Startseite Design and Analysis of a Hybrid Solar and Vibration Energy Harvester
Artikel Öffentlich zugänglich

Design and Analysis of a Hybrid Solar and Vibration Energy Harvester

  • M Shafiqur Rahman und Uttam K. Chakravarty EMAIL logo
Veröffentlicht/Copyright: 17. Dezember 2019
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

The performance of the small-scale stand-alone energy harvesters can be improved by implementing a hybrid energy harvesting technique. This paper aims at presenting the design and characterization of a hybrid energy harvester that can simultaneously harvest energy from mechanical vibration and solar radiation by combining piezoelectric, electromagnetic, electrostatic, and photovoltaic mechanisms. The hybrid device consists of a small high-efficiency solar panel and a bimorph PZT cantilever beam having a cylindrical tip magnet and two sets of capacitors (comb electrodes) attached on two sides of an ASTM 6061 T-6 Aluminum substrate. All the transducing sections of the configuration are interconnected by a smart hybrid electric circuit having a common optimum load resistance, an energy storage, and a microcontroller to generate and store combined power output when subjected to transverse vibration and solar radiation. The initial bias-voltage input required for the electrostatic mechanism is either obtained from the photovoltaic system or taken from the storage through the microcontroller. Results for the maximum power output are obtained at the fundamental resonance frequency of the vibrating cantilever beam. As the hybrid design allows a combined power harvesting method, more power is generated with better conversion efficiency than those obtained by stand-alone mechanisms. In addition to the power calculation, the study includes a stress and fatigue analysis of the cantilever beam using the finite element method to investigate the stress-life criteria of the hybrid structure.

Introduction

Energy harvesting, also known as power harvesting, is the process of scavenging energy from ambient sources and storing that in a consumable form, typically, in the form of electrical energy. The recent development of autonomous and self-powered electronic systems is momentously facilitated by the technology of energy harvesting as it eliminates the drawbacks of the battery-dependent electronics (Ahmed, Mir, and Banerjee 2017). As a matter of fact, significant researches on energy harvesting have been going on over the last few decades (Ahmed, Mir, and Banerjee 2017; Roundy 2003; Rantz and Roundy 2017; Lee et al. 2016; Priya et al. 2017; Davidson and Mo 2014). This increase in research is invigorated by the modern advancements in low-power electronics such as the microelectromechanical systems (MEMS) and wireless technology such as wireless sensor networks (WSNs) (Priya et al. 2017). Applications of these electronic devices can be vigorously found in the fields of sensing and actuation, surveillance, biomedical science, animal migration tracking, structural health monitoring (Davidson and Mo 2014), aerospace and automobile engineering, and so on. In most cases, the low-powered wireless sensors and portable electronic devices are powered by conventional batteries which create complications such as limited lifespan, bulky weight, and maintenance issues including periodical recharging. Moreover, an environmental hazard can also be caused by the disposal of chemical batteries. Therefore, it becomes incumbent to find an effective alternative to the battery or similar nonregenerative sources for achieving the full benefits of the self-powered MEMS and wireless sensors nodes. An energy harvester (EH) is intuitively that alternative as it can extract energy from the ambient environment and can ensure an extended lifespan of the power supply by not only replacing the batteries but also recharging them.

In recent years, studies on energy harvesting from the ambient sources to power the low-powered electronic devices has gained excessive attention. The typical power requirement for the duty-cycled commercial WSNs ranges from 100 µW to 1 mW (Waterbury and Wright 2013) which can be extracted from the renewable energy sources such as, solar, (Sharma, Haque, and Jaffery 2018; Miles, Hynes, and Forbes 2005; Sivula 2015; Lewis, Zhang, and Jiang 2009; Grover, Nehra, and Kedia 2019), thermal (Al-Haik, Alothman, and Hajj 2018; Chavez et al. 2017), wind, hydroelectric, radio frequency (Sample and Smith 2009), and kinetic energy (including mechanical vibration and human locomotion) (Castagnetti 2011; Lee and Tovar 2013; Yuan et al. 2018). Among all the energy sources, solar energy and mechanical vibration are two of the most prolific ones that can be utilized while designing an efficient EH (Rahman 2016). Based on these two energy sources, significant number of investigations on the stand-alone EHs were reported by the researchers employing the photovoltaic (PV) (Sharma, Haque, and Jaffery 2018; Miles, Hynes, and Forbes 2005; Sivula 2015; Lewis, Zhang, and Jiang 2009; Grover, Nehra, and Kedia 2019), piezoelectric (PE) (Erturk and Inman 2008, 2009; De Pasquale, Somà, and Fraccarollo 2012; Zhang 2014; Hwang et al. 2015; Wang, Kobayashi, and Lee 2015; Cui et al. 2015; Jiang et al. 2015; Luo et al. 2016; Kang et al. 2016; Wurpts, Twiefel, and Brouet 2017), electromagnetic (EM) (Waterbury and Wright 2013; Glynne-Jones et al. 2004; Arnold 2007; Beeby, Tudor, and White 2006; Moss et al. 2015; Yildirim et al. 2017; Zhu et al. 2019), electrostatic (ES) (Roundy, Wright, and Rabaey 2003; Mitcheson et al. 2004; Tao et al. 2015; Khan and Qadir 2016), and magnetostrictive (Zhang 2011; Liu et al. 2015) transduction mechanisms. Due to simplicity in design and ability to obtain higher power densities, the PE technique is extensively used in standalone vibration energy harvesting. In addition to the PE technique, vibration EHs based on stand-alone EM or ES mechanism were also heavily investigated by many researchers at micro and meso scales. However, studies showed that a small-scale (e. g. micro scale) EH applying only one mechanism hardly provide necessary energy even for a microelectronic device. Therefore, a smart and expedient way to increase the power density and hereby improve the efficiency of the conventional EHs is the development of a hybrid energy harvester (HEH) that incorporates multi-transduction mechanisms at a time (Rahman 2016; Challa, Prasad, and Fisher 2009; Khaligh, Zeng, and Zheng 2010; Yang et al. 2010; Shan et al. 2013; Larkin and Tadesse 2013; Chen, Cao, and Xie 2015; Xu et al. 2016; Khan and Izhar 2016; Edwards, Aw, and Hu 2015; Su et al. 2016; Xu et al. 2017; Khbeis, McGee, and Ghodssi 2009; Eun et al. 2014; Madinei et al. 2016; Gambier et al. 2012; Goudarzi, Niazi, and Besharati 2013; Yu et al. 2014; Colomer-Farrarons et al. 2011; Rahman and Chakravarty 2018).

The concept of the hybrid energy harvesting can be illustrated into two points of view – harvesting from single source (e. g. vibration) and harvesting from several sources (e. g. vibration, solar, and thermal energy) (Rahman 2016). However, one of the most common types of HEH is the one that harvests energy from the same source, e. g. vibration-based hybrid piezo-electromagnetic harvesters (Challa, Prasad, and Fisher 2009; Khaligh, Zeng, and Zheng 2010; Yang et al. 2010; Shan et al. 2013; Larkin and Tadesse 2013; Chen, Cao, and Xie 2015; Xu et al. 2016; Khan and Izhar 2016; Edwards, Aw, and Hu 2015; Su et al. 2016; Xu et al. 2017). Challa, Prasad, and Fisher (2009), Khaligh, Zeng, and Zheng (2010), Yang et al. (2010), Shan et al. (2013), and Larkin and Tadesse (2013) investigated such types of HEHs that combine piezoelectric and electromagnetic mechanisms. More advancements on the hybrid piezo-electromagnetic EHs can be attributed to the investigations of Chen, Cao, and Xie (2015), Xu et al. (2016), Khan and Izhar (2016), Edwards, Aw, and Hu (2015), Su et al. (2016), and Xu et al. (2017) where attempts were taken to increase the power output. Design and characterization of the combined piezoelectric and electrostatic EHs (Khbeis, McGee, and Ghodssi 2009; Eun et al. 2014; Madinei et al. 2016) were studied by several researchers including Khbeis, McGee, and Ghodssi (2009), Eun et al. (2014), and Madinei et al. (2016) where simultaneous transductions of piezoelectric and electrostatic technique showed promising outcomes. Gambier et al. (2012) developed an effective multifunctional energy harvesting system which was able to successfully harvest the ambient energy combining the piezoelectric, photovoltaic, and thermoelectric methods. Goudarzi, Niazi, and Besharati (2013) offered a hybrid energy harvesting technique by simultaneous exploitation of the pyroelectric and piezoelectric effects. To take advantage from both light and vibration energy sources, a hybrid indoor ambient light and vibration energy harvester was presented by Yu et al. (2014) where low light illumination was captured by a small-scale amorphous-silicon solar panel. Colomer-Farrarons et al. (2011) presented a hybrid system combining photovoltaic, piezoelectric, thermoelectric, and electromagnetic mechanisms which could deliver more power than a stand-alone EH. Again, a hybrid energy harvesting technique based on solar radiation and mechanical vibration was proposed by Rahman and Chakravarty (2018), where the self-energized hybrid device combined the photovoltaic, piezoelectric, electromagnetic, and electrostatic mechanisms for simultaneous energy harvesting. All these studies showed that, the HEHs offer greater power output than the stand-alone harvesters because of their high energy-density, multi-functionality, and a wide range of frequency bandwidth. However, further studies are required on the efficient hybridization techniques to obtain better efficiency than the existing hybrid systems.

In this study, a HEH is designed and analyzed that can scavenge energy from mechanical vibration and solar radiation without any power supply from nonregenerative sources such as, batteries or fuel cells. This device offers increased power output and good efficiency as it combines the output obtained from the PV, PE, EM, and ES mechanisms simultaneously. A combined EH circuit operating at 5 V is designed in Proteus 8.6. The mathematical model of the hybrid energy harvester is coded and solved in MATLAB R2015a and the modal and structural analyses are conducted in ANSYS Workbench 15.0 using finite element (FE) method. Results for power outputs from the PE, EM, and ES mechanisms are obtained at the fundamental natural frequency of the hybrid beam using 1 g base excitation. As the hybrid structure is subjected to transverse vibration, it considers the necessity of the fatigue analysis to verify the structural reliability of the structure. Therefore, the fatigue behavior of the hybrid beam is also determined by conducting the FE analysis and studying the standard Stress-Life Curves (S-N Curves) for the base material of the structure, i. e. the aluminum substrate of the HEH model.

Design of the Hybrid Energy Harvester

Hybrid Cantilever Beam

The HEH consists of a bimorph Lead Zirconate Titanate (PZT) cantilever beam having a 6061-T6 Aluminum (Al) substrate and a cylindrical NdFeB magnet as a tip mass. The cylindrical magnet is surrounded by an induction coil of 1200 turns which is coaxially configured with the magnet on a frame. The volume of the Al substrate is 100 mm × 10 mm × 0.75 mm while each of the PZT layers have a volume of 90 mm × 8 mm × 0.40 mm. The PZT layers operate in d31 mode (i. e. with transverse piezoelectric strain coefficient) as the beam vibrates in transverse direction. Two sets of copper electrodes, in the shape of a hair comb, are attached on two sides of the Al substructure by keeping insulations in between. When the beam undergoes harmonic excitation, they act as the variable capacitance electrodes (Cvar electrodes). The PZT layers and the two sets of Cvar electrodes are assumed to be perfectly bonded to the Al substrate. A simplified two-dimensional view of the configuration is shown in Figure 1 while the actual three-dimensional configuration is demonstrated in Figures 2 and 3.

Figure 1: 
						Schematic diagram of the hybrid cantilever beam showing top and front views (figure not drawn to scale).
Figure 1:

Schematic diagram of the hybrid cantilever beam showing top and front views (figure not drawn to scale).

Figure 2: 
						(a) Configurations of the 3-D hybrid energy harvester without the fixed electrodes, (b) exploded 3-D view of the hybrid structure showing the sets of fixed electrodes on two sides of the beam.
Figure 2:

(a) Configurations of the 3-D hybrid energy harvester without the fixed electrodes, (b) exploded 3-D view of the hybrid structure showing the sets of fixed electrodes on two sides of the beam.

Figure 3: 
						Top view of the hybrid energy harvester showing the details of fixed and Cvar electrodes.
Figure 3:

Top view of the hybrid energy harvester showing the details of fixed and Cvar electrodes.

Each set of comb-shaped Cvar electrodes contains 70 electrodes, each having the dimension of 9 mm × 0.75 mm × 0.3 mm. The gap between two successive Cvar electrodes in the comb is 1.1 mm (Rahman and Chakravarty 2018). Two sets of similar combs of electrodes (fixed or, stationary capacitors) are attached with the base on two sides of the beam where each electrode has the dimension of 10 mm × 1 mm × 0.2 mm. This is illustrated in Figure 2(b) where the fixed electrodes are complementing the Cvar electrodes by fitting into the gaps. When the fixed electrodes and Cvar electrodes overlap, the gap between the electrodes is 0.45 mm as shown in Figure 3. Since the beam oscillates in transverse direction, this gap is assumed to be constant throughout the electrostatic power generation. Among the three fundamental electrostatic configurations – in-plane overlap varying, in-plane gap closing, and out-of-plane gap closing, the third one is selected for this design to avoid contacts between the moving and stationary electrodes. The relative motion between the precharged moving capacitors and stationary capacitors governs the electrostatic transduction.

The geometric and physical properties of the HEH considered for the analysis are shown in Table 1. The hybrid structure of the device is scalable and tunable. The tip mass and the length of the comb electrodes can be adjusted to operate at the resonance frequency (Rahman and Chakravarty 2018; Challa, Prasad, and Fisher 2011; Chen, Yang, and Deng 2009). Since the performance of the EH is sensitive to the smallest amount of change of its geometry (Yi, Shih, and Shih 2002), attention must be paid to the effective scaling of the design.

Table 1:

Geometric and physical parameters of the hybrid cantilever beam (Rahman 2016; Rahman and Chakravarty 2018).

Parameter Value
Length of the cantilever beam, L 100 mm
Length of the PZT layer, Lp 90 mm
Length of the comb-electrodes, Le 100 mm
Height and Width of the substrate, hs and bs 0.75 mm and 10 mm
Height and width of each PZT layer, hp and bp 0.4 mm and 8 mm
Piezoelectric strain constant, e31 13.74 C/m2
Dielectric permittivity of the PZT layer, ε33s 6.46 nF/m
Height and effective width of the comb-electrodes, he and be 0.75 mm and 2.9 mm
Area moment of inertia of the beam, Ic 0.001066 m4
Height and radius of the magnet 20 mm and 4.2 mm
Number of turns and height of the coil 2000 and 25 mm
Inner and outer radii of the coil 5.2 mm and 7.2 mm
Length, l and radius of the coil wire 12 m and 0.0615 mm
Distance between the coil and the magnet, d 1 mm
Magnetic flux density, B 1.18 T
Density of comb electrodes, ρe 8960 kg/m3
Density of PZT, ρp 7800 kg/m3
Density of the substrate, ρs 2700 kg/m3
Density of the magnet, ρm 7400 kg/m3
Young’s modulus of comb electrodes, Ee 117 Gpa
Young’s modulus of PZT, Ep 66 GPa
Young’s modulus of substrate, Es 70 GPa
Young’s modulus of the magnet, Em 160 GPa

Photovoltaic Panel

The PV panel includes an array of 10 high-efficiency monocrystalline solar cells as shown in Figure 4. The size of the PV unit is 55 mm × 55 mm × 3 mm weighing 36 grams. The item number of PV panel is KS-M5555 which is selected from the China Solar LTD (http://www.solars-china.com/solar-panel/5v-80mA-oem-solar-panel.html). The data sheet information is available in the manufacturer’s website (http://www.solars-china.com/solar-panel/5v-80mA-oem-solar-panel.html). The output from the PV panel is DC while the PE, EM, and ES outputs are AC outputs. Therefore, the AC outputs need to be rectified to DC outputs in the combined circuit. The HEH is designed in such a way that it allows freedom of customization of the PV panel based on the power requirement (e. g. using different scale or type) because the PV panel is not an integrated part of the hybrid cantilever beam. However, the power (bias input) supplied to the Cvar electrodes form the PV panel must be adequate to excite the electrodes in order to generate electrostatic power output.

Figure 4: 
						Configuration of the 5 V 80 mA solar panel (Item no: KS-M5555) (http://www.solars-china.com/solar-panel/5v-80mA-oem-solar-panel.html).
Figure 4:

Configuration of the 5 V 80 mA solar panel (Item no: KS-M5555) (http://www.solars-china.com/solar-panel/5v-80mA-oem-solar-panel.html).

The operating voltage and current of the PV panel are 5 V and 80 mA, respectively. The other important specifications are shown in Table 2.

Table 2:
Parameter Value
Open circuit voltage, VOC 6 V
Short circuit current, ISC 87 mA
Maximum power, PPV 0.4 W
Nominal operating cell temperature 45 ± 2 °C
Temperature coefficient of ISC (0.045 ± 0.01) %/°C
Temperature coefficient of VOC (0.34 ± 0.01) %/°C
Temperature coefficient of power (0.47 ± 0.05) %/°C
Efficiency, ηSolar 17 %
Operating temperature 40 °C to 85 °C
Table 3:

Parameters considered for analytical calculation and simulation.

Parameter Value Parameter Value Parameter Value
ε 0 8.854 × 10–12 F/m L 0.1 m keff 581.26 N/m
Fe 5.534 × 10–5 N Lp 0.09 m δc 47.518 Tm
B 1.1 T Li 0.0171 H ωc 1.05 × 107
μ 0 1.257 × 10–6 H/m R C o i l 67.2 Ω α 2.6 × 10−4
RL 180 kΩ R E l e c t r o d e 12.5 Ω Y0 1.8 × 10–4 m
Cp 1.16 × 10–8 F N 2000 de 1.274 × 10–3 kg/s
Fn 4.56 N rc 0.0052 m dm 0.099 kg/s
Mt 0.0082 kg rm 0.0042 m bs 0.01 m
mL 0.0104 kg hc 0.025 m bp 0.008 m
ρe 8960 kg/m3 hm 0.020 m be 0.0029 m
ρm 7400 kg/m3 hs 0.00075 m χ n 7.9 × 10−5
ρp 7800 kg/m3 hp 0.00040 ζ n 0.02
ρs 2700 kg/m3 d 0.001 m φ n 7.9 × 10−5

A Hybrid Power Conditioning Circuit

A simplified circuit diagram of the proposed HEH is shown in Figure 5 where the PV, ES, PE, and EM generators are parallelly connected to operate at 5 V. The PV panel generates DC while the vibration-based methods induce AC which must be converted to DC by rectifiers when combined with the PV generator. Voltage regulators maintain a constant operating voltage of 5 V. The combined current passes through a microcontroller that decides where to pass the current through a load resistance or to store it in a low voltage (e. g. 3.6 V) battery. The microcontroller also plays an important role in biasing the ES system by supplying a portion of the harvested energy or charge from the battery to the Cvar electrodes. If the harvested power is low, the battery will provide the back-up power to the load resistance through the microcontroller. The design of the power harvesting circuit thus allows a continuous power supply to the MEMS and WSNs even though the solar energy or mechanical vibration is inadequate or absent.

Figure 5: 
					Schematic diagram of the combined circuit of the hybrid energy harvester.
Figure 5:

Schematic diagram of the combined circuit of the hybrid energy harvester.

Mathematical Model

The mathematical model of the proposed HEH is developed including the photovoltaic equations and the governing partial differential equation of a vibrating cantilever beam with a tip mass in a distributed parameter system (Rahman 2016; Rahman and Chakravarty 2018). Under a small-amplitude harmonic excitation, the Euler–Bernoulli beam theory is applicable to the hybrid beam. Therefore, the shear deformation and rotary inertia of the beam are neglected. Expressions of power output are presented for each mechanism and finally, the total power output and efficiency formulations are derived.

Photovoltaic Equations

Considering a PV array with NS number of cells in series and NP number of strings in parallel. the output current I and output voltage V can be obtained from eqs. (1)(4) (Rahman and Chakravarty 2018; Sood and Bhalla 2013).

(1) I = N P I P H N P I S [ exp ( q ( V / N S + I R S ) N p / k T C A ) 1 ] ( N p V / N s + I R s ) R S H

where IPH is the light-generated current or photocurrent, IS is the cell saturation of dark current, q = 1.6 × 10–19 C is the electron charge, k = 1.38 × 1023 J/K is Boltzmann’s constant, Tc is the cell’s working temperature, A is diode’s ideality factor (typically between 1 and 2), RSH is the shunt resistance, and RS is the series resistance. The photocurrent IPH can be calculated as follows:

(2) I P H = I S C + K 1 T C T r e f λ

where ISC is the short-circuit current of a cell at reference temperature of 25 °C and reference irradiance of 1000 W/m2, K1 = 0.032 is the cell’s short-circuit current temperature coefficient (Sood and Bhalla 2013), Tref is the cell’s reference temperature, and λ is the solar insolation (irradiance) in W/m2. The cell’s saturation current varies with the change in cell temperature, which is expressed as

(3) I S = I R S T C T r e f 3 exp q E G 1 T r e f 1 T C k A

where IRS is the cell’s reverse saturation current at a reference temperature and irradiance, and EG is the band energy of the semiconductor used in the cell. The reverse saturation current IRS at the reference temperature can approximately be obtained as

(4) I R S = I S C e x p q V O C N s k T C A 1

Several parameters including the open-circuit voltage, VOC (at I = 0), short-circuit current, ISC (at V = 0), and the fill factor (FF) determine the performance of a solar cell, which are related as

(5) F F = I m p V m p I S C V o c = P p v I S C V o c

where Imp and Vmp are the current and voltage operating points for the maximum power Ppv, respectively. The output power is calculated by taking the product of either I and V, or, I2 and RL (Rahman and Chakravarty 2018). However, in terms of the FF, the maximum PV power output is determined by

(6) P p v = F F × I S C V O C

Piezoelectric Equations

Assuming the Euler-Bernoulli beam, the coupled electromechanical equation for a bimorph PZT cantilever beam of length L with a tip mass Mt and mass per unit length mL can be represented as follows (Rahman 2016; Erturk and Inman 2008; Rahman and Chakravarty 2018):

(7) E c I c 4 w r e l x , t x 4 + C s I c 5 w r e l x , t x 4 t + C a w r e l x , t t + m L 2 w r e l x , t t 2 + ϑ V t d δ x d x d δ x L d x = m L + M t δ x L 2 w b x , t t 2

Here, CsIc is the internal (Kelvin-Voigt) damping and Ca is the viscous air damping. The sum of the harnonic base displacement wb=Y0ejωt and relative displacement wrel gives the total transverse displacement w at any time t and location x of the beam as follows:

(8) w x , t = w b x , t + w r e l x , t

The electromechanical coupling term ϑ for series connection of bimorph PZT is given by

(9) ϑ = e 31 b 2 h p h s 2 4 h p + h s 2 2

where e31 is the piezoelectric strain constant. The vibration response relative to the base can be represented as a convergent series of the Eigen functions by following modal damping (Erturk and Inman 2008, 2009; Rahman and Chakravarty 2018) as follows:

(10) w r e l x , t = n = 1 ϕ n x η n t

where n and ηn are the mass normalized Eigen function and the modal coordinate of the clamped-free beam for the nth mode, respectively. Now, the governing equation indicates that, four boundary conditions are required in terms of x and two initial conditions are needed in terms of t for the solution (Rahman and Chakravarty 2018). At the fixed end (i. e. x = 0),

(11) w r e l = 0 , w r e l x = 0

At the free end (i. e. x = L), moment M is zero, but shear force is equal to the applied force Fn at the tip, i. e.

(12) M = E c I c 2 w r e l x 2 = 0
(13) F n = x E c I c 2 w r e l x 2 = m e f f g

where meff represents the effective mass of the hybrid beam and g is the gravitational acceleration. The modal mechanical response ηn can be obtained from the mass normalized reduced partial differential equation as follows:

(14) η n . . t + 2 ζ n ω n η n . t + ω n 2 η n t + χ n V t = f n t

Here, fn is the mass normalized tip force on the beam. The backward modal electromechanical coupling term χn can be given by

(15) χ n = ϑ d ϕ n x d x

Applying the Kirchhoff laws to the circuit depicted in Figure 6, the capacitance CP and piezoelectric current ipt for the series connection of the PZT layers (Rahman and Chakravarty 2018) can be obtained by

(16) C P 2 d V t d t + V t R L = i p t
Figure 6: 
						(a) Series connection of the two PZT layers and, (b) electrical circuit corresponding to the series connection (Erturk and Inman 2008, 2009).
Figure 6:

(a) Series connection of the two PZT layers and, (b) electrical circuit corresponding to the series connection (Erturk and Inman 2008, 2009).

where the capacitance and the piezoelectric current for each layer of PZT can be defined as,

(17) C P = ε 33 s b p L p h p
(18) i p t = n = 1 φ n d η n t d t

In eq. (17), ε33s is the dielectric permittivity of the PZT layer for the series connection.

The forward modal coupling term φn can be expressed as,

(19) φ n = e 31 b h p + h s 2 d n x d x

The cross-section of the hybrid beam through the Cvar electrodes in y-z plane is shown in Figure 7.

Figure 7: 
						Cross-section of the beam through the Cvar electrodes.
Figure 7:

Cross-section of the beam through the Cvar electrodes.

The equivalent bending stiffness EcIc and the mass of the beam m without the tip mass can be given by,

(20) E c I c = E s 24 b s h s 3 + E p 24 b p 2 h p + h s 3 b p h s 3 + 2 E e 24 b e h s 3 × 2
(21) m = ρ A c r o s s L = ρ s b s h s L s + 2 ρ p b p h p L p + 2 ρ e b e h s L e

Therefore, the fundamental natural frequency ωn of the hybrid beam can be found by,

ω n = k e f f m e f f
= 3 E c I c L 3 0.236   m + M t
(22) = 6 E s 24 b s h s 3 + E p 24 b p 2 h p + h s 3 b p h s 3 + 2 E e 24 b e h s 3 0.236 ρ s b s h s L + 2 ρ p b p h p L p + 2 ρ e b e h s L + M t L 3

where the effective stiffness and effective mass of the beam are keff=3EI/L3 and meff=0.236m+Mt, respectively (Rahman and Chakravarty 2018). The maximum output power is obtained at the resonance condition, i. e. at ω=ωn. The expression of power output PPEω for the PE generator at resonance can be given by eq. (23) (Rahman and Chakravarty 2018; Cottone 2011).

(23) | P P E ( ω ) | = V 2 ( t ) R L = R L ( ω φ n F n ) 2 [ ω n 2 ω 2 ( 1 + 2 ζ n ω n R L C P ) ] 2 + [ 2 ζ n ω n ω + ω R L [ C P ( ω n 2 ω 2 ) + φ n χ n ] ] 2

where RL is the load resistance and ζn is the damping ratio of the bimorph PZT cantilever beam. At resonance, Fn is equal to resonance-equivalent static load (Challa, Prasad, and Fisher 2009) at the tip, i. e. Fn = Ftip= m+Mtg/2ζn.

Electromagnetic Equations

The expression for the maximum EM power output PEMω at the resonance frequency (Rahman and Chakravarty 2018; Cottone 2011; Kim et al. 2010) is given by the following equation:

(24) | P E M ( ω ) | = Y 0 2 2 R L [ m e f f δ c ω c ω n 2 ω ( k e f f ω c m e f f ω c ω 2 d ω 2 ) 2 + ( k e f f ω m e f f ω 3 + d ω c ω + α δ c ω c ω ) 2 ] 2

where the electrical coupling force factor α, characteristic cut-off frequency ωc, electromechanical conversion factor δc, and coil self-inductance Li are expressed as follows (Rahman and Chakravarty 2018; Cottone 2011; Kim et al. 2010; Miller et al. 2016):

(25) α = B l R L
(26) ω c = R L L i
(27) δ c = B l
(28) L i = μ 0 N 2 π r c 2 h c

where B is the magnetic flux density, l is the length of wire in the coil, N is the number of turns in the coil, μ0 is the permeability of the coil material, rc is the coil radius, and hc is the coil height.

Electrostatic Equations

The maximum power output expression for the ES mechanism PESω at the resonance frequency (Cottone 2011) is given by,

(29) P E S ω = Y 0 2 2 d e m e f f 2 ω n 4 ω 2 k e f f m e f f ω 2 2 + ( d e + d m ) 2 ω 2

In eq. (29), the mechanical damping dm and the electrical damping de can be expressed as,

(30) d m = 2 ζ n m e f f ω n
(31) d e = F e Y o ω n

where the electric force Fe can be given by,

(32) F e = ε 0 A c V 2 2 d c 2

where ε0 is the permittivity at free space, Ac is the overlap area of the comb electrodes, and dc is the gap between two electrodes.

Combined Power Output and Efficiency

The combined maximum power output Pcombined is not simply the sum of all the stand-alone outputs because the damping of the system causes power loss Ploss. Therefore, the total power output combining all the generators is determined by

(33) P c o m b i n e d = P o u t P l o s s

where the total power output Pout is given by

(34) P o u t = P P V + P P E ω + P E M ω + P E S ω

The power loss due to the mechanical and electrical damping (Xu et al. 2017) of the hybrid system can be calculated by

(35) P l o s s = P D + P C o i l + P E l e c t r o d e s

In eq. (35), the average power losses PD, PCoil, and PElectrodes due to the mechanical damping dm, internal resistance of the coil RCoil, and internal resistance of variable capacitance electrodes RElectrodes, respectively, are expressed as

(36) P D = 1 2 d m ω n 2 Y 0 2
(37) P C o i l = P E M ω R C o i l R L
(38) P E l e c t r o d e s = P E S ω R E l e c t r o d e s R L

The bias input Pin,ES for the ES mechanism, which is supplied by either the battery or the PV panel to the Cvar electrodes, should be considered as an input while calculating the efficiency of the HEH. If the input power Pin is a sum of the Pin,ES, solar power input, and vibratory power input, then the overall efficiency η of the HEH (Miller et al. 2016) becomes,

(39) η = P C o m b i n e d P i n = P C o m b i n e d P s o l a r + P v i b + P i n , E S × 100 %

The magnitude of the vibrational power input Pvib is calculated as follows (Rahman 2016; Rahman and Chakravarty 2018), i. e.

(40) P v i b = 1 2 m e f f g Y o ω n

The solar power input Psolar is given by,

(41) P s o l a r = λ A P V = P P V η S o l a r

where λ is the irradiance and APV is the effective photovoltaic area.

Model Validation with Finite Element Analysis

Fundamental Natural Frequency

The fundamental natural frequency ωn of the proposed HEH is found 37.16 Hz by eq. (22). Since the power output of the HEH is sensitive to the natural frequency ωn, finite element (FE) modeling is conducted in ANSYS Workbench 15.0 to verify the analytical result for ωn. The FE analysis result for ωn of the HEH is found 35.867 Hz with a deviation of 3.48 % as compared to the analytical result. Figure 8 shows the first resonant mode shape of the HEH obtained by the FE analysis which indicates a good agreement with the first normalized mode shape of a typical cantilever beam with a tip mass (Rahman 2016; Rahman and Chakravarty 2018).

Figure 8: 
						First mode shape of the hybrid cantilever beam with tip mass at ωn{\omega _n} = 35.866 Hz.
Figure 8:

First mode shape of the hybrid cantilever beam with tip mass at ωn = 35.866 Hz.

Based on the fundamental natural frequency, a mesh independence or, convergence study is conducted using ANSYS where the value of ωn converges to 35.867 Hz with the increase of degrees of freedom. Figure 9(a) depicts the default converged 3-D mesh of the hybrid beam where 38,402 elements are connected with 121,087 nodes. The convergence of the value of ωn with the number of nodes is shown in Figure 9(b).

Figure 9: 
						(a) Computational mesh of the hybrid beam and, (b) Convergence of ωn{\omega _n} value with the increase of DOF.
Figure 9:

(a) Computational mesh of the hybrid beam and, (b) Convergence of ωn value with the increase of DOF.

Stress and Fatigue Analysis

Stress and fatigue analyses of the HEH are also done in ANSYS Workbench 15.0 to study the stress and failure criteria of the hybrid structure. The maximum bending stress σ of the beam is found 2.939 MPa analytically following the equation σ=Mc/Ic where c = hp + hs/2 (Rahman 2016). The FE analysis result for the bending stress is found 2.868 MPa which substantiated the accuracy of the analytical calculation by yielding 2.12 % deviation. A fatigue analysis is also done in ANSYS for the substrate material, i. e. Aluminum, since it is clamped to the base, carries the PZT layers (which are not clamped to the base, starts from 0.1 mm on the substrate), Cu electrodes, and tip magnet, and undergoes the maximum stress at the root. Using a fully reversed cyclic load of F=meff.g = 0.10455 N at the tip, a fatigue life of nearly 107 cycles is found following the Goodman stress-life approach. Figure 10 shows the distribution of bending stress and its maximum value at the fixed end of the hybrid cantilever beam which is obtained from the FE analysis in ANSYS Workbench.

Figure 10: 
						Distribution of stress on the hybrid beam due to tip loading.
Figure 10:

Distribution of stress on the hybrid beam due to tip loading.

Results and Discussion

Power and Optimum Load Resistance at the Resonance

The results for the output power of the HEH are obtained in MATLAB R2015a following the mathematical model. The parameters considered for the analytical calculation and numerical simulation are listed in Table 3.

The PV power output along with the I-V characteristics is obtained by developing a MATLAB algorithm based on the maximum power point tracking (MPPT) technique at an irradiance level of 1000 W/m2 (AM 1.5) at 25 °C temperature (Rahman, Sarker, and Chakravarty 2019). Figure 11 shows the maximum power output from the PV panel along with the I-V characteristics as obtained from the MPPT algorithm.

Figure 11: 
						Power output from the PV panel.
Figure 11:

Power output from the PV panel.

The calculations for PE, EM, and ES power outputs are done considering 1 g harmonic base excitation which shown in Figure 12.

Figure 12: 
						Harmonic base excitation with 1 g (9.81 m/s2) amplitude.
Figure 12:

Harmonic base excitation with 1 g (9.81 m/s2) amplitude.

The peak of the output power for the PE, EM, and ES generators occurred at resonance, i. e. the fundamental natural frequency ωn which is found 37.16 Hz. The power outputs around the resonance frequency bandwidth for the PE, EM, and ES methods are shown in Figure 13.

Figure 13: 
						Power outputs at the optimum load resistance from the three vibration-based standalone mechanisms: (a) PE, (b) EM, and (c) ES.
Figure 13:

Power outputs at the optimum load resistance from the three vibration-based standalone mechanisms: (a) PE, (b) EM, and (c) ES.

The optimal load resistance is considered as 180 kΩ for the maximum PE output at 37.16 Hz. This is shown in Figure 14 where the resistance at the peak power is the optimal load resistance.

Figure 14: 
						Optimal load resistance RL at resonance.
Figure 14:

Optimal load resistance RL at resonance.

In general, energy harvesters are designed to operate in the fundamental natural frequency, i. e. the first mode of vibration which typically provides the maximum deflection and therefore gives the maximum electrical energy (Rahman and Chakravarty 2018). Erturk and Inman (2008, 2009) mentioned that the sign of the mode shapes of a cantilever beam during vibration is analogous to the sign of axial strain distribution along the length and any change in sign reduces the output voltage (Rahman and Chakravarty 2018). As the first mode shape does not change its sign (i. e. stays at one side of the x axis), the corresponding axial strain curve also remains unchanged in its sign. This ensures the maximum voltage generation by the PZT layers. Therefore, strong emphasis is given on determining the fundamental natural frequency and the first mode shape of the structure before calculating the power output and efficiency.

Calculation of Total Power Output and Efficiency

The efficiency of the HEH is determined by eq. (39). The magnitude of Pvib is found 2196 µW for 1 g base excitation. On the other hand, considering 17 % efficiency and an irradiance level of 1000 W/m2 at AM 1.5, the solar energy input Psolar yielded a value of 2.35 W. 10 % of the PV power output is utilized for biasing the ES mechanism. The alternate way to provide the bias input for the ES generator is to use the low voltage battery of the circuit. In that case, the rated amount of PV output, i. e. 0.4 W is harvested which contributes to its maximum efficiency. The total power loss due to damping of the system is calculated by using eqs. (36)(38). The power inputs and the total amount of power loss are shown in Figure 15.

Figure 15: 
						Amounts of input power and power loss for the hybrid energy harvester.
Figure 15:

Amounts of input power and power loss for the hybrid energy harvester.

The total power output of the HEH is determined by adding all the stand-alone power outputs and deducting the power losses from that amount. While calculating the efficiency, the total input power is considered as the summation of the solar input, base excitation input, and the initial bias input to the ES generator. First, it is assumed that the bias input for the ES generator is supplied by the battery through the microcontroller. The total power output of the HEH in that case is found 772.972 mW with an overall efficiency of 32.3 %, which is greater than the stand-alone power outputs of the PV, PE, EM, and ES generators. The results are shown in Figure 16.

Figure 16: 
						Results for power outputs and efficiency (%) with biasing by the battery.
Figure 16:

Results for power outputs and efficiency (%) with biasing by the battery.

Instead of the battery, if the PV panel provides 40,000 µW solar power for biasing the ES generator, the total power output of the HEH drops slightly due to the lack of PV power output. In that case, efficiency of the PV panel decreases to 15.3 % and the total power output becomes 732.973 mW with an overall efficiency of 30.6 %. This is illustrated in Figure 17.

Figure 17: 
						Results for power outputs and efficiency (%) with biasing by the solar panel.
Figure 17:

Results for power outputs and efficiency (%) with biasing by the solar panel.

Results displayed in Figures 16 and 17 indicate that the PV and EM generators are the major contributors in the total power output. Results also indicate that, if the battery or the PV panel fails to supply the bias input, the EM generator will still be able to provide the bias input to the ES generator in order to maintain a continuous power transduction.

Effect of Change in Scale of the Beam

Since the fundamental natural frequency ωn plays an important role in the performance of the vibration energy harvesters, the effects of change in scale or size of the harvester are studied considering the value of ωn. A parametric study is conducted to observe the effects of change in length and width on the natural frequency of the beam and the total power output from the PE, EM, and ES generators. As the expression for the ωn suggests, the increases in length L and mass meff result in the decrease in the value of ωn for a constant width and thickness. Figures 18 shows the effect of changing length on the fundamental natural frequency of the beam.

Figure 18: 
						Effect of change in length on the fundamental natural frequency of the beam.
Figure 18:

Effect of change in length on the fundamental natural frequency of the beam.

On the other hand, at a given length and thickness, the fundamental natural frequency increases with the increase of width of the beam. This effect is illustrated in Figure 19 where the additional width increases the effective mass of the beam but contributes to the increase of the fundamental natural frequency value by increasing the area moment of inertia.

Figure 19: 
						Effect of change in width on the fundamental natural frequency of the beam.
Figure 19:

Effect of change in width on the fundamental natural frequency of the beam.

Knowing the variation of natural frequency with the structural parameters is also critical for tuning the vibration energy harvesters. For a given input excitation, any change in natural frequency as a result of adjusting the tip mass or changing the scale, can change the power output of the energy harvester (Rahman 2016). Therefore, it is necessary to optimize the length and width of the beam depending on the value of the fundamental natural frequency. The variations of natural frequency and power output with the change in length and effective width at a given thickness of the vibration-based generators (i. e. PE, EM, and ES generators) are shown in Table 4.

Table 4:

Power output from the vibration-based generators at different scales.

Case Length (m) Effective width (m) Thickness (m) Effective mass (Kg) Natural frequency (Hz) PE + EM + ES power output (mW)
1 0.1 0.006 0.00155 0.00994 29.170 125.37
0.1 0.01 0.00155 0.01066 37.167 373.15
0.1 0.015 0.00155 0.01156 44.176 378.52
2 0.6 0.01 0.00155 0.00963 84.14 51.40
0.8 0.01 0.00155 0.01014 53.246 225.35
0.12 0.01 0.00155 0.011173 27.615 378.56

As mentioned earlier, the size of the PV panel (also the number of arrays) can be adjusted based on the power requirement. The design of the HEH is very flexible in incorporating any small-size PV panel as it is externally connected to the vibration-based system.

Conclusions

In this paper, a self-energized hybrid solar and vibration energy harvester is designed and its performance for a standard base excitation and irradiance level is investigated. Physical and mathematical models of the hybrid structure are presented with illustrations. The proposed hybrid device generates more power than the stand-alone energy harvesters with decent efficiency. In absence of solar energy, the electrostatic components can still obtain the initial bias input from the energy storage or the harvested power with the help of a microcontroller which implies that failure of one generator will not result in disruption of the entire power generation. The proposed design thus, ensures a simultaneous harvesting of energy and a continuous supply of power to the low-powered electronic components. The reciprocal interactions (e. g. electromechanical damping) of the mechanisms are handled by considering a reasonable value of damping of the hybrid system. While calculating the actual power output, considering the effect of damping is important because the material damping, surrounding air damping, internal resistances of the coil, and electrostatic action of the electrodes cause a power loss during the hybrid power generation. The effect of changing scale of the geometry on the performance of the hybrid device is also studied.

A finite element analysis is also done to validate the model and investigate the fatigue behavior of the structure. The analytical results for the fundamental frequency and bending stress show a good agreement with the finite element analysis results. The results for the expected life and potential failure zone of the aluminum substrate are quite reasonable compared to the standard S-N curves. Overall, the coupled electromechanical model, combined circuit, and finite element analysis presented in this study facilitate the knowledge and development of an efficient hybrid energy harvester for meeting the growing demand of power management for wireless sensors and microelectromechanical systems.

Funding statement: The authors received no financial support for this research.

Conflict of Interest

There is no conflict of interest in this work. However, the authors are thankful to Bikrant Poudel from the Department of Electrical Engineering, University of New Orleans for his cooperation and electrical engineering concepts.

References

Ahmed, R., F. Mir, and S. Banerjee. 2017. “A Review on Energy Harvesting Approaches for Renewable Energies from Ambient Vibrations and Acoustic Waves Using Piezoelectricity.” Smart Materials and Structures 26 (085031): 1−33.10.1088/1361-665X/aa7bfbSuche in Google Scholar

Al-Haik, M., A. Alothman, and M. Hajj. 2018. “Integrated Thermoelectric Energy Generator and Organic Storage Device.” Energy Harvesting and Systems 5 (3–4): 73–79.10.1515/ehs-2018-0009Suche in Google Scholar

Arnold, D. P. 2007. “Review of Microscale Magnetic Power Generation.” IEEE Transactions on Magnetics 43: 3940–3951.10.1109/TMAG.2007.906150Suche in Google Scholar

Beeby, S. P., M. J. Tudor, and N. M. White. 2006. “Energy Harvesting Vibration Sources for Microsystems Applications.” Measurement Science and Technology 13: R175–95.10.1088/0957-0233/17/12/R01Suche in Google Scholar

Castagnetti, D. 2011. “Fractal-Inspired Multifrequency Structures for Piezoelectric Harvesting of Ambient Kinetic Energy.” ASME Journal of Mechanical Design 133 (11): 111005:1–8.10.1115/1.4004984Suche in Google Scholar

Challa, V. R., M. G. Prasad, and F. T. Fisher. 2009. “A Coupled Piezoelectric–electromagnetic Energy Harvesting Technique for Achieving Increased Power Output through Damping Matching.” Smart Materials and Structures 18 (005029): 1−11.10.1088/0964-1726/18/9/095029Suche in Google Scholar

Challa, V. R., M. G. Prasad, and F. T. Fisher. 2011. “Towards an Autonomous Self-tuning Vibration Energy Harvesting Device for Wireless Sensor Network Application.” Smart Materials and Structures 20 (025004): 1−11.10.1088/0964-1726/20/2/025004Suche in Google Scholar

Chavez, L., F. Zayas Jimenez, B. Wilburn, L. C. Delfin, H. Kim, N. Love, and Y. Lin. 2017. “Characterization of Thermal Energy Harvesting Using Pyroelectric Ceramics at Elevated Temperatures.” Energy Harvesting and Systems 5 (1–2): 3–10.10.1515/ehs-2018-0002Suche in Google Scholar

Chen, W., Y. Cao, and J. Xie. 2015. “Piezoelectric and Electromagnetic Hybrid Energy Harvester for Powering Wireless Sensor Nodes in Smart Grid.” Journal of Mechanical Science and Technology 29 (10): 4313−18.10.1007/s12206-015-0928-xSuche in Google Scholar

Chen, Z. S., Y. M. Yang, and G. Q. Deng. 2009. “Analytical and Experimental Study on Vibration Energy Harvesting Behaviors of Piezoelectric Cantilevers with Different Geometries.” International Conference on Sustainable Power Generation and Supply Nanjing, China: IEEE, pp.1–6.10.1109/SUPERGEN.2009.5348290Suche in Google Scholar

Colomer-Farrarons, J., P. Miribel-Catala, A. Saiz-Vela, and J. Samitier. 2011. “A Multiharvested Self-Powered System in A Low-Voltage Low-Power Technology.” IEEE Transactions on Industrial Electronics 58 (9): 4250−63.10.1109/TIE.2010.2095395Suche in Google Scholar

Cottone, F. 2011, “Introduction to Vibration Energy Harvesting.” Retrieved from http://wwwnipslab.org/files/file/nips%20summer%20school%202011/Cottone%20Introduction%20to%20vibration%20harvesting.pdf.Suche in Google Scholar

Cui, Y., Q. Y. Zhang, M. L. Yao, W. J. Dong, and S. Q. Gao. 2015. “Vibration Piezoelectric Energy Harvester with Multi-beam.” AIP Advances 5: 041332−9.10.1063/1.4919049Suche in Google Scholar

Davidson, J., and C. Mo. 2014. “Recent Advances in Energy Harvesting Technologies for Structural Health Monitoring Applications.” Smart Materials Research 2014 (410316): 1−14.10.1155/2014/410316Suche in Google Scholar

De Pasquale, G., A. Somà, and F. Fraccarollo. 2012. “Piezoelectric Energy Harvesting for Autonomous Sensors Network on Safety-improved Railway Vehicles.” Proceedings of the Institution of Mechanical engineers, Part C: Journal of Mechanical Engineering Science 226 (4): 1107–1117.10.1177/0954406211418158Suche in Google Scholar

Edwards, B., K. C. Aw, and A. P. Hu. 2015. “Hybrid Electromagnetic-piezoelectric Transduction for a Frequency Up-converted Energy Harvester.” IEEE International Conference on Advanced Intelligent Mechatronics, pp. 1149−54.10.1109/AIM.2015.7222694Suche in Google Scholar

Erturk, A., and D. J. Inman. 2008. “On Mechanical Modeling of Cantilevered Piezoelectric Vibration Energy Harvesters.” Journal of Intelligent Material Systems and Structures 19: 1311–1325.10.1177/1045389X07085639Suche in Google Scholar

Erturk, A., and D. J. Inman. 2009. “An Experimentally Validated Bimorph Cantilever Model for Piezoelectric Energy Harvesting from Base Excitations.” Smart Materials and Structures 18 (025009): 1−18.10.1088/0964-1726/18/2/025009Suche in Google Scholar

Eun, Y., D. Kwon, M. Kim, I. Yoo, J. Sim, H. Ko, K. Cho, and J. Kim. 2014. “A Flexible Hybrid Strain Energy Harvester Using Piezoelectric and Electrostatic Conversion.” Smart Materials and Structures 23 (045040): 1−6.10.1088/0964-1726/23/4/045040Suche in Google Scholar

Gambier, P., S. R. Anton, N. Kong, A. Erturk, and D. J. Inman. 2012. “Piezoelectric, Solar and Thermal Energy Harvesting for Hybrid Low-power Generator Systems with Thin-film Batteries.” Measurement Science and Technology 23 (01510): 1−11.10.1088/0957-0233/23/1/015101Suche in Google Scholar

Glynne-Jones, P., M. J. Tudor, S. P. Beeby, and N. M. White. 2004. “An electromagnetic, Vibration-powered Generator for Intelligent Sensor Systems.” Sensors and Actuators A 110: 344−49.10.1016/j.sna.2003.09.045Suche in Google Scholar

Goudarzi, M., K. Niazi, and M. K. Besharati. 2013. “Hybrid Energy Harvesting from Vibration and Temperature Gradient by PZT and PMN-0.25PT Ceramics.” Materials Physics and Mechanics 16: 55–65.Suche in Google Scholar

Grover, M., M. Nehra, and D. Kedia. 2019. “Simulative Parametric Study on Heterojunction Thin Film Solar Cells Incorporating Interfacial Nanoclusters Layer.” Energy Harvesting and Systems 6 (1–2): 23–28.10.1515/ehs-2019-0004Suche in Google Scholar

Hwang, G. T., M. Byun, C. K. Jeong, and K. J. Lee. 2015. “Flexible Piezoelectric Thin-film Energy Harvesters and Nanosensors for Biomedical Applications.” Advanced Healthcare Materials 4: 646–658.10.1002/adhm.201400642Suche in Google Scholar PubMed

Jiang, X., J. Wang, Y. Li, J. Li, and J. Yao. 2015. “Energy Harvesting for Powering Wireless Sensor Networks in Low-frequency and Large-force Environments.” Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science 229 (11): 1953–1964.10.1177/0954406214551038Suche in Google Scholar

Kang, M. G., W. S. Jung, C. Y. Kang, and S. J. Yoon. 2016. “Recent Progress on PZT Based Piezoelectric Energy Harvesting Technologies.” Actuators 5 (5): 1−17.10.3390/act5010005Suche in Google Scholar

Khaligh, A., P. Zeng, and C. Zheng. 2010. “Kinetic Energy Harvesting Using Piezoelectric and Electromagnetic technologies—State of the Art.” IEEE Transactions on Industrial Electronics 57: 850–860.10.1109/TIE.2009.2024652Suche in Google Scholar

Khan, F. U., and Izhar. 2016. “Hybrid Acoustic Energy Harvesting Using Combined Electromagnetic and Piezoelectric Conversion.” Review of Scientific Instruments 87 (025003): 1−10.10.1063/1.4941840Suche in Google Scholar PubMed

Khan, F. U., and M. U. Qadir. 2016. “State-of-the-art in Vibration-based Electrostatic Energy Harvesting.” Journal of Micromechanics and Microengineering 26: 103001−28.10.1088/0960-1317/26/10/103001Suche in Google Scholar

Khbeis, M., J. McGee, and R. Ghodssi. 2009. “Development of a Simplified Hybrid Ambient Low Frequency Low Intensity Vibration Energy Scavenger System.” Transducers 2009, Denver, CO, pp. 525−28.10.1109/SENSOR.2009.5285377Suche in Google Scholar

Kim, J., F. Cottone, S. Goyal, and J. Punch. 2010. “Energy Scavenging for energy Efficiency in Networks and Applications.” Bell Labs Technical Journal 15: 7–29.10.1002/bltj.20438Suche in Google Scholar

Larkin, M., and Y. Tadesse. 2013. “HM-EH-RT: Hybrid Multimodal Energy Harvesting from Rotational and Translational Motions.” International Journal of Smart and Nano Materials 4: 257–85.10.1080/19475411.2014.902870Suche in Google Scholar

Lee, J., J. Kim, T. Kim, M. A. Hossain, S. Kim, and J. Kim. 2016. “All-in-one Energy Harvesting and Storage Devices.” Journal of Materials Chemistry A 4 (21): 7983−99.10.1039/C6TA01229ASuche in Google Scholar

Lee, S., and A. Tovar. 2013. “Topology Optimization of Piezoelectric Energy Harvesting Skin Using Hybrid Cellular Automata.” ASME Journal of Mechanical Design 135 (3): 031001:1–11.10.1115/1.4023322Suche in Google Scholar

Lewis, J., J. Zhang, and X. Jiang. 2009. “Fabrication of Organic Solar Array for Applications in Micro-electromechanical Systems.” Journal of Renewable and Sustainable Energy 1 (013101): 1−8.10.1063/1.2998825Suche in Google Scholar

Liu, H., S. Wang, Y. Zhang, and W. Wang. 2015. “Study on the Giant Magnetostrictive Vibration-power Generation Method for Battery-less Tire Pressure Monitoring System.” Proceedings of the Institution of Mechanical engineers, Part C: Journal of Mechanical Engineering Science 229 (9): 1639–1651.10.1177/0954406214545821Suche in Google Scholar

Luo, Y., R. Gan, S. Wan, R. Xu, and H. Zhou. 2016. “Design and Analysis of a MEMS-based Bifurcate-shape Piezoelectric Energy Harvester.” AIP Advances 6 (045319): 1−9.10.1063/1.4948592Suche in Google Scholar

Madinei, H., H. H. Khodaparast, S. Adhikari, and M. I. Friswell. 2016. “A Hybrid Piezoelectric and Electrostatic Vibration Energy Harvester.” Conference Proceedings of the Society for Experimental Mechanics Series. In Shock & Vibration, Aircraft/ Aerospace,Energy Harvesting, Acoustics & Optics, Vol. 9, edited by A. Brandt, and R. Singhal, 189–195. Cham: Springer.10.1007/978-3-319-30087-0_17Suche in Google Scholar

Miles, R. W., K. M. Hynes, and I. Forbes. 2005. “Photovoltaic Solar Cells: an Overview of State-of-the-art Cell Development and Environmental Issues.” Progress in Crystal Growth and Characterization of Materials 51: 1–42.10.1016/j.pcrysgrow.2005.10.002Suche in Google Scholar

Miller, L. M., A. D. T. Elliott, P. D. Mitcheson, E. Halvorsen, I. Paprotny, and P. K. Wright. 2016. “Maximum Performance of Piezoelectric Energy Harvesters When Coupled to Interface Circuits.” IEEE Sensors 16 (12): 4803−15.10.1109/JSEN.2016.2546684Suche in Google Scholar

Mitcheson, P., P. Miao, B. Start, E. Yeatman, A. Holmes, and T. Green. 2004. “MEMS Electrostatic Micro-Power Generator for Low Frequency Operation.” Sensors and Actuators A 115: 523−29.10.1016/j.sna.2004.04.026Suche in Google Scholar

Moss, S. D., O. R. Payne, G. A. Hart, and C. Ung. 2015. “Scaling and Power Density Metrics of Electromagnetic Vibration Energy Harvesting Devices.” Smart Materials and Structures 24 (023001): 1−14.10.1088/0964-1726/24/2/023001Suche in Google Scholar

Priya, S., H. Song, Y. Zhou, R. Varghese, A. Chopra, S. Kim, I. Kanno, L. Wu, D. S. Ha, J. Ryu, and R. G. Polcawich. 2017. “A Review on Piezoelectric Energy Harvesting: Materials, Methods, and Circuits.” Energy Harvesting and Systems 4 (1): 3–39.10.1515/ehs-2016-0028Suche in Google Scholar

Rahman, M. S. 2016. “A Hybrid Technique of Energy Harvesting from Mechanical Vibration and Ambient Illumination.” Master’s thesis, Louisiana, USA: University of New Orleans.Suche in Google Scholar

Rahman, M. S., and U. K. Chakravarty. 2018. “A Hybrid Energy Harvesting System Based on Solar Radiation and Mechanical Vibration.” Proceedings of. ASME 2018 International Mechanical Engineering Congress & Exposition, Vol. 1, Advances in Aerospace Technology, Pittsburgh, Pennsylvania, USA, pp. 1–12.10.1115/IMECE2018-86928Suche in Google Scholar

Rahman, M. S., P. Sarker, and U. K. Chakravarty. 2019. “A Hybrid Energy Harvester Based on Solar Radiation and Mechanical Vibration.” AIP Conference Proceedings 2121 (1): 120008:1–6.10.1063/1.5115945Suche in Google Scholar

Rantz, R., and S. Roundy. 2017. “Characterization of Real-world Vibration Sources and Application to Nonlinear Vibration Energy Harvesters.” Energy Harvesting and Systems 4 (2): 67–76.10.1515/ehs-2016-0021Suche in Google Scholar

Roundy, S. 2003. “Energy Scavenging for Wireless Sensor Nodes with a Focus on Vibration to Electricity Conversion.” Ph.D. thesis, Berkeley: University of California.Suche in Google Scholar

Roundy, S., P. Wright, and J. Rabaey. 2003. Energy Scavenging for Wireless Sensor Networks. Boston: Kluwer Academic Publishers.10.1007/978-1-4615-0485-6Suche in Google Scholar

Sample, A., and J. R. Smith. 2009. “Experimental Results with Two Wireless Power Transfer Systems.” IEEE Radio and Wireless Symposium, San Diego, CA, pp. 16–18.10.1109/RWS.2009.4957273Suche in Google Scholar

Shan, X. B., S. W. Guan, Z. S. Liu, Z. L. Xu, and T. Xie. 2013. “A New Energy Harvester Using A Piezoelectric and Suspension Electromagnetic Mechanism.” Journal of Zhejiang University-Science A (Applied Physics and Engineering) 14 (12): 890−97.10.1631/jzus.A1300210Suche in Google Scholar

Sharma, H., A. Haque, and Z. A. Jaffery. 2018. “Solar Energy Harvesting Wireless Sensor Network Nodes: A Survey.” Journal of Renewable and Sustainable Energy 10 (2): 023704:1–33.10.1063/1.5006619Suche in Google Scholar

Sivula, K. 2015. “Toward Economically Feasible Direct Solar-to-fuel Energy Conversion.” The Journal of Physical Chemistry Letters 6: 975–976.10.1021/acs.jpclett.5b00406Suche in Google Scholar PubMed

Sood, V. K., and P. Bhalla. 2013. “EMTP Model of Grid Connected PV System.” Proceedings of International Conference on Power System Transients, Vancouver, Canada, pp. 1−8.Suche in Google Scholar

Su, D., R. Zheng, K. Nakano, and M. P. Cartmell. 2016. “Stabilisation of the High Energy Orbit for a Nonlinear energy Harvester with Variable Damping.” Proceedings of the Institution of Mechanical engineers, Part C: Journal of Mechanical Engineering Science 230 (12): 2003–2012.10.1177/0954406215590169Suche in Google Scholar

Tao, K., J. Miao, S. Lye, and X. Hu. 2015. “Sandwich-structured Two-dimensional MEMS Electret Power Generator for Low-level Ambient Vibrational Energy Harvesting.” Sensors and Actuators A: Physical 228: 95–103.10.1016/j.sna.2015.02.021Suche in Google Scholar

Wang, T., T. Kobayashi, and C. Lee. 2015. “Micromachined Piezoelectric Ultrasonic Transducer with Ultra-wide Frequency Bandwidth.” Applied Physics Letters 106: 013501−5.10.1063/1.4905441Suche in Google Scholar

Waterbury, A. C., and P. K. Wright. 2013. “Vibration Energy Harvesting to Power Condition Monitoring Sensors for Industrial and Manufacturing Equipment.” Proceedings of the Institution of Mechanical engineers, Part C: Journal of Mechanical Engineering Science 227 (6): 1187–1202.10.1177/0954406212457895Suche in Google Scholar

Wurpts, W., J. Twiefel, and F. Brouet. 2017. “Equivalent Circuit Parametrization Utilizing FE Model Order Reduction and Its Application to Piezoelectric Generators and Actuators.” Energy Harvesting and Systems 4 (3): 115–129.10.1515/ehs-2017-0002Suche in Google Scholar

Xu, Z., X. Shan, D. Chen, and T. Xie. 2016. “A Novel Tunable Multi-Frequency Hybrid Vibration Energy Harvester Using Piezoelectric and Electromagnetic Conversion Mechanisms.” Applied Science 6 (10): 1−16.10.3390/app6010010Suche in Google Scholar

Xu, Z., X. Shan, H. Yang, W. Wang, and T. Xie. 2017. “Parametric Analysis and Experimental Verification of a Hybrid Vibration Energy Harvester Combining Piezoelectric and Electromagnetic Mechanisms.” Micromachines 8 (189): 1–19.10.3390/mi8060189Suche in Google Scholar

Yang, B., C. Lee, L. W. Kee, and P. S. Lim. 2010. “Hybrid Energy Harvester Based on Piezoelectric and Electromagnetic Mechanisms.” Journal of Micro/Nanolithography, MEMS and MOEMS 9 (023002): 1−10.10.1117/1.3373516Suche in Google Scholar

Yi, J. W., W. Y. Shih, and W. H. Shih. 2002. “Effect of length, width, and Mode on the Mass Detection Sensitivity of Piezoelectric Unimorph Cantilevers.” Journal of Applied Physics 91: 1680−86.10.1063/1.1427403Suche in Google Scholar

Yildirim, T., M.H. Ghayesh, T. Searle, W. Li, and G. Alici. 2017. “A Parametrically Broadband Nonlinear Energy Harvester.” ASME Journal of Energy Resources Technology 139: 032001:1–8.10.1115/1.4034514Suche in Google Scholar

Yu, H., Q. Yue, J. Zhou, and W. Wang. 2014. “A Hybrid Indoor Ambient Light and Vibration Energy Harvester for Wireless Sensor Nodes.” Sensors 14: 8740−55.10.3390/s140508740Suche in Google Scholar PubMed PubMed Central

Yuan, Y., M. Liu, W. Tai, and L. Zuo. 2018. “Design and Treadmill Test of a Broadband Energy Harvesting Backpack with a Mechanical Motion Rectifier.” ASME Journal of Mechanical Design 140 (8): 085001:1–8.10.1115/1.4040172Suche in Google Scholar

Zhang, H. 2011. “Power Generation Transducer from Magnetostrictive Materials.” Applied Physics Letters 98 (232505): 1−7.10.1063/1.3597222Suche in Google Scholar

Zhang, Y. 2014. “Piezoelectric Based Energy Harvesting on Low Frequency Vibrations of Civil Infrastructures.” Ph.D, Baton Rouge, LA: Lousiana State University.Suche in Google Scholar

Zhu, J., N. S. Yuksek, M. Almasri, and Z. Feng. 2019. “Numerical Modeling of Dynamic Response of Miniature Multi-impact Electromagnetic Device for Low and Wide Range Frequencies Energy Harvesting.” Proceedings of the Institution of Mechanical engineers, Part C: Journal of Mechanical Engineering Science 233 (7): 2400–2409.10.1177/0954406218791638Suche in Google Scholar

Published Online: 2019-12-17
Published in Print: 2021-03-26

© 2019 Walter de Gruyter Inc., Boston/Berlin

Heruntergeladen am 10.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/ehs-2019-0006/html
Button zum nach oben scrollen