramics with a specific tf in the range of 215 to þ15 ppm/uC are selected. In ceramic production, tf and 1r specifications must be produced to within demanding tolerances typically + 1%.2 Electronic circuits for the automotive industry, home electronics and telecommunications have to handle a steadily increasing amount of functionality within as tiny a space as possible. In the development of complex miniaturized circuits, flexible glass–ceramic composites, the so called low-temperature cofired ceramics (LTCCs), play a decisive role as a base material. LTCCs have become crucial in the development of various modules and substrates. This technology enables fabrication of three-dimensional ceramic modules with embedded silver or copper electrodes, and LTCCs with relative permittivity from *4 up to w100 have been developed showing low dielectric loss. These advantages make LTCC technology very attractive for a variety of microand millimetre-wave applications.3 The important characteristics required for LTCCs are (a) densification temperature v950uC (b) 1r in the range 5–70 (c) Q f w1000 (d) tf close to zero (e) high thermal conductivity (f) preferably low thermal expansion and (g) chemical compatibility with the electrode material. Low sintering temperatures are required to avoid melting metallic conductors like silver or gold in the fabrication of dielectric devices.3 Most conventional electroceramics do not meet the basic requirements with regard to sinterability for LTCC technology since they have relatively high sintering temperatures. The different methods used to reduce the sintering temperature of dielectrics include: (1) addition of low melting-temperature glass phases, (2) addition of low melting-point compounds such as Bi2O3, B2O3, V2O5 or CuO and (3) the use of chemical processing in order to achieve smaller particle sizes. The first method, while commonly found effective in decreasing the sintering temperature, usually results in a degradation of microwave dielectric properties. The selection of glass materials is very important for sintering glass–ceramic composites, since the liquidation of glass takes a dominant role in the viscous flow mechanism during sintering; hence, this method remains the focus of intense research. The dielectric table (supplementary file) lists the key property data of microwave dielectric materials available from published and, to a far lesser extent, reputable unpublished sources. These data are the relative permittivity (1r), the product of the Q factor and the frequency (Q f ), the frequency of measurement ( f ), the temperature coefficient of the resonant frequency (tf), sintering temperature and crystal structure or structural family. Measurement of microwave dielectric properties The three important characteristics of an ideal low-loss dielectric material are application optimised value of relative permittivity (1r), low dielectric loss (loss tangent, tand) and low temperature coefficient of resonant frequency (tf). These three properties and different measurement methodologies to measure them are briefly discussed in the following sections. International Materials Reviews 2015 VOL 60 NO 7 393 Sebastian et al. Low-loss dielectric ceramic materials Permittivity When microwaves enter a dielectric medium, they are ; therefore slowed down by a factor equal to e21/2 r l0 c ld ¼ pffiffiffiffi ¼ pffiffiffiffi 1r n 1r [ n¼ c pffiffiffiffi ld 1r ð1Þ At resonant frequency, l ¼ f0 and ld , D (diameter of resonator); therefore c c 2 ð2Þ f 0 ¼ pffiffiffiffi [ 1r ¼ D 1r Df0 Downloaded by [Penn State University] at 11:01 26 November 2015 Equation (2) is only valid in the case of resonators in free space. It fails for resonators in more realistic situations ( e.g., on microstrips, in cavities, between shorting plates, etc.). In order to calculate permittivity in these geometries, several techniques have been developed and variously discussed. Perturbation techniques rely on the shift of f0 (and Q) of a resonant cavity caused by the presence of a dielectric disc or sphere. Optical methods at microwave frequencies are suited to measurements at which l,1 cm and require a large amount of material. Transmission-line methods have the practical difficulty of requiring a very small waveguide for l,4 mm. All of these methods have an accuracy of approximately ¡1%. The exact resonance method proposed by Karpova4 and further developed by Hakki and Coleman,5 Courtney6 and others yields errors of only ¡0.1% but is limited to the accuracy of the measurements of resonant frequency and sample dimensions. The reader is referred to the recent book2 for details of these techniques. In this paper, we restrict the discussion to the measurement of the relative permittivity and loss tangents of low-loss dielectric materials. Hakki – Coleman method Karpova4 used a re-entrant cavity for the measurement of dielectric properties, but the physical size of the resonant structure required could be problematic for the low-millimetre range. In order to avoid the problem of physical size while maintaining high accuracy, Hakki and Coleman5 instead proposed an open-boundary resonant structure in which a dielectric rod was positioned between much larger conducting plates (Fig. 2). The characteristic equation which describes this condition for an isotropic resonator in a TE0mp mode: a J 0 ðaÞ K 0 ðbÞ ¼ 2b J 1 ðaÞ K 1 ðbÞ K1(b) are modified Bessel functions of the second kind of orders zero and one, respectively. The parameters a and b are functions of geometry, resonant wavelength and permittivity: sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2ffi 2pa c 1r 2 ð4Þ a¼ vp l0 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 2pa c b¼ 21 ð5Þ l0 vp where c is the speed of light, a is resonator radius and vp is the phase velocity in the resonator such that: c pl0 ð6Þ ¼ vp 2t where p is the number of longitudinal variations of field along the axis and l0 ¼ c/f0. Clearly vp can be calculated from thickness and resonant frequency alone; and b can then be calculated from vp, frequency and radius. The characteristic equation (3) is transcendental and requires a graphical solution. Hakki and Coleman5 used analogue mode charts to relate various {am} to each corresponding value of b, resulting in somewhat limited accuracy (Fig. 3). Although this technique is sometimes called the Courtney method,6 ‘Courtney, actually, only perfected and scrutinised a parallel-plate arrangement introduced [10 years] earlier’ by Hakki and Coleman.5 Courtney also adapted the technique to the use of coaxial probes (an innovation introduced 4 years earlier by Cohn and Kelly7), allowing a greater range of sample dimensions. An improvement in accuracy over a purely graphical approach can be achieved by numerically solving for each Bessel/modified Bessel function rather than trying to read values off the mode charts of Hakki and Coleman5 or even relying on curve fits. With modern computers, ordinary Bessel functions and modified Bessel functions can be numerically calculated, and these numerical methods make it possible to solve equation (3) for b<10. The algorithm employed in the HakCol program8 starts by calculating b from the resonator radius and resonant frequency. Next an approximate corresponding value for a is calculated using a curve fit to the m ¼ 1 (TE01p) mode chart of Hakki and Coleman5 (Fig. 3). The polynomial which describes the curve in Fig. 3 is: ð3Þ where J0(a) and J1(a) are Bessel functions of the first kind of orders zero and one, respectively. K0(b) and 2 Schematic sketch of Courtney set-up for measuring the dielectric constant under end shorted condition (after Ref. 6) 394 International Materials Reviews 2015 VOL 60 NO 7 3 Mode chart (after Ref. 5) Sebastian et al. Downloaded by [Penn State University] at 11:01 26 November 2015 a ¼ 2:3508 þ 0:34969b – 0:051220b2 þ 0:0044392b3 – 0:00020633b4 þ 3:9411 £ 1026 b5 ð7Þ The mode TEnml, the integer n denotes the azimuthal variation, m radial variation and l the axial variation. The a so calculated is used as a first approximation in order to calculate er. Next, equation (3) is evaluated and if the two sides are unequal then er is adjusted accordingly, a re-calculated, and the process iterates until equation (3) is satisfied. The entire algorithm and the HakCol program is detailed in ref.9 The TE011 mode is used for the measurements since this mode propagates inside the sample but is evanescent outside; therefore, a large amount of electrical energy can be stored in high-Q DRs.10 In the end-shorted condition, the E field becomes zero close to the metal wall and electric energy vanishes in the air gap.7 The TE and TM modes do not contain electric and magnetic fields in the axial (z) direction. For the TE011 mode only the azimuthal component of the electric field exists and the error because of the air gap is practically eliminated.11 For cylindrical resonators, TE and TM modes exist only if the azimuthal mode index m ¼ 0 otherwise all other modes are hybrid, i.e., they have all six electromagnetic components. Hybrid modes are usually divided into two mode families: HE and TM. They are only occasionally used in measurements of dielectrics (e.g., for uniaxially anisotropic crystals). This method is proposed as one of the international standard IEC techniques12 for measurements of the complex permittivity of low-loss solids. Hennings and Schnabel13 studied the reproducibility of the 1r measured by this end-shorted method using 10 different samples prepared in a batch. Their results showed a maximum variation of 0.6% in 1r. In this method, the 1r is measured only at one resonant frequency corresponding to the TE011 mode. If one can identify other resonant modes, then it is possible to measure 1r at other resonant frequencies. By using the resonant modes TE011, TE021, TE031 and TE041, the 1r of a sample can be measured over a range of frequencies. It may be noted that as the 1r increases, the resonant frequency decreases and as the dimensions of the sample decrease the resonant frequency increases. Shielded resonator in dielectric-rod waveguide method For a high-Q material in a cavity, as proposed by Itoh and Rudokas14 and modified by Kajfezz and Guillon15 (Fig. 4), most of the electrical field is contained within the resonator itself (region 6), and very little exists in regions 1 and 2, and even less in regions 3 and 5. To a fair first approximation, then, the fields in regions 3 and 5 can be ignored. For the TE01d modes in this geometry, the requirement for continuity of fields leads to two simultaneous eigenvalue equations: J o ðkr1 aÞ kr2 K 0 ðkr2 aÞ ¼2 J 1 ðkr1 aÞ kr1 K 1 ðk2 aÞ bL ¼ w1 w2 þ þ lp l ¼ 0; 1; 2; 3. . . 2 2 Low-loss dielectric ceramic materials ð8Þ ð9Þ The symbol k represents the radial propagation constants in the different regions of the model, which are functions of both frequency and dielectric constant; and 4 Resonator in a cavity (after Ref. 9) r is the radial distance from the geometric centre. The arguments of the various Bessel functions are the eigenvalues of the system, where kr1a is called the eigenvalue of the TE0n mode, and kr2 is given by: qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð10Þ kr2 a ¼ ðk0 aÞ2 ð1r6 2 1r4 Þ 2 ðkr1 aÞ2 where k0 is called variously the propagation constant, wavenumber or phase constant of free space, and has units of m21: pffiffiffiffiffiffiffiffiffi k0 ¼ v0 10 m0 ð11Þ In equation (9), b is the propagation constant of the resonator. If p is the number of axial variations of the field along the resonator’s height, then p ¼ l þ d, where l is an integer and d is a non-integer number smaller than unity which depends in a complicated way on propagation constants and geometry. Whereas for the TE011 mode discussed above, l ¼ 1 and d ¼ 0, for the TE01d mode, l ¼ 0 and d ?.0. The symbols (w1 and w2 are called the phase angles and are complex hyperbolic functions of the cavity geometry and the propagation constant of the resonator. The entire algorithm and the ErCalc program is detailed in reference.9 Correction for porosity The porosity in the sintered ceramic disc influences the measured 1r and thus the measured 1r should be corrected to isolate the actual dielectric permittivity. This correction can be performed in a variety of ways. The Maxwell Garnett16 approximation treats one of the components as a host in which inclusions of the other component are embedded. Lichtenecker’s17 logarithmic mixing rule assumes a randomly connected second phase and, although it is much used, is actually one of the least accurate mixture rules available. By contrast, the Bötcher mixture rule18 assumes a dispersion of spherical porosity (or another second phase) in a mixture of both solid and air (or another second phase), like that of Bruggeman,19 thereby allowing for International Materials Reviews 2015 VOL 60 NO 7 395 Sebastian et al. Low-loss dielectric ceramic materials the interaction between the two phases and increasing the accuracy even for high values of porosity: 1rm 2 1r2 d1 ð1r1 2 1r2 Þ ¼ 31rm 1r1 þ 21rm ð12Þ where erm, er1, and er2 are the permittivities of the mixture, phase 1 and phase 2, respectively, and d1 is the volume fraction of phase 1. This rule is also based in part on the work of Wiener20 and Stratton21 and supported by Reynolds.22 The various equations typically only diverge significantly for very high or very low densities of second phase and re-converge for densities of 0 and 100%. Maxwell’s equation, in particular, slightly inflates the value at intermediate densities, presumably because it does not allow for the interaction between the two phases. QU ¼ ð1 þ bc1 þ bc2 ÞQL Measurement of loss tangent/quality factor Downloaded by [Penn State University] at 11:01 26 November 2015 The measured Q value is commonly the loaded quality factor (QL) taking into account the external circuit (the network analyser with coupling probes). However, if the measurement is arranged under very weak coupling the QL is the same as unloaded one Qu and is obtained from the following equation. QL ¼ Qu ¼ f Df confinement is not complete in the z direction. As shown in Fig. 5, the spacer isolates the sample from the effects of losses because of the finite resistivity of the metallic cavity. After identifying the mode, the resonant frequency and 3dB bandwidth are determined. The network analyser is then calibrated and S11 and S22 are measured at the resonant frequency (Fig. 6). From these values, the coupling coefficients bc1 and bc2 for the coupling ports are determined using the relations bc1 ¼ (12S11)/(S11 þ S22) and bc2 ¼ (12S22)/(S11 þ S22), where S11 and S22 are reflection coefficients of ports 1 and 2.30 Figure 7 shows the typical resonance spectra in reflection and transmission configuration of a Ba(Mg1/3Ta2/3)O3 ceramic sample having 1r ¼ 38. The TE01d mode frequency is noted and the unloaded Q factor is measured. From the measured QL, QU can be calculated as ð13Þ where f is the measured resonance frequency and Df is 3 dB band width of the peak. There are various methods which enable measurement of the quality factors of low-loss dielectrics.23–31 One should however keep in mind that not all of them take into account practical effects introduced by a real measurement system, such as noise, cross-talk, coupling losses, transmission-line delay and impedance mismatch. Inadequate accounting of these effects may lead to significant uncertainty in the measured Q-factor. For example the quality factor can be measured by Hakki and Coleman’s end-shorted method,2,5–7,25–28 but the quality factor measured by this method will be somewhat low since loss occurs because of the conducting plates and radiation effects. Fortunately, corrections for conductor losses can be applied knowing the surface resistance of the conducting plates. ð14Þ In some cases the desired mode (TE01d) may be close to other modes but this method allows slight change of the cavity volume by rotating the top screw, which enable separation of the modes. This action enables the identification of the desired resonant mode and in addition allows the cavity to measure samples of different dimensions. Figure 8 shows a typical test 5 The cavity set-up for the measurement of Q factor TE01d mode DR method To avoid the problems of the conduction and radiation losses, the Q of a DR sample can be measured by using the cavity method in which the DR is placed on a lowloss (e.g., single crystal quartz or Teflon) spacer inside the cavity. This method is proposed by Krupka et al.23,29 using a transmission-mode cavity. It enables measurement of the quality factor (Q), permittivity (1r) and temperature coefficient of resonant frequency (tf) of the DRs, which is placed inside a cylindrical metallic cavity usually made of copper. The inner surfaces are polished and gold or silver coated. A loop coupling is used to feed microwave to the DR Since the cavity has an infinite number of modes, the diameter and height ratio of the sample is commonly kept on the level 2–2.5 to get maximum mode separation. Since the electric field is symmetric in this measurement method, the sources of loss owing to the cavity are reduced. In this method the TE011 mode is designated as TE01d, since the field 396 International Materials Reviews 2015 VOL 60 NO 7 6 The TE011 resonance of a ceramic puck with 1r 5 38 under end shorted condition Downloaded by [Penn State University] at 11:01 26 November 2015 Sebastian et al. Low-loss dielectric ceramic materials 7 Microwave resonance spectra of Ba(Mg1/3Ta2/3)O3 ceramic with 1r 5 24 a reflection b transmission configuration 8 The cavity manufactured by QWED for quality factor measurement (courtesy, J Krupka QWED, Warsaw, Poland) fixture manufactured by QWED, Warsaw, Poland. The evaluation of the permittivity and the dielectric loss tangent of the sample under test require rigorous electromagnetic analysis. QWED uses the Rayleigh–Ritz method in their software. The TE01d mode DR method is one of the most accurate techniques for measuring especially loss tangent of isotropic low-loss materials.29,31 The inverse of measured unloaded Q-factor is approximately equal to the dielectric loss tangent if all parasitic losses can be neglected (true in the cases when the permittivity of the sample is large) and if the electric energy filling factor can be assumed to be equal to unity. One must keep in mind that these assumptions are not valid when the sample has very low dielectric loss or permittivity value. In the first case the conductor losses must be taken into account. What comes to the low-permittivity materials, the electric energy filling factor in the sample is substantially smaller than 1. However, the advantages of the cavity method using the TE01d mode are easy mode identification, small parasitic losses and lack of mode degeneracy.23 On the other hand, the evaluation of tand requires advanced numerical computations, which can only be done employing dedicated computer programs because of the absence of exact solutions of Maxwell’s equation. The uncertainty in dielectric loss tangent using TE01d mode cavity method with optimized enclosure is of the order of 0.03 tand. The frequency band this method is feasible depends on the size and permittivity of the samples, and the cavity geometry. Higher frequency measurements are performed by using smaller cavities and samples, or by using several higher order quasi-TE0nm modes.32 Valant et al.33 reported the effect of the test cavity dimensions on the microwave dielectric properties of the ceramic resonator. The electromagnetic field could penetrate into the conducting walls of the test cavity (skin effect) lowering the Q factor. With large size of the test cavity this source of error can be avoided. Thus in order to derive the unloaded Q value, the test cavity should be large enough. A good practice is to select the test cavity size in such a way that the TE01d mode of the DR is the lowest resonance and hence it can be easily identified. This is especially true in the case DRs with permittivity .20 when increase of the size of the test cavity is needed to move the resonant modes of the cavity to lower frequencies. Figure 9 shows how the measured quality factor decreases with the cavity diameter/disc diameter ratio. Thus it is advisable to use 3–5 times larger cavity compared to the size of the test sample. In addition the surface resistance of cavity walls can be calculated from the quality factor of the TE011 resonance of the empty cavity.15 Strip line excited by cavity method Magnetic coupling of the DR to a 50 V microstrip line is used in the microstrip line excited cavity method, International Materials Reviews 2015 VOL 60 NO 7 397 Sebastian et al. Low-loss dielectric ceramic materials shielded resonator configuration like in shielded cavities, the power dissipated in the resonator is given by Pd ¼ 1 2 jS 110 j2 2 jS 210 j2 ð16Þ The coupling factor bc is a function of the distance between the DR and the microstrip line under fixed shielding conditions. According to Khanna and Garault34 the unloaded voltage transmission coefficient S21u is sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 S 21u ¼ S210 ð17Þ 1 þ S2210 9 Variation of Qf with ratio of cavity diameter/sample diameter (after Ref. 33) Downloaded by [Penn State University] at 11:01 26 November 2015 as shown in Fig. 10 along with the equivalent circuit.34 In this method the Q factor is estimated through the so called coupling factor, bc, which is the ratio of the resonator-coupled resistance R at the resonant frequency to the resistance external to the resonator. bc ¼ R S 110 ¼ Rext S 210 ð15Þ Here S110 and S210 are the real quantities of the reflection and transmission coefficients, respectively, at the resonant frequency. When the coupling factor bc is equal to one, the power dissipated in the external circuit is the same as the power dissipated in the resonator (Pd), which is equally divided into the power reflected to the generator (Pr ¼ S2110) and the power transmitted to the load (Pt ¼ S2210). In the 10 Schematic diagram of a dielectric resonator (DR) coupled to a microstrip line a and b equivalent circuit (after Ref. 34) 398 International Materials Reviews 2015 VOL 60 NO 7 S21u corresponds to the voltage transmission coefficient of the unloaded resonator. The frequencies f1 and f2 corresponding to S21u given by equation (17) are measured (Fig. 11) and their difference Df ¼ ( f22f1) is calculated. When the resonance frequency, f, corresponds to the peak of the S21 curve and Df is known, the unloaded quality factor, Qu, is calculated using equation (13). The experimental set-up for the Q measurement by the microstrip line excited by cavity method is shown in Fig. 12. In this a 50 V microstrip line of width 3 mm is etched on RT-Duroid 5880 (1r ,2.2 and thickness 1.9 mm) to the bottom wall of a rectangular cavity made of copper. The cavity is then excited using 3.5 mm microstrip edge connectors. The DR is placed near the microstrip line and the TE01d mode is identified. After system calibration, the resonant frequency is measured, the transmission coefficient S210 corresponding to f is recorded, and S21u is calculated using equation (17). Finally the unloaded quality factor is calculated through Df corresponding S21u and f. Whispering gallery mode resonator method TE01d, TM01d or HE11d modes of the DRs are normally measured by the end-shorted TE011, TE01d (cavity) or stripline methods.5,6,25,31 However, the measured Q of these modes depends not only on the material’s tand but also on the radiation and conduction losses of the cavity as stated earlier. Thus simple measurement by the above methods for very low-loss dielectrics are not accurate enough. The Whispering Gallery modes (WGMs)35–39 11 Typical resonant curve of a dielectric resonator (DR) coupled to a microstrip line used in determining the quality factor by the stripline method Sebastian et al. Low-loss dielectric ceramic materials the diameter of the DR the number of modes in a bandwidth increases. This means that samples with small resonator diameter, the frequency interval between two successive modes will be large. Dielectric resonators acting in WGMs can be excited in different ways. In the low-frequency range, an electric or magnetic dipole is used. However, this type of excitation is stationary, and travelling WGMs cannot be excited. In the millimetrewavelength frequency region dielectric image waveguides or microstrip transmission lines are used to excite travelling WGMs. Split-postdielectric resonator method Downloaded by [Penn State University] at 11:01 26 November 2015 12 The experimental set-up for measuring quality factor by the stripline method. The dielectric resonator (DR) is coupled to the stripline has been reported to be able to confine the entire field within the resonator which yields negligible radiation and conductor losses at microwave frequencies. The Q factor of WGM dielectric resonators is limited only by the intrinsic losses in the dielectric material leading to the situation where the measured WGM Q factor is approximately equal to 1/tand. In this method, most of the electromagnetic energy is confined to the dielectric near the perimeter of the air-dielectric.37,38 One additional advantage of using the WGMs technique is that it allows measurements of two permittivity components of uniaxially anisotropic materials and is very useful, for example, in the case of several single-crystals. In this technique measurements of resonant frequencies and Q factors of two modes belonging to different mode families employing rigorous numerical analysis, e.g. mode-matching, are needed. The electrical energy filling factors for E (quasi-TM mode) and H (quasi-TE mode) modes are then calculated.37,39 P1’ ¼ 2 Lf 1’ L1’ f ð18Þ P1ll ¼ 2 Lf 1ll L1ll f ð19Þ The dielectric tand can be solved38 using the equation Q21 ðEÞ ¼ tan dðP1’ þ P1II Þ þ Rs =G ðEÞ ð20Þ Q21 ðHÞ ¼ tan dðP1’ þ P1II Þ þ Rs =G ðHÞ ð21Þ where Rs is the surface resistance of the cavity, 1ll is the permittivity parallel to the anisotropic axis and 1His the one perpendicular to it. In general the conductor losses decrease as the surface resistance becomes smaller and as the geometric factor (G) increases.39 However for the WGM the geometric factor, G, is sufficiently large and thus the effect of the cavity can be ignored. In addition the radiation losses are negligible for these modes. The spurious modes in WGM method dominate since the propagation constant along the z axis is very small and unwanted modes leak out axially. The WGM dielectric resonators are classified as WGEn,m,l or WGHn,m,l.These correspond to the case where the electric field is essentially transverse, or axial. WGMs are periodic according to the azimuthal number, and with The split-post-dielectric resonator (SPDR), shown in Fig. 13, is an accurate method for measuring the complex permittivity and loss tangent of substrates and thin films at a single frequency in the range of 1–20 GHz.40–43 This arrangement allows the formation of an evanescent electromagnetic field, not only in the air gap, but also in the cavity region for radii greater than the radius of the dielectric resonators which simplifies the numerical analysis and reduces possible radiation effects. In the SPDR method, a flat sample of the test material is inserted through one of the open sides of the fixture and positioned between two low-loss dielectric rods or resonators kept in a metallic enclosure. The sample is in contact with one of the resonators and separated from the other by a small air gap. The electric field in resonators is parallel to the surface. This means that the test sample should have strictly parallel faces, the thickness of the sample should be less than that of the fixture air gap, and the sample should have enough area to cover the inside of the fixture. In these conditions, the accuracy of the measurement is not affected by the air gap between the sample and the resonator. The required thickness of the sample also depends on the 1r of the material and materials with high 1r must be thinner. Several modes are commonly excited, but the TE01d mode is preferred since it is insensitive to the presence of air gaps perpendicular to z-axis of the fixture. The complex permittivity is calculated based on electromagnetic modelling of the split-post-resonant structure using the Rayleigh–Ritz technique.40 The real part of the complex permittivity can be iterated from the measured resonant frequencies and thickness of the test sample, h, using the following equation 41 19r ¼ 1 þ f0 2fs hf0 K 1 ð19r ; hÞ ð22Þ where f0 is the resonant frequency of the empty SPDR, fs is the resonant frequency of the SPDR with the dielectric sample. K1 is a function of 1’r and h and is evaluated for a number of 1r using the Rayleigh–Ritz technique.40 The tand of the sample is calculated from the measured unloaded Q factors of the SPDR with and without the dielectric sample from 1 1 1 2 2 ð23Þ Pe tan d ¼ Qu Qd Qc 21 denote losses of the dielectric and where Q21 d and Qc metallic parts of the resonator, respectively. Pe is the electric energy filling factor of the sample given by the following equation International Materials Reviews 2015 VOL 60 NO 7 399 Sebastian et al. Low-loss dielectric ceramic materials 13 Schematic sketch of split-postdielectric resonator (SPDR) Downloaded by [Penn State University] at 11:01 26 November 2015 N X 1 ¼ Pei tan di Qd i¼1 ÐÐÐ Pei ¼ ÐÐÐ vd 1i jEj2 dv 2 vt 1ðvÞjEj dv ð24Þ accuracy and is convenient and fast for low-loss laminar dielectrics such as substrates or LTCCs, printed circuit boards and even for thin films. ð25Þ Measurement of dielectric properties of powder samples where Pei and tandi are the electric energy filling factor and the dielectric loss tangent for the ith dielectric region, respectively. Uncertainty of the permittivity depends on the sample thickness h as follows: Der/er ¼ ¡(0.0015þDh h21) and the accuracy of the loss tangent is , ¡0.03 tand. For the complex permittivity of a sample, the resonant frequencies and Q factors of the empty SPDR and the SPDR containing the test sample must be measured. The SPDR is operated in a particular mode with a particular resonant frequency which depends on resonator dimensions and to a limited extent the electrical properties of the test sample; thus each SPDR is designed for a particular nominal frequency and the actual measurement is taken close to this frequency. The size of the sample is determined by the nominal frequency. This means that, for example, a nominal frequency of 5–6 GHz requires a minimum sample size of 3063062.1 mm. QWED provides SPDRs (Fig. 14) with dedicated software for the evaluation of permittivity and loss tangents. Compared to the reflection-transmission methods, the SPDR provides superior Low-loss dielectric ceramic powders are also more and more used to enhance dielectric properties of polymers. In the last 10 years interesting polymer ceramic composites have been introduced enabling free adjustment of permittivity of devices being important in several applications areas where design of telecommunication devices has limited space. Additionally one overwhelming example is advanced printed electronics, where inks are based on low temperature curing polymers with embedded dielectric ceramic particles. In these cases it is crucial to know the dielectric properties of powder particles, which can differ significantly if compared to the bulk properties. However, very few papers are published44–50 on the measurement of powder samples. More recently Tuhkala et al.48–50 reported an indirectly coupled open-ended coaxial cavity resonator method operating in TEM mode at 4 GHz to estimate the relative permittivity and loss tangents of powder samples. In this method the open-ended coaxial cavity with optimised dimensions and conductivity is filled with dielectric materials in powder form and the effective dielectric properties are determined by the shift of resonant frequency and change in Q factor between an empty resonator and a filled resonator. A schematic set-up for the measurement is shown in Fig. 15. For a completely filled cavity, the effective dielectric constant is given by 1r ¼ ðc=4Lfr Þ2 14 Photographs of split-postdielectric resonators (SPDRs) produced by QWED, Poland. (Courtesy, QWED Poland) 400 International Materials Reviews 2015 VOL 60 NO 7 ð26Þ where c is the speed of the light in vacuum, L is the resonator length and fr is the resonant frequency. The dielectric constant of powder samples can be obtained from measured resonator response, volume fractions of each phase (powder, air and different phases) and calculating the effective dielectric constant using Bruggeman symmetric19 and the Looyenga51 mixing rules. In the same manner the effective loss tangent of the powder sample (with different phases) is estimated from the difference in Q factor between an empty resonator and a filled resonator using the equation Sebastian et al. Low-loss dielectric ceramic materials The tf value is measured by following the drift in the resonant peak frequency ( fo) as a function of temperature. Choosing an arbitrary temperature to use as standard, say 258C, the ratio fo(T )/fo(258C) can be defined at each temperature. Then, tf is obtained from the slope of a graph of fo(T )/fo(258C) versus the reduced temperature T9 ¼ T2258C: Downloaded by [Penn State University] at 11:01 26 November 2015 tf ¼ 15 Schematic set-up for microwave measurements of powder samples (after Ref. 53) tandeff ¼ 1=Qfilled 2 1=Qempty ð27Þ and utilising general mixing rules. The method expects careful estimation of the volume fraction of the powder, homogeneous distribution of the powder throughout the cavity, management of measurement environment (mainly humidity) and probe coupling, which should be loose enough not disturbing the measurement itself and producing symmetric resonance peak. Tuhkala et al.52,53 estimated the microwave dielectric properties of several materials with reasonable accuracy by this technique. The method is shown to be useful for studying the properties of powders in several ways. They also reported estimation of humidity level of the powders,53 effect of surfactant treatment,54 evaluation of the amount of two powder phases52 using the above method.48 Measurement of temperature coefficient of resonant frequency (tf) The temperature coefficient of resonant frequency, tf, is the parameter which indicates how much the resonant frequency drifts with temperature. In many cases microwave devices in order to operate correctly require tf value close to zero. The tf relates to the linear expansion coefficient of the sample itself, aL, and dependence of the material’s dielectric permittivity with temperature.2 Mathematically tf relates to the temperature coefficient of the permittivity, t1, as follows tf ¼ 2aL 2 t1 2 ð28Þ It may be noted that equation (31) is valid for 100% electric energy storage in the sample and thermal expansion of the metal cavity enclosing the DR is negligible. In general the electronic ceramic materials has aL close to þ10 ppm/8C meaning a significant influence of t1 on tf. dð f o =f o ð25o ÞÞ 1 dfo ¼ dT9 f o ð25Þ dT9 ð29Þ The actual curve of the graph is never linear but very nearly parabolic and is easily fitted with a quadratic equation in terms of fo(T)/fo(258) and T2258. In this equation, the first-order coefficient is tf, which is the slope of the curve at 258C. The second-order coefficient is the so-called non-linearity or double-derivative factor (tf’). Its sign shows whether the parabola is concave up (þ) or down (2), and its magnitude indicates the severity of curvature. Both numbers are usually scaled up by 106 and reported in parts per million per degree (ppm/8C) or MK21. If tf is measured by the cavity method, then the thermal expansion of the cavity during heating (or contraction during cooling – measurements are best done upon cooling) limits the accuracy of the method, in which case very low-aL materials (e.g., invar, aL ¼ 1.2 ppm/8C) can be used for the cavity. The WGMs and also for TE01d mode resonant structures are reliable methods for tf measurements since the thermal expansion of the cavity is negligibly small especially if the relative permittivity of the sample is large and the sample is situated away from cavity walls. The temperature coefficient of dielectric permittivity t1 can be obtained by the parallel-plate capacitor method using an LCR meter at low frequency. Factors affecting dielectric losses The tand is known to be very sensitive to humidity6,55 meaning that the microwave measurements should be done in a humidity-controlled room. Before the experiments the samples should be heated in an oven to remove adsorbed moisture. However, the content of the dielectric has the main effect on the loss values. When especially low-loss materials are desired, the starting materials should be selected the way they have the lowest possible concentration of dipoles and charge carriers with the lowest possible mobility.56 Dielectric loss is also affected by disordered charge distributions in the crystal lattice56,57 which occur if the charge distribution in a crystal deviates from perfect periodicity. In 1964 Schlömann56 reported that the loss tangent increases in ionic non-conducting crystals when ions are disordered in such a way that they violate periodicity. The loss tangent thus depends strongly on the spatial correlation between charge deviations and is negligible if the disordered charge distribution in the crystal maintains charge neutrality within a short range of the order of the lattice constant. The intrinsic quality factor (QU ¼ 1/tand) of any given material is frequency dependent. For many materials tand almost linearly increases as the frequency increases and thus often the intrinsic quality factor is reported as (QU f ¼ f/tand) (in GHz) as a first approximation. This is most valid for well-densified ceramics within a limited International Materials Reviews 2015 VOL 60 NO 7 401 Sebastian et al. Low-loss dielectric ceramic materials Downloaded by [Penn State University] at 11:01 26 November 2015 frequency range. In practice, higher QUf values for samples measured at higher frequencies (5–12 GHz) than at lower frequencies. More recently Li and Chen reported58 that the product Q f is frequency dependent and increases with frequency. The frequency dependence of Q f value is attributed to the presence of defectsinduced extrinsic dielectric loss. It may be noted that larger samples resonating at lower frequencies statistically contain more imperfections than smaller ceramic discs resonating at higher frequencies. The presence of porosity decreases the Q factor further because of presence of moisture in the pores. A fundamental theory of intrinsic losses set the lower limit of losses found in pure defect-free single crystals.59 In a dielectric several phonon processes contribute to intrinsic losses and their importance depends on the ac field frequency, temperature range and symmetry of the crystal under consideration. The loss mechanisms are different for a crystal with and without a centre of symmetry. Gurevich & Tagantsev59 obtained numerical estimates of tand of ideal crystals. For an ideal crystal with a hexagonal symmetry when T%TD. tan d ¼ gvðkTÞ5 1r rv 5s h 2 ðkT D Þ2 ð30Þ and for rhombohedral or cubic symmetry gv 2 ðkTÞ4 tan d ¼ 1r rv 5s hðkT D Þ2 ð31Þ where g is a dimensionless anharmonicity parameter ranging between 1 and 100, v is the angular frequency, k ¼ Boltzmann constant, T ¼ absolute temperature, vs ¼ sound velocity, TD ¼ Debye temperature and r is the mass density. Owing to the complicating factors introduced by a variety of extrinsic mechanisms, there is no predictive theory to account for the microwave loss in dielectric ceramics meaning that finding new dielectric resonator materials is largely done by trial and error and involves the preparation and testing a large number of samples. This is a laborious and time-consuming job. The Q factor is highly dependent on not only the extrinsic and intrinsic quality of the ceramic sample but also the method of measurement, the measurement environment and the frequency at which the sample is measured. A given material sample may exhibit greatly differing Q values when tested in different test fixtures and environments which may vary in size, shape, conductor quality, coupling, type of sample support, ambient temperature and relative humidity. data on materials of identical composition and manufactured in different laboratories using different processing conditions would be expected to lead to small variations in properties. The dielectric data measured by impedance methods at low frequencies are not included in the Table since it is unreliable when the loss tangent is less than 1023. The Table shows nearly 4000 low-loss dielectric ceramic compositions reported in the literature. About 35% of them belong to the interesting, widely applicable perovskite family. The analysis of crystal systems shows that many of them enable interesting low-loss dielectric properties. The most common one is orthorhombic (35%), followed by hexagonal (18%), monoclinic (12%), cubic (12%) and tetragonal (10%) crystal systems. About 60% of the reported lowloss dielectric ceramics are based on alkaline earth metals like Ba, Sr, Ca or Mg. Additionally, titanates (46%) and compositions containing rare earths (40%) or tantalates/niobates (39%) are widely reported. Silicates and tungstates are also well represented. Understanding the relation between bonding mechanisms and the microwave dielectric properties is essential. The silicates, to mention one example, have in general predominantly covalent bonding which geometrically restricts the movement of atoms and leads to low dielectric loss. On the other hand, the low dielectric polarisability of silicon and the strong covalent bonds in silicates yield low 1r. Thus, in general, the silicates and tungstates have low 1r, niobates and tantalates have medium 1r, and titanates have relatively larger 1r. Another example is formed by an octahedral arrangement of anions within a perovskite family of materials where octahedral tilting, brought on by a geometrical instability related to the relative sizes of A and B cations, is accompanied by symmetry lowering and affects the dielectric loss. As expected the reported quality factors of the microwave dielectric ceramics decrease significantly with increasing relative permittivity as shown in Fig. 16. The inset in Fig. 16 shows the variation of quality factor frequency product with relative permittivity in the logarithmic scale. The 0.993MgO–0.007B2O3 material has the highest quality factor (Qf ¼ 773 700 GHz, with er ¼ 9.3 and tf of 255 ppm/uC). On the other hand, the composition 0.8SiO2–0.2B2O3 has the lowest relative permittivity (er ¼ 3.6, Qf ¼ 70 600 GHz and tf of 211 ppm/uC). Its relatively low Qf value can be explained by the glassy nature of this material. AlPO4 is difficult to densify and Low-loss dielectric ceramics A list of low-loss ceramic dielectric materials with sintering temperature, crystal structure, relative permittivity, quality factor-frequency product, measurement frequency, temperature variation of resonant frequency and references are given in the supplementary file. In tabulating these data, we make no judgement on the measurement method and the reliability of the result. The ceramic properties such as porosity, grain size, raw materials used, measurement methods and equipment used for measurements affect the dielectric properties and readers should be aware that exact comparison of 402 International Materials Reviews 2015 VOL 60 NO 7 16 Variation of Qf as a function of relative permittivity Downloaded by [Penn State University] at 11:01 26 November 2015 Sebastian et al. it has a very low permittivity of 3.0. These low-er materials are important for increasing the signal speed in communication systems. At the other end of the scale Ba0.6Sr0.4TiO3 þ 0.5 wt-% MgCo2(VO4)2 composite represents the highest relative permittivity (Qf ¼ 300 GHz, with er ¼ 2763). Figure 17 shows the variation of tf with relative permittivity. In general the materials with lower relative permittivity show negative tf and high permittivity materials have a positive tf. The Bi6Ti5TeO22 has the highest temperature variation of resonant frequency (Qf ¼ 220 GHz, with er ¼ 350 and tf of þ2600 ppm/uC) and BaNb2O6 (hexagonal) has the highest negative tf (Qf ¼ 4000 GHz, with er ¼ 42 and tf of 2800 ppm/uC); however, this is a question of optimisation since there are several ways to tune the tf value such as by forming composites with positive and negative tf materials. The table (supplementary file) shows the availability of materials with almost any desired relative permittivity especially in the range of 5–100; however, simultaneously satisfying a desired relative permittivity with excellent Qf and tf values is difficult. Tailoring microwave dielectric properties Microwave dielectric properties can be tailored by chemical methods like doping, slight deviations from stoichiometry, or the formation of composites of dielectrics with oppositely signed tf values.60–66 Luiten et al.64,67 used paramagnetic effects of impurity ions to compensate for the permittivity–temperature dependence (t1), which is related to tf by equation (28); but this technique is not applicable at cryogenic temperatures or even room temperature because of the finite energy gap of paramagnetic resonance. Hartnett et al.68,69 proposed a method of compensating for the frequency–temperature dependence (tf) of high-Q monolithic sapphire resonators near liquid-nitrogen temperatures by doping single-crystal sapphire with Ti3þ ions. Breeze et al.70 reported a new method of achieving temperature compensation by coating a film of TiO2 on the surface of an alumina disc. The composite resonators obtained by firing at 1400uC showed a temperature compensation depending on the volume fraction of TiO2. Materials having negative tf are usually tailored 17 Variation of the coefficient of temperature variation of the resonant frequency as a function of relative permittivity Low-loss dielectric ceramic materials by adding TiO2, CaTiO3 or SrTiO3, all of which have high positive tf values.66,71–79 Similarly, positive tf materials can be tailored by adding negative tf materials.80,81 For example, the addition of about 17 mol% TiO2 in ZnAl2O4 results in a nearly zero tf as shown in Fig. 18. The quality factor and the relative permittivity also vary with TiO2 content. Of course, this technique can only be used when the additive material does not react with the parent material. It is also possible to tailor tf by stacking positive and negative tf resonators. The resultant properties depend on the volume fraction or thickness of the two different resonator materials.82–85 Fig. 19 shows a sketch of the stacking and the variation of tf as a function of the volume fraction of the negative tf (266 ppm/uC) resonator Sr(Y1/2Nb1/2)O3 in a composite with the positive tf (þ 78 ppm/uC) resonator Ba5Nb4O15. It was reported83 that the properties slightly change on reversing the bottom and top resonator samples. The samples can be joined using low-loss adhesives, but the use of adhesives lowers the quality factor.84 It is also possible to tailor properties by solid-solution formation between positive and negative tf materials provided they have similar crystallographic structures.86–91 If the end members have different crystal structures, then a phase transition at some intermediate composition may result in sudden change in the 18 Variation of dielectric properties of (12x)ZnAl2O4–xTiO2 as a function of x a tf b quality factor. Inset of figure b shows variation of resonant frequency with TiO2 content (after Ref. 66) International Materials Reviews 2015 VOL 60 NO 7 403 Sebastian et al. Low-loss dielectric ceramic materials Downloaded by [Penn State University] at 11:01 26 November 2015 dielectric properties.87 Fig. 20 shows the variation of the dielectric properties of (12x)CaTiO3–xNdAlO3 solid solution. A zero tf is observed for x ¼ 0.3. The solid solution can be represented by Ca12xNdxTi12xAlxO3. A slight non-stoichiometry is also sometimes found to improve the densification and microwave dielectric properties.92–96 The presence of vacancies can facilitate atomic diffusion and thereby increase densification. For example, slight Ba or Mg deficiencies in Ba(Mg1/3Ta2/3)O3 is found to improve densification, order parameter and quality factor, as shown in Fig. 21. The addition of suitable dopants can improve the microwave dielectric properties, and a study of the dielectric table reveals that the microwave dielectric 19 a Schematic sketch of stacking of positive and negative tf resonators having varying thickness b Variation of tf of Ba5Nb4O15 ceramic as a function volume fraction of stacked Sr(Y1/2Nb1/2)O3ceramic (after Ref. 83) 20 Variation of dielectric properties a tf b relative permittivity and c Qf as a function x in (12x)CaTiO3–xNdAlO3 solid solution (after Ref. 92) 404 International Materials Reviews 2015 VOL 60 NO 7 21 a Variation of bulk density and order parameter as a function of x in Ba(Mg0.33xTa0.67)O3 ceramics b Variation of the relative permittivity and tf as a function of x in Ba(Mg0.33x Ta0.0.67)O3 ceramics (after Ref 95) Downloaded by [Penn State University] at 11:01 26 November 2015 Sebastian et al. properties can also be tailored to some extent by suitable chemical substitution.97–102 Usually, the dopant partially substitutes at appropriate sites in the parent material. For example, it is reported that the quality factor reaches a maximum when the ionic radius of the dopant is close to the average ionic radius of the B-site ion in Ba(Mg1/3Ta2/3)O3 (BMT) and Ba(Zn1/3Nb2/3)O3 (BZN) ceramics.103,104 In BMT the Qf reaches a maximum when the ionic radius of the dopant is between 0.6 and 0.7 Å, and the weighted average ionic radius of Mg and Ta is 0.653 Å. Figure 22a shows the variation of Qf in BMT ceramics as a function of the concentration of various dopants. A very small amount of dopant is found to improve the quality factor, with slight changes in relative permittivity and tf. Figure 22b shows the variation of Qf in BMT as a function of the ionic radius of the dopant. Many of the materials are difficult to densify even sintering at high temperatures, and such materials are usually densified by adding a small amount of lowmelting-temperature compounds or glasses. The high sintering temperatures can also be lowered and the densification improved by liquid-phase sintering via the addition of low-melting-temperature compounds such as V2O5, Bi2O3, CuO, LiF, MgF2, CuO, B2O3, Nb2O5, Li2CO3, BaCuB2O5, MoO3, Li2WO4, CuV2O6, PbO, etc. Low-loss dielectric ceramic materials and glasses such as Li2O–B2O3–SiO2, Li2O–MgO–ZnO– B2O3–SiO2, MgO–B2O3–SiO2, ZnO–B2O3, CaO–B2O3– SiO2, B2O3–P2O5, MgO–CaO–Al2O3–SiO2, ZnB2O4, Bi2O3–B2O3, Al2O3–B2O3–SiO2, ZnO–B2O3–SiO2, BaO–B2O3–SiO2, Bi2O3–B2O3–ZnO–SiO2, PbO–B2O3, PbO–B2O3–SiO2, Li2O–Zn–B2O3, Li2O–B2O3–SiO2, Bi2O3–B2O3–ZnO–SiO2, BaO– Li2O–MgO–B2O3, B2O3–SiO2–CaO–Al2O3, BaO–B2O3–Li2O–CuO–, PbO– Al2O3–SiO2, La2O3–ZnO–B2O3, PbO–Bi2O3–B2O3– ZnO–TiO2. The addition of a small amount of several glasses is found to be effective in lowering the sintering temperature and improving microwave dielectric properties of BMT ceramics. Figure 23 shows the effect of some selected glasses on the Qf and tf of BMT ceramics. Although the addition of larger amounts of glass considerably lowers the sintering temperature, it also degrades the microwave dielectric properties. In general glasses have negative tf values, and glass addition improves the tf of materials with positive values of tf. Compounds like CeO2, MnCO3, SnO2, NiO, ZnO, WO3, TiO2, Yb2O3, ZrO2 etc. have also been used to aid solid-state sintering and improving the dielectric properties.103,106–112 Partial substitution by elements with higher dielectric polarisability can increase the relative permittivity.113 For example, as shown in Fig. 24, substituting 44% of the Sr2þ(ai ¼ 4.24 Å3)114 in Sr9Ce2Ti12O36 with Pb2þ(ai ¼ 6.58 Å3)114 increases the relative permittivity from 183 to about 800.113 In several complex perovskites an order–disorder transition is found to affect the microwave dielectric properties. Improvement in ordering by annealing or doping is found to improve the quality factor considerably.103–105 The purity and origin of the initial raw materials can also influence the phase formation, densification and microwave dielectric properties. The presence of porosity decreases the relative permittivity and a correction for porosity can be performed using mixture rules, as discussed in section Correction for Porosity. The presence of porosity considerably increases the loss tangent for otherwise dense ceramics, as shown in Fig. 25 for alumina.115 A dense ceramic usually optimises the microwave dielectric properties. Figure 26 shows a typical microstructure of thermally etched dense ceria ceramic sintered at 1675uC. Ceria has a relative permittivity of 24 and Qf of 65 000 GHz.116 All types of defects contribute to extrinsic dielectric losses. For an ideal material, loss is mainly a manifestation of the interaction of the phonons with microwaves, hence it is possible to improve the quality factor by suppressing the phonons by cooling the ceramics. Figure 27 shows the variation of the quality factor of ceria ceramic as a function of cooling. The quality factor reaches a maximum of about 10 000 at 6 GHz at 50 K. Applications of low-loss dielectric ceramics Materials for LTCC applications 22 a Variation of the quality factor of Ba(Mg1/3Ta2/3)O3 ceramics as a function of the dopant concentration b Variation of the quality factor of Ba(Mg1/3Ta2/3)O3 ceramics as a function of the dopant ionic radii (after Ref. 98) High and low temperature co-fired ceramics, HTCC and LTCC respectively, have created a new generation of small and lightweight electronic multilayer components with application area such as capacitors and microwave products. The LTCC tapes are fabricated from suitable choice of low-temperature sinterable dielectric materials. International Materials Reviews 2015 VOL 60 NO 7 405 Sebastian et al. Low-loss dielectric ceramic materials Downloaded by [Penn State University] at 11:01 26 November 2015 24 Variation of the relative permittivity as a function of Pb substitution for Sr in Sr9Ce2Ti12O36 ceramics (after Ref. 107) 25 Variation of loss tangent as a function of porosity in alumina (after Ref. 108) 23 Variation of a quality factor b tf of BMT as a function of glass content (after Ref. 100) Currently these LTCC substrates are being developed by industrial organisations like DuPont, Ferro and Motorola. The developmental activities (basic, applied and product development) of dielectric materials have shown substantial increase in the last decade resulting in a variety of dielectric materials for choosing the required compositions with respective properties. 406 International Materials Reviews 2015 VOL 60 NO 7 Ceramic composition that has a sintering temperature from 700 to 950uC can be categorised to belong to LTCCs. The upper limit comes from the requirement that the tapes made of it should densify in co-firing with high conductive electrode material like Ag or Cu. At lower sintering temperatures than 700uC, other electrode material like Al, Pd or different mixtures should be selected and the resistance of the electrodes increases highly. One must note in order to enable multilayer co-fired structures, the tapes made of the LTCC have to be co-fired with Ag or Cu pastes without excess reactions. Low-melting and low-loss glasses are usually added to low-loss dielectric ceramics in order to decrease the sintering temperature below the melting point of the silver electrode. The addition of glasses degrades the dielectric and mechanical properties. Another option is to add sintering aids which is the most common way with LTCCs having high relative permittivity. The Table includes several compositions such as vanadates, telleurates, tungstates, molybdnates and phosphates based on Li, Mg, suitable for glass free LTCC applications and several ceramic glass composites. The reader is referred to the review on LTCC for more details in reference.3 Sebastian et al. Downloaded by [Penn State University] at 11:01 26 November 2015 26 SEM microstructure of thermally etched ceria sintered at 16758C (courtesy P S Anjana) 27 Variation of quality factor of ceria on cooling (after Ref. 109) It may be noted that many of these reported LTCC materials are not prepared in tape form and the reactivity with electrode, thermal expansion, thermal conductivity, etc. are not reported in the literature. Although several glass free LTCC materials are available, tapes of very few glass free materials are reported in the literature.117,118 The important characteristics required for LTCC applications are as follows: (a) relative permittivity erw4 (b) tand v1022 at 5 GHz (c) tf in the range of 210 to þ10 ppm/uC (d) no reactivity with the electrode materials (e) coefficient of linear thermal expansion less than 20 ppm/uC or matching with that of silicon (f) high thermal conductivity. Materials for ULTCC applications There is a clear need for electroceramic compositions feasible for co-firing with organic or semiconductive structures expecting sintering temperatures less than 650uC using aluminium or less than 400uC using nano silver ink electrodes. In semiconductors, metal electrode should be deposited on top of the dielectric layers with Low-loss dielectric ceramic materials low temperature process. Additionally, multilayer packages similar nowadays to those made by LTCC technology but with much lower sintering temperature would enable co-firing of semiconductor devices into the package. In the recent decade several electroceramic compositions with sintering temperature below 700uC have been reported as shown in the Table. These materials fall in two categories. The first one (category II) covers compositions having sintering temperature over 400 up to 700uC. This category is justified since in these temperatures only Al, Pd or different metal mixture electrodes with relative low conductivity can be used. These ULTCC II category materials can be used on some metal, glass or ceramics substrates, but their feasibility for real multilayer applications is somewhat limited although there are many interesting compositions like Li2Mo4O12 sintered at 630uC has the highest Qf of 108 000 GHz with er ¼ 8.8 and tf ¼ 289,119 Zn2Te3O8 þ 30 wt-% TiTe3O8 sintered at 610uC has the lowest tf of 3 ppm/uC with er ¼ 19.8 and Qf ¼ 50 000 GHz.120 The [(Li0.5Bi0.5)x Bix][MoxV12x]O4 with x ¼ 0.098 when sintered at 650uC has the highest er of 81 with Qf of 8000 GHz, tf of 10 ppm/uC.121 The main application areas can be found at moderately low frequency areas. Important application fields could be multilayer capacitors and packages. The category I, with sintering temperature at 400uC or below, should be feasible with commercially available highly conductive nano silver inks in co-firing. These compositions have commonly ultra-low sintering temperature inherently. Although only very few compositions so far belonging to this category are reported, they will in the future provide great opportunities with integrated applications with semiconductor devices or on organic substrates. The NaAgMoO4 has the lowest sintering temperature of 400uC among the reported materials. It has a relative permittivity of 7.9 with Qf of 33 000 GHz and tf of 2120 ppm/uC.122 On the other hand, most of these ULTCC I materials are based on vanadates and molybdates which are soluble in water. This means the ultimate device needs suitable encapsulation. The research and development of ULTCC materials are still in their initial stage. There is an urgent need for developing materials with sintering temperature less than 400uC for future applications. Materials for dielectric resonators Ceramic dielectric resonators are widely used for commercial and military purposes from MHz frequencies up to 50 GHz. Their main advantages are compact size, temperature stability and high unloaded Q factor. Commonly used products are dielectric resonator oscillators (DROs), low-loss filters and combiners, and in unmetallized form are intended to operate in the TE01d mode. With different kind of cavity shielding or metallic coating, the performance of the DR is adjusted for the product demands. In the case of DROs a low phase noise is needed and thus materials with high-Q factor are used. In commercial DROs, the ceramic resonators have relative permittivity from 20 up to 50 with Q f values as high as 100 000 GHz. The dielectric properties are commonly measured in the frequency range from 2 GHz up to 10 GHz. Regardless of the application, thermal stability of the resonant frequency (210vTf v10 ppm/uC) is desirable. International Materials Reviews 2015 VOL 60 NO 7 407 Sebastian et al. Low-loss dielectric ceramic materials Materials for dielectric ceramic antennas Downloaded by [Penn State University] at 11:01 26 November 2015 The ceramic dielectric resonators are often enclosed inside metal cavities to confine radiation and to maintain a high-quality factor which is important for filter and oscillator applications. When the metallic shield is removed and with suitable feeding to excite appropriate mode, the resonators could become efficient radiators. McAllister and Long123 proposed the use of resonators for antenna applications. For details of dielectric resonator antennas (DRA), the reader is referred to the recent reviews.124–126 There is no inherent conductor loss in dielectric resonators which leads to high radiation efficiency. Simple coupling schemes can be used for DRA to most of the transmission lines used in microwave and millimetre-wave frequencies. Experimental and theoretical studies are extensively done on DRA with different shapes or geometries such as cylindrical, rectangular, circular, ring, conical, hemispherical and square-shaped structures. The dimension of the resonator is related to free space wavelength and er by equation (2) and by choosing a high er, the size of the DRA can be significantly reduced at the expense of bandwidth. The operating band width can be varied for a wide frequency range by suitably selecting the resonator parameters and a band width of 117% have been reported.127 The lowest frequency of DRA reported is 55 MHz128 and the highest 94 GHz.129 DRA’s with dimensions ranging from a few millimetre with er in the range of 6–100 have been reported.124–126 Very thin (v4 mm) structures are needed especially when integrated to portable terminals. Ceramic dielectric materials are widely used for GPS patch antennas leading to high performance and miniaturisation. Ceramic antennas have also been proposed for multi-purpose targets like machine-to-machine communication.130 They are supposed to operate in the Zigbee, ISM and cellular bands including LTE in the frequency band between 700 and 2500 MHz. Ceramic chip antenna is calculated to provide 80% reduction in PCB space for 2.45 GHz applications. Materials for millimetre-wave applications The millimetre-wave radio spectrum is expected to be used in the future (e.g. 5G networks) since higher carrier frequencies are possible as compared to the current systems, such as 4G and Wi-Fi. For millimetre-wave applications the relative permittivity should be approximately in the range of 6–20 having very highquality factors greater than 75 000 GHz with temperature stable dielectric properties. High permittivity materials in general have lower quality factors and also have the problem of fabricating extremely small sized resonators. ZnAl2O4–TiO2-, Mg2SiO4-, Mg4Ta2O9- and Al2O3-based materials are some of the examples for possible millimetre-wave communication in radars, space, 5G and military applications. Materials for future applications and conclusions The DR table indicates a large number of materials with very useful microwave dielectric properties. However, important emerging technologies will require seamless co-firing with plastics or paper substrates (printed electronics) or semiconductor devices. The DR table 408 International Materials Reviews 2015 VOL 60 NO 7 exhibits about 120 materials with Ultra Low Sintering Temperature (ULTCC – sintering temperature less than 700uC) and with materials like NaAgMoO4 sintering temperature even lower than 400uC. However, further research is required to enable low-loss dielectric microwave ceramics to integrate with plastics and feasible with nano silver and other electrode materials. Recently room temperature curable silica ink has been screen printed on flexible mylar substrate for printed applications.131 The screen printed silica has a relative permittivity of 2.4 and tand of 0.003 at 15 GHz. Another field of application, the health care systems or monitoring, requires suitable antenna materials for bio-implantable communication devices.132 Apatite-type materials with reasonably good microwave dielectric properties are reported133 (see the Table). However, their biocompatibility needs to be investigated for practical applications. One interesting field for microwave ceramics is lowloss polymer–ceramic dielectric composites reported for antenna and printed circuit board applications.1 However, they are rigid and not bendable or stretchable especially if the loading level of ceramic is high. Flexible, bendable and stretchable dielectrics which can cover even curved surfaces are important for applications in electronic control systems, consumer electronics, heart pacemakers, body worn antenna, etc. The requirements for a material to be used as a flexible dielectric waveguide are mechanical flexibility, high relative permittivity, low dielectric loss, high thermal conductivity, low coefficient of thermal expansion (CTE), etc. Recently low-loss ceramic-filled butyl rubber and silicon rubberbased composites have been reported,134–136 but further work is needed for device optimisation. Several applications areas like the semiconductor industry are in constant need for low-loss materials with ultra-low relative permittivity (low k materials) to reduce RC signal delay. Lowering the relative permittivity decreases power consumption and reduces crosstalk between nearby interconnects. Silica, which has the lowest permittivity of about 4.0, is commonly used as the low k material. Further decrease of the permittivity can be achieved by introducing porosity. However, the presence of porosity degrades the mechanical properties. Fluorination of silica (SiOF) lowered the permittivity to about 3.6.138 SiCOH with k of about 2.4 have also been reported.137 Recently several organic or hybrid dielectric materials have been developed137 with even lower permittivities, but they are not suitable for very large-scale integration (VLSI) chips because of their poor chemical, mechanical and thermal properties. There remains a need to develop materials with lower relative permittivities and good mechanical, chemical and thermal properties in order to increase the signal speed. As a conclusion, the study of the DR table reveals that many tantalates, niobates, titanates, silicates, tungstates, molybdanates, vanadates or tellurates based on alkali earth metal and rare earths show low dielectric loss. It seems that most of the low-loss dielectric microwave ceramic materials have an octahedral or tetrahedral arrangement of atoms. However, further investigation, including especially spectroscopic and XRD studies, is needed to understand the relationship between chemical bonding, lattice vibrations, atomic coordination, secondary phases, impurities and microwave dielectric properties. Such studies would be useful for finding new Sebastian et al. low-loss dielectric ceramic compositions for present and future applications. Attempts should be also done to lower the cost of production of microwave materials with emphasis on use of environment friendly materials with the possibility of recycling.138 In the near future, the new emerging communication applications like 5G network machine-to-machine connection and IoT will need novel dielectric ceramics with feasible component fabrication technologies. This means that the low-loss microwave ceramics will continue to be an active area of research in years to come. The future will show their importance for improved performance with cost-efficient and miniaturised devices. The operational details of 5G networks and the IoT are still not available and hence the material requirements are yet to be determined. Acknowledgement Downloaded by [Penn State University] at 11:01 26 November 2015 The authors are grateful to European Research Council (ERC project) and the US National Science Foundation (DMR 1052788) for financial support. References 1. M. T. Sebastian and H. Jantunen: ‘Polymer-ceramic composites of 0-3 connectivity for circuits in electronics: a review’, International J. Appl. Cer. Techn., 2010, 7, 415–434. 2. M. T. Sebastian: ‘Dielectric materials for wireless communication’, 2008, Oxford, UK, Elsevier. 3. M. T. Sebastian and H. Jantunen: ‘Low loss dielectric materials for LTCC applications. A review’, Int. Mater. Rev, 2008, 53, 57–90. 4. O. V. Karpova:""' ‘On an absolute method of measurement of dielectric properties of a solid using a Л – shaped resonator’, Sovt. Phys., 1959, 1, 220–228. 5. B. W. Hakki and P. D. Coleman: ‘A dielectric resonator method of measuring inductive capacities in the millimeter range’, IRE Trans. Microwave Theory Tech, 1960, 8, 402–410. 6. W. Courtney: ‘Analysis and evaluation of a method of measuring complex permittivity and permeability of microwave materials’, IEEE Trans. Microwave Theory & Tech., MTT-, 1970, 476, 485. 7. S. B. Cohn and K. C. Kelly: ‘Microwave measurements of high dielectric constant materials’, IEEE Trans. Microwave Theory Tech., 1966, 14, 406–410. 8. Available at: http://coen.boisestate.edu/rickubic/files/2012/05/ HakCol.exe, 2013. 9. R. Ubic: ‘Software for calculating permittivity of resonator’, in HakCol & ErCalc’, Ceramic Transactions, (Ed. J. Akedo, T.-Y.Tseng, X.M. Chen and H.-T. Lin), 2014, 65–75. 10. Y. Kobayashi and S. Tanaka: ‘Resonant modes of dielectric rod resonator short circuited at both ends by parallel plate conducting plates’, IEEE Microwave Theory Tech., 1980, 28, 1077–1085. 11. D. L. Rebsh, D. C. Webb, R. A. Moore and J. D. Cawlishaw: ‘Mode chart for accurate design of cylindrical dielectric resonators’, IEEE Microwave Theory and Tech., 1965, 468–469. 12. IEC 61338, section1-1,1-2 and 1-3 (1996-2003): ‘Wave guide type dielectric resonators. Part 1. General information and test conditionssection 3. Measurement method of complex permittivity for dielectric resonator materials at microwave frequency’, International Electrotechnical Commission (IEC), 2005, Geneva, Switzerland, . 13. D. Hennings and P. Schnabel: ‘Dielectric characterization of Ba2Ti9O20 type ceramics at microwave frequencies’, Philips Res. Rept., 1983, 38, 295–311. 14. T. Itoh and R. S. Rudokos: ‘New method for computing the resonant frequencies of dielectric resonators’, IEEE trans. Microwave Theory Tech., 1977, 25, 52–54. 15. D. Kajfezz and P. Guillon: ‘Dielectric Resonators’, 1986, Dedham, MA, Artech House Inc. 16. C. M. Garnet: ‘Colours in metal glasses and metal films’, Phil. Trans. Royal Society of London. Series A: Mathematical,, Physical, and Engineering Sciences, 1904, 3, 385–420. 17. K. Lichtenecker: ‘Die Dielektrizittskonstante natrlicher und knstlicher Mischkrper’, Physikalische Zeitschrift, 1926, XXVII 115–158. 18. C. J. F. Bötcher: ‘The Dielectric Constant of Crystalline Powders, Recueil des Travaux Chimiques des Pays-Bas’, Journal of the Royal Netherlands Chemical Society, 1945, 64, 47–51. Low-loss dielectric ceramic materials 19. D. A. G. Bruggeman: ‘Calculation of various physics constants in heterogenous substances I Dielectricity constants and conductivity of mixed bodies from isotropic substances’, Annalen der Physik., 1935, 24, 636–664. 20. O. Wiener: ‘Theorie d. Mischkörpers F.D Feld d. Station. Strömung. I. Herkunft und Stellung der Aufgabe’, Abhandlungen Math. Phys. Klasse Sächs. Akademie der Wissenschaften (Leipzig), 1912, 32, 509–523. 21. J. A. Stratton: ‘Electromagnetic Theory’, 1941, New York, McGraw-Hill. 22. J. A. Reynolds and J. M. Hough: ‘Formulae for Dielectric Constant of Mixtures’, Proceedings of the Physical Society, 1957, 70, 769–775. 23. J. Krupka: ‘Frequency domain complex permittivity measurements at microwave frequencies’, Meas. Sc. & Tech., 2005, 16, R1–R16. 24. P. Whelss and D. Kajfezz: ‘The use of higher order resonant modes in measuring the dielectric constant of dielectric resonators’, IEEE. Symp. Digest. St. Louis, 1985, MTT-S, 473–476. 25. Y. Kobayashi and M. Katoh: ‘Microwave measurement of dielectric properties of low loss materials by the dielectric rod resonator method’, IEEE Microwave Theory and Tech., 1985, 33, 586–592. 26. M. W. Pospieszalski: ‘On the theory and applications of the dielectric post resonator’, IEEE Microwave Theory and Tech., 1977, 25, 697–7000. 27. H. Takagi, N. Funnami, H. Tamura and K. Akino: ‘Measurement accuracy of complex permittivity in the dielectric rod resonator method’, Jpn. J. Appl. Phys., 1992, 31, 3269–3271. 28. Y. Kobayashi and H. Tamura: ‘Round robin test on a dielectric resonator method for measuring complex permittivity at microwave frequency’, ICIE Trans. on Electronics, 1994, E77-C, 882–887. 29. J. Krupka, K. Derzakowski, B. Riddle and J. B. Jarvis: ‘A dielectric resonator for measurements of complex permittivity of low loss materials as a function of temperature’, Meas. Sci. Techn., 1998, 9, 1751–1756. 30. K. Leong, J. Mazierska and J. Krupka: ‘Measurement of unloaded Q factor of transmission mode dielectric resonators’, IEEE Symposium Digest, 1997, 97, 1639–1642. 31. H. Takamura, H. Matsumoto and K. Wakino: ‘Low temperature properties of microwave dielectrics’, Proc. 7 th Meeting on Ferroelectric Materials and their Applications. Jpn. J. Appl. Phys., 1989, (Supl. S-28-2), 21–23. 32. J. Krupka, W-T. Huang and M-J. Tung: ‘Complex permittivity measurements of low loss microwave ceramics employing higher order quasi TEonp modes excited in a cylindrical dielectric sample’, Measurement Science & Technology, 2005, 16, 1014–1020. 33. M. Valant, D. Suvorov and S. Macek: ‘Measurement error analysis in the determination of microwave dielectric properties using cavity reflection method’, Ferroelectr., 1996, 176, 167–177. 34. A. P. S. Khanna and G. Garault: ‘Determination og loaded, unloaded and external quality factors of a dielectric resonator coupled to a microstrupline’, IEEE Trans. Microwave Theory and Tech., 1983, 31, 261–264. 35. J. R. Wait: ‘Electromagnetic Whispering Gallery Modes in a dielectric rod’, Radio Sci., 1967, 2, 1005–1017. 36. D. Cros and P. Guillon: ‘Whispering Gallery dielectric resonator modes for w-band devices’, IEEE Microwave Theory and Tech., 1990, 38, 1667–1674. 37. J. Krupka, D. Cros, M. Aubourg and P. Guillon: ‘Study of Whispering Gallery Modes in anisotropi single crystal dielectric resonators’, IEEE Trans. Microwave Theory and Tech., 1994, 42, 56–61. 38. E. N. Ivanov, D. G. Blair and V. I. Kalinichev: ‘Approximate approach to the design of shielded dielectric disk resonators with whispering Gallery Modes’, IEEE Trans. Microwave Theory and Tech., 1993, 41, 632–637. 39. J. Krupka, K. Derzakowski, A. Abramowicz, M. E. Tobar and R. G. Geyer: ‘Whispering gallery modes for complex permittivity measurements of ultra low loss dielectric materials’, IEEE Microw. Theory Tech., 1999, 47, 752–759. 40. J. Krupka, A. P. Gregonry, O. C. Kochard, R. N. Clarke, B. Riddle and J. Baker-Jarvis: ‘Uncertainity of complex permittivity measurement by split post dielectric resonator techniques’, J. Eur. Ceram. Soc., 2001, 21, 2673–2676. 41. J. Krupka, S. Gabelich, K. Derzakowski and B. M. Pierce: ‘Comparison of split post dielectric resonator and ferrite disc resonator techniques for microwave permittivity measurements of polycrystalline ytterium iron garnet’, Measur. Sci. Tech., 1999, 10, 1004–1008. International Materials Reviews 2015 VOL 60 NO 7 409 Sebastian et al. Low-loss dielectric ceramic materials Downloaded by [Penn State University] at 11:01 26 November 2015 42. G. Kent: ‘An evanescent mode tester for ceramic dielectric substrates’, IEEE Microwave Theory and Tech., 1988, 36 1451–1454. 43. J. Krupka: ‘Precise measurement of complex permittivity of dielectric materials at microwave frequencies’, Mater. Chem. Phys., 2003, 79, 195–198. 44. N. L. Conger and S. E. Tung: ‘Measurement of dielectric constant and loss factor of powder materials in the microwave region’, Rev. Sci. Instr., 1967, 38, 384–386. 45. S. O. Nelson and A. W. Kraszewski: ‘Sensing pulverized material mixture proportions by resonant cavity measurements’, IEEE Trans. Instr. Meas., 1998, 47, 1201–1204. 46. S. O. Nelson and P. G. Bartley: ‘Open ended coaxial line permittivity measurements on pulverised sample’, IEEE Trans. Meas., 1998, 47, 133–137. 47. H. Ebara, T. Inoue and O. Hashimoto: ‘Measurement method of complex permittivity and permeability of a powdered material using waveguide in microwave band’, Sci. Tech. Adv. Mater., 2006, 7, 77–83. 48. M. Tuhkala, J. Juuti and H. Jantunen: ‘Method to characterize dielectric properties of powdery substances’, J. Appl. Phys., 2013, 114, 014108. 49. D. L. Gershon, J. P. Carmel, T. M. Antonsen and R. M. Hutcheon: ‘Open ended coaxial probable for high temperature and broad band dielectric measurements’, IEEE Trans. Micro. Theo. Tech., 1999, 47, 1640–1648. 50. K. Murata, A. Hanawa and R. Nozaki: ‘Broad band complex permittivity measurement technique of material with thin configuration at microwave frequencies’, J. Appl. Phys., 2005, 98, 084107. 51. H. Looyenga: ‘Dielectric constants of mixtures’, Physica, 1965, 31, 401–406. 52. M. Tuhkala, J. Juuuti and H. Jantunen: ‘An indirectly coupled open ended resonator applied to characterize dielectric properties of MgTiO3-CaTiO3 powders’, J. Appl. Phys., 2014, 115, 184101. 53. M. Tuhkala, J. Juuti and H. Jantunen: ‘Use of an open ended coaxial cavity method to characterize powders substances exposed to humidity’, Appl. Phys. Lett., 2013, 103, 142907. 54. M. Tuhkala, J. Juuti and H. Jantunen: ‘Determination of complex permittivity of surfactant treated powders using an open ended coaxial cavity resonator’, Powder Technology, 2014, 256, 140–145. 55. J. Molla, M. Gonzalez, R. Villa and A. Ibara: ‘Effect of humidity on microwave dielectric losses of porous alumina’, J. Appl. Phys., 1999, 85, 1727–1730. 58. E. Schloman: ‘Dielectric losses in ionic crystals with disordered charge distributions’, Phys. Rev., 1964, 135, A413–A419. 57. V. S. Vinogradov: ‘Fiz. Tverd. Tela. 2(1960)2622. (English Translation)’, Sovt. Phys. Sol. St., 1961, 2, 2338. 58. L. Li and X. M. Chen: ‘Frequency dependent QF value of microwave dielectric ceramics’, J. Amer. Ceram. Soc., 2014. 59. G. L. Gurevich and A. K. Tagantsev: ‘Intrinsic dielectric los in crystals’, Adv. Physics, 1991, 40, 719–767. 60. K. S. Hong, I. T. Kim and S. J. Yoon: ‘Effect of nonstoichiometry and chemical inhomogeneity on the order-disorder phase transformation in the complex perovskite compound Ba(Ni1/3Nb2/3)O3 and Ba(Zn1/3Nb2/3)O3’, J. Mater. Sci., 1995, 30, 514–521. 61. L. A. Khalam and M. T. Sebastian: ‘Effect of cation substitution and non-stoichiometry on the microwave dielectric properties of Sr(B1/2B’’1/2)O3 [B0¼lanthanides] ’, J. Am. Ceram. Soc., 2007, 89, 3689–3695. 62. S. G. Santiago, R. T. Wang and G. J. Dick: ‘Improved performance of a temperature compensated LN2 cooled sapphire oscillator’ ‘IEEE 1995, San Fransisco, Ca, USA. Proc. IEEE Int. Freq. Control Symp.1995’, 397–400. 63. J. G. Hartnett, M. E. Tobar, E. N. Ivanov and D. Cross: ‘High frequency compensated sapphire/rutile resonator’, Electron. Lett., 2000, 36, 726–727. 64. A. N. Luitenen, A. G. Mann and D. G. Blair: ‘Paramagnetic susceptibility and permittivity measurements at microwave frequencies in cryogenic sapphire resonators’, J. Phys. D. Appl. Phys., 1996, 29, 2082–2090. 65. J. Krupka, D. Cross, A. N. Luitenen and M. E. Tobar: ‘Design of very high Q sapphire resonators’, Electron. Lett., 1996, 32, 670–671. 66. K. P. Surendran, N. Santha, P. Mohanan and M. T. Sebastian: ‘Temperature stable low permittivity ceramic dielectrics in (1-x)ZnAl2O4-xTiO2 system for microwave substrate applications’, Eur. Phys. J.B., 2004, 41, 301–306. 410 International Materials Reviews 2015 VOL 60 NO 7 67. A. N. Luiten, A G Mann, N. J. Mc Donald and D. G. Blair: ‘Latest results of the UWA cryogenic sapphire oscillator’ ‘IEEE 1995, San Fransisco, Ca, USA. Proc. Freq. Control Symp.’,1995, 433–437. 68. J. G. Hartnett, M. E. Tobar, A. G. Mann, J. Krupka and E. N. Ivanov: ‘Temperature dependence of Ti3þ doped sapphire whispering gallery mode resonator’, Electron. Lett., 1998, 34, 195–196. 69. J. G. Hartnett, M. E. Tobar, A. G. Mann, E. N. Ivanov, J. Krupka and R. Geyer: ‘Frequency-temperature compensation in Ti4± doped sapphire whispering gallery mode resonators’, IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 1999, 46, 993–999. 70. J. Breeze, S. J. Penn, M. Poole, N. Mc and N. Alford: ‘Layered Al2O3–TiO2 composite dielectric resonators’, Electron. Lett., 2000, 36, 883–884. 71. D. Thomas and M. T. Sebastian: ‘Temperature compensated LiMgPO4: A new glass free low temperature cofired ceramic’, J. Amer. Ceram. Soc., 2010, 93, 3828–3831. 72. Y. Guo, H. Ohsato and K. Kakimoto: ‘Characterization and dielectric behaviour of willemite and TiO2 doped willemite at millimeterwave frequency’, J. Eur. Ceram. Soc., 2006, 26, 1827–1830. 73. K. P. Surendran, P. V. Bijumon, P. Mohanan and M. T. Sebastian: ‘(1-x) MgAl2O4-xTiO2 dielectrics for microwave and millimeter wave applications’, Appl. Phys., 2005, A81, 823–826. 74. S-F. Wang, C-Y. Huang and Y-L. Liu: ‘Effects of CaTiO3 and SrTiO3 additions on the microstructure and microwave dielectric properties of ultra-low fire TeO2 ceramics’, J. Amer. Ceram. Soc., 2010, 93, 3272–3277. 75. C-L. Huang and J-Y. Chen: ‘Low loss microwave dielectrics using SrTiO3 modified (Mg0.95Co0.05)2TiO4-ceramics’, J. Alloys & Compounds, 2009, 485, 706–710. 76. J. Guo, D. Zhou, L. Wang, H. Wang, T. Shao, Z. M. Qi and Xi Yao: ‘Infra red spectra, raman spectra, microwave dielectric properties and simulation for effective permittivity of temperature stable ceramics AMO4-TiO2 (A¼Ca, Sr)’, Dalton Transactions, 2013, 42, 1483–1491. 77. C-L. Huang, J-Y. Chen and C-C. Liang: ‘Dielectric properties and mixture behaviour of Mg4Nb2O9-SrTiO3 ceramic system at microwave frequeny’, J. Alloys & Compounds, 2009, 478, 554–558. 78. J-W. Choi, J-Y. Ha, S-J. Yoon, H-J. Kim and K. H. Yoon: ‘Microwave dielectric characteristics of 0.75(Al1/2Ta1/2)O20.25(Ti1-xSnx)O2 ceramics’, J. Eur. Ceram. Soc., 2003, 23, 2507–2510. 79. M. Hu, Y. Fu, C. Luo, H. Gu, W. Lei and Q. Shen: ‘Microstructure and microwave dielectric properties of xCa(Al0.5Nb0.5)O3 þ (1-x)SrTiO3 solid solution’, J. Amer. Ceram. Soc., 2010, 93, 3354–3359. 80. N. Ichinose and K. Mutoh: ‘Microwave dielectric properties in the (1-x) (Na1/2La1/2)TiO3-x(Li1/2Sm1/2)TiO3 ceramic system’, J. Eur. Ceram. Soc., 2003, 23, 2455–2459. 81. J-S. Kim, C. I. Cheon, H-J. Kang, C-H. Lee, K-Y. Kim, S. Nahm and J-D. Byun: ‘Crystal structure and microwave dielectric properties of CaTiO3-(Li1/2Nd1/2)TiO3-(Ln1/3Nd1/3) TiO3(Ln¼La, Dy) ceramics’, Jpn. J. Appl. Phys., 1999, 38, 5633–5637. 82. Y. Zhang, B. Fu, M. Hong, M. Xiang, L. Liu and T. Lang: ‘Microstructure and microwave dielectric properties of MgNb2O6- ZnTa2O6 composite ceramics prepared by layered stacking method’, J. Mater. Sc. Mater. Electr., 2014, 25, 5475–5480. 83. I. N. Jawahan, P. Mohanan and M. T. Sebastian: ‘A novel method of achieving temperature compensation by stacking positive and negative (f resonators’, Materials Science and Engineering, 2003, B97, 258–264. 84. L. Li and X. M. Chen: ‘Adhesive bonded Ca(Mg1/3Nb2/3)O3/ Ba(Zn1/3Nb2/3)O3 layered dielectric resonators with tunable temperature coefficient resonant frequency’, J. Amer. Ceram. Soc., 2006, 89, 544–549. 85. L. Li, X. M. Chen and X. Cheng: ‘Characterization of MgTiO3CaTiO3 layered microwave dielectric resonators with TE01d mode’, J. Amer. Ceram. Soc., 2006, 89, 557–561. 86. M. W. Lufasso, E. Hopkins, S. M. Bell and A. Llobet: ‘Crystal chemistry and microwave dielectric properties of Ba3MNb2-xSbxO9 (M¼Mg, Ni, Zn)’, Chem. Mater., 2005, 17, 4250–4255. 87. K. P. Surendran, M. R. Varma and M. T. Sebastian: ‘Microwave dielectric properties of RE1-xRE’xTiNbO6 [RE¼Pr, Nd, Sm; RE’¼Gd, Dy, Y] ceramics’, J. Am. Ceram. Soc., 2003, 86, 1695–1699. Downloaded by [Penn State University] at 11:01 26 November 2015 Sebastian et al. 88. J. Kato, H. Kagata and K. Nishimoto: ‘Dielectric properties of (Pb,Ca)(Me,Nb)O3 at microwave frequencies’, Jpn. J Appl. Phys., 1992, 31, 3144–3147. 89. K. Wakino, M. Murata and H. Tamura: ‘Far infrared reflection spectra of Ba(Zn,Ta)O3-BaZrO3 dielectric resonator material’, J. Amer. Ceram. Soc., 1986, 69, 34–37. 90. P. P. Ma, L. Yi, X. Q. Liu, L. Li and X. M. Chen: ‘Effects of post densification annealing upon microstructures and microwave dielectric characteristics in Ba((Co.06-x/2Zn0.4-x/2 Mgx)1/3Nb2/3)O3 ceramics’, J Amer. Cer. Soc., 2013, 96, 3417–3424. 91. B. Jancar, D. Suvorov, M. Valant and G. Drazic: ‘Characterization of CaTiO3-NdAlO3 dielectric ceramics’, J. Eur. Ceram. Soc., 2003, 3, 1391–1400. 92. C-S. Chen, C-C. Chou, W-J. Shih, K-S. Liu, C-S. Chen and I-N. Lin: ‘Microwave dielectric properties of glass ceramic composites for low temperature cofireable ceramics’, Mater. Chem. Phys., 2003, 79, 129–134. 93. D. A. Durilin, O. V. Ovchar and A. G. Belous: ‘Effect of nonstoichiometry on the structure and microwave dielectric properties of Ba1-x(Zn1/2W1/2)O3-x and Ba(Zn1/2-yW1/2)O3-y/2’, Inorganic Mater., 2011, 47, 313–316. 94. K. P. Surendran, M. T. Sebastian, P. Mohanan, R. L. Moreira and A. Dias: ‘Effect of stoichiometry on the microwave dielectric properties of Ba(Mg1/3Ta2/3)O3 dielectric ceramics’, Chem. Mater., 2005, 17, 142–151. 95. A. G. Belous, O. V. Ovchar, A. V. Krameeranko, B. Jancar, J. Bezjak and D. Suvorov: ‘Effect of nonstoichiometry on the structure and microwave dielectric properties of Ba(Co1/3Nb2/ 3)O3’, Inorganic Mater., 2010, 46, 529–533. 96. H. Wu and P. K. Davies: ‘Influence of non-stoichiometry on the structure and properties of Ba(Zn1/3Nb2/3)O3 microwave dielectrics. I. Substitution of Ba3W2O9’, J. Am. Ceram. Soc., 2006, 89, 2239–2249. 97. M. Terada, K. Kuwamura, I. Kagomiya, K. Kakimoto and H. Ohsato: ‘Effect of Ni substitution on the microwave dielectric properties of cordierite’, J. Eur. Ceram. Soc., 2007, 27, 3045–3048. 98. H-J. Lee, I-T. Kim and K. S. Hong: ‘Dielectric properties of AB2O6 comopunds at microwave frequencies (A¼Ca, Mg, Mn, Co, Ni, Zn and B¼Nb,Ta)’, Jpn. J. Appl. Phys., 1997, 36, L1318–L1320. 99. G. Sumesh and M. T. Sebastian: ‘Microwave dielectric properties of Ca[Li1/3A2/3)1-xMx]O3-d [A¼Nb,Ta and M¼Ti,Zr,Sn] complex perovskite’, in ‘Perovskite:Structure properties and Uses’, (ed. M. Borowski), 283–317; 2010, Huntington, NY, USA, Nova science Publishers Inc. 100. L. A. Khalam and M. T. Sebastian: ‘Low loss dielectrics in the Ca(B’1/2Nb1/2)O3 [B’¼lanthanides, Y] system’, J. Am. Ceram. Soc., 2007, 90, 1467–1474. 101. L. Fang, Q. Liu, Y. Tang and H. Zhang: ‘Adjustable dielectric properties of Li2CuxZn1-xTi3O8 (x¼0 to 1) ceramics with low sintering temperature’, Ceram. Intl., 2013, 38, 6431–6434. 102. X. Kuang, M. M. B. Allix, J. B. Claridge, H. J. Niu, M. J. Rossiensky, R. M. Ibberson and D. Iddles: ‘Crystal structure, microwave dielectric properties and AC conductivity of B-cation deficient hexagonal perovskites La5MxTi4-xO15 (x¼0.5, 1,: M¼Zn, Mg, Ga, Al)’, J. Mater. Chem, 2006, 16, 1038–1045. 103. K. P. Surendran, M. T. Sebastian, P. Mohanan and M. V. Jacob: ‘The effects of dopants on the microwave dielectric properties of Ba(Mg1/3Ta2/3)O3 ceramics’, J. Appl. Phys., 2005, 98, 094114. 104. M. R. Varma and M. T. Sebastian: ‘The effects of dopant additives on the microwave dielectric properties of Ba(zn1/3Nb2/3)O3 ceramics’, J. Eur. Ceram. Soc., 2007, 27, 2827–2833. 105. K. P. Surendran, P. Mohanan and M. T. Sebastian: ‘The effects of glass additives on the microwave dielectric properties of Ba(Mg1/3Ta2/3)O3 ceramics’, Solid St. Chem., 2004, 177, 4031–4046. 106. A. Ioachin, M. G. Banau, M. I. Toacsan, L. Nedelcu, D. Ghetu, H. V. Alexander, G. Toica, G. Annino, M. Cassettari and M. Martnelli: ‘Nickel doped (Zr0.8Sn0.2)TiO4 for microwave and millimeter wave applications’, Mater. Sci. Eng. B, 2014, 118, 205–209. 107. C. -L. Huang, C. -S. Hsu and R. -J. Lin: ‘Improved high-Q microwave dielectric resonator using ZnO andWO3-doped Zr0. 8Sn0.2TiO4 ceramics’, Mater. Res. Bull., 2001, 36, 1985–1993. 108. K. H. Yoon, Y. S. Kim and E. S. Kim: ‘Microwave dielectric properties of (Zr0.8Sn0.2)TiO4 ceramics with pentavalent additives’, J. Mater. Res., 1995, 10, 2085–2090. 109. W. -Y. Lin, R. F. Speyer, W. S. Hackenberger and T. R. Shrout: ‘Microwave properties of Ba2Ti9O20 doped with zirconium and tin oxide’, J. Am. Ceram. Soc., 1999, 82, 1207–1211. Low-loss dielectric ceramic materials 110. S. F. Wang, Y. F. Hsu, T. -H. Weng, C. -C. Chiang, J. P. Chu and C. -Y. Huang: ‘Effects of additives on the microstructure and dielectric properties of Ba2Ti9O20 microwave ceramics’, J. Mater. Res., 2003, 18, 1179–1187. 111. L. A. Khalam and M. T. Sebastian: ‘Microwave dielectric properties of Sr(B1/2Nb1/2)O3 [B¼La, Pr, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Y, Er, Yb and In] ceramics’, Int. J. Appl. Ceram. Tech., 2006, 3, 364–374. 112. J.-I. Yang, S. Nahm, C.-H. Choi, H.-J. Lee and H.-M. Park: ‘Microstructure and microwave dielectric properties of Ba(Zn1/3Ta2/3)O3 with ZrO2 addition’, J. Am. Ceram. Soc., 85, 165–168. 113. S. Kamba, M. Savinov, F. Luafek, O. Tkac, C. Kadlec, E. J. John, G. Subodh, M. T. Sebastian, M. Klementova, V. Bovtun, J. Pokorny, V. Goian and J. Petzelt: ‘Ferroelectric and incipent ferroelectric properties of a novel Sr9-xPbxCe2Ti2O36 (x¼0-9) ceramic system’, Chem. Mater., 2009, 21, 811–819. 114. N. W. Grimes and R. W. Grimes: ‘Dielectric polarizability of ions and the corresponding effective number of electrons’, J.Phys.: Condens. Matter, 1998, 10, 3029–3034. 115. N. McN Alford, J. Breeze, X. Wang, S. J. Penn, S. Dalla, S. J. Webb, N. Ljepojevic and X. Aupi: ‘Dielectric loss of oxide single crystals and polycrystalline analogues from 10 to 320 K’, J. Eur. Ceram. Soc., 2001, 21, 2605–2611. 116. N. Santha, M. T. Sebastian, P. Mohanan, N. Mc, N. Alford, K. Sarma, R. C. Pullar, S. Kamba, A. Pashkin, P. Samukhina and J. Petzelt: ‘Effect of doping on the dielectric properties of CeO2 in the microwave and sub-millimeter frequency range’, J. Amer. Ceram. Soc., 2004, 87, 1233–1237. 117. D. Thomas, P. Abhilash and M. T. Sebastian: ‘Casting and characterization of LiMgPO4 glass free LTCC tape for microwave applications’, J. Eur. Ceram. Soc., 2013, 33, 87–93. 118. P. Abhilash, M. T. Sebastian and K. P. Surendran: ‘Glass free, non-aqueous LTCC tapes of Bi4(SiO4)3’, J. Eur. Cer. Soc.(submitted). 119. D. Zhou, C. A. Rindall, H. Wang, L-X. Pang and Xi Yao: ‘Microwave dielectric ceramics in Li2O-Bi2O3-MoO3 system with ultra low sintering temperatures’, J. Amer. Cer. Soc., 2010, 93, 1096–1100. 120. S. F. Wang, Y-F. Hsu, Y-R. Wang and C-C. Sung: ‘Ultra low fire Zn2Te3O8-TiTe3O8 ceramic Composites’, J. Amer. Ceram. Soc., 2010, 94, 812–816. 121. D. Zhou, C. A. Rindall, H. Wang, L-X. Pang and X. Yao: ‘Ultra low firing high-k scheelite structures based on [(Li0.5Bi0.5)xBi1-x][MoxV1-x]O4 microwave ceramics’, J. Amer. Cer. Soc., 2010, 93, 2147–2150. 122. D. Zhou, L-X. Pang, Z-M. Qi, B-B. jin and Xi Yao: ‘Novel ultralow temperature cofired microwave dielectric ceramic at 400 degree and its chemical compatibility with base metal’, Scientific Reports, 2014, 4, 05980. 123. M. W. Mcallister and S. A. Long: ‘Rectangular dielectric resonator antenna’, IEEE Electronic Letters, 1983, 19, 218–219. 124. A. Petoso and A. Ittipiboon: ‘Dielectric resonator antennas. a historical review and the current state of the art’, IEEE antennas and Propagation Magazine, 2010, 52, 91–116. 125. A. Petosa: ‘Dielectric Resonator Antenna Handbook’, 2007, Norwood, Ma, Artech House. 126. D. Soren, R. Ghatak, R. K. Mishra and D. Poddar: ‘Dielectric resonator antennas: designs and advances’, Progr. Electromag. Res. B., 2014, 60, 195–213. 127. M. N. Jazi and T. A. Denidni: ‘Design and implemenatation of an ultrawideband hybrid skirt monople dielectric resonator antenna’, IEEE and Wireless Propagation Lett, 2008, 7, 493–496. 128. S. P. Kingley and S. G. OKeefe: ‘beam steering and monopulse processing of probe fed dielectric resonator antenna’, Proc. Rada Sonar Navigation, 1999, 3, 121–125. 129. J. Svedin, L. G. huss, d. karlen, P. enoksson and C. Rusu: ‘A micromachined 94 GHz dielectric resonator antenna for focal plane array applications’ ‘IEEE intl. Microwave symposium, Honolulu, IEEE MTT-S’, 1375–1378; 2007. 130. J. liflander: ‘Ceramic chip antennas vs. PCB trace antennas: a comparison’, ‘MPDIGEST. Feature article’; 2010, White paperwww.pulseelectronics.com/download/3721/g041/pdf. 131. J. Varghese, K. P. Surendran and M. T. Sebastian: ‘Room temperature curable silica ink’, Royal Soc. Chem. Adv., 2014, 4, 47701–4707. 132. A. K. Skirvervik: ‘Implantable antennas:The challenge of efficiency’, ‘IEEE Proc.7 th European conference on Antennas International Materials Reviews 2015 VOL 60 NO 7 411 Sebastian et al. Low-loss dielectric ceramic materials Downloaded by [Penn State University] at 11:01 26 November 2015 and Propagation (EUCAP 2013), Gothenberg, Published by IEEE’, 3617–3631. 133. D. Thomas, P. Abhilash and M. T. Sebastian: ‘Effect of isovalent substitutions on the microwave dielectric properties of Ca4La6(SiO4)4(PO4)2O2 apatite’, J Alloys & Compounds, 2013, 546, 72–76. 134. L. K. Namitha and M. T. Sebastian: ‘Microwave dielectric properties of flexible silicone rubber- Ba(Zn1/3Ta2/3)O3 composite substrates’, Mater. Res. Bull., 2013, 48, 4911–4916. 135. J. Chameswary and M. T. Sebastian: ‘Development of butyl rubber-rutile composites for flexible microwave substrate applications’, Ceram. Intl., 2014, 40, 7439–7448. 412 International Materials Reviews 2015 VOL 60 NO 7 136. J. Chameswary, L. k. Namitha, m. Brahmakumar and M. T. Sebastian: ‘Material characterization and microwave Substrat applications of alumina-filled butyl rubber composites’, Int. J. Appl. Ceram. Technol., 2014, 11, 916–926. 137. A. Grill, S. M. Gates, T. E. Ryan, S. V. Nguyen and D. Priyadarshini: ‘Progress in the development and understanding of advanced low k and ultralow k dielectrics for very large scale interconnects-state of the art’, Appl. Phys. Reviews, 2014, 1, 011306. 138. I. Reaney and D. Iddles: ‘Microwave ceramics for resonators and filters in mobile phone networks’, J. Amer. Ceram. Soc., 2006, 89, 2063–2072.