1. Introduction
Vortex bladeless wind turbines (VBWTs) are vibration-based energy harvesters that have been successfully demonstrated to scavenge wind energy [
1]. As the name implies, one of the main features of this type of turbine is that they do not have rotating blades or moving parts. As such, they are considered an environmentally friendly, economical, and efficient alternative to traditional propeller-type turbines [
2]. In comparison to horizontal and vertical axis wind turbines, VBWTs can generate power at lower wind speeds, which, in turn, expands their operating range and improves their power density.
The turbine is excited to oscillate due to the vortex-shedding force induced on the body mast. Once the frequency of the shedding force is close to the fundamental frequency of the turbine, the lock-in phenomenon is initiated [
3]. At this particular point, the mast will oscillate at the shedding frequency of the flow vortices, which is directly proportional to the wind speed. This phenomenon allows expanding the turbine bandwidth by sustaining the oscillation between the turbine structure and the shedding vortices unaltered beyond the fundamental frequency of the turbine. In general, the maximum amount of energy is extracted from the wind within the lock-in region [
4]. However, the upper bound at which the lock-in region collapses is determined by the mass ratio of the mast to the added mass by the airflow. The bandwidth of the lock-in region is narrowed down as the ratio increases (i.e., added mass increases with the increasing speed of the wind) [
3].
The lock-in region is also narrow-banded because the turbine structure is often unchanged and damped. There have been many research attempts to improve and optimize the output power, bandwidth, and efficiency of the wind turbine. Yazdi [
5,
6] proposed a tuning mechanism with a linear actuator and an optimal control strategy to slide a mass inside the hollow mast of the piezoelectric-based turbine; as such, the structure resonance could be regulated close to the shedding frequency at a different airflow speed. At a wind speed of 1.6 m/s, the numerical results suggested an increase in the 6 m long turbine bandwidth and power of 190% and 294%, respectively. This method, however, is complex and requires external power to operate the actuator and the controller. Meliga et al. [
7] presented a feedback control approach that utilizes an actuator located in the cylindrical mast of the turbine. The actuator ejects a controlled stream of flow at an optimized velocity, position, and angle to suppress the vibration of the cylinder to remain within the limit cycle orbits of the shedding vortices of the flow. This control approach leads to a 3.5% increase in the harvested energy. Using this approach, the system has more robustness against small inaccuracies of the structural parameters.
Optimizing the shape of the body mast is another method to increase the harnessed power. Ding et al. [
8] numerically investigated several cross-sectional shapes of the body mast. A quasi-trapezoid shape (i.e., a shape between square and triangle) generates more power than the other designs, including triangular prism, square cylinder, and circular cylinder. However, for VBWTs, the cylindrical body mast remains preferred over the other designs due to its less sensitivity to air flow direction. Square cylinders are best used for energy harvesters utilizing the galloping effect of the flow. On the other hand, Villarreal et al. [
9,
10] introduced an arrangement of two pairs of annular magnets to the structure of the VBWT. The magnets have opposite polarities to repel each other to form a repulsive restoring force. This nonlinear restoring force tunes the overall stiffness of the structure. As the wind speed increases, the structure becomes stiffer, and, thus, the resonance frequency increases to match the shedding frequency of the flow at higher wind speeds. This method, in turn, widens the lock-in region of the turbine.
In this research work, a tunable mechanism based on a progressive-rate spring is presented. The spring stiffness varies as the wind speed changes. The spring is attached to a permanent magnet, allowing it to oscillate vertically inside the electromagnetic coil. The key feature of this mechanism is to keep the fundamental frequency of the turbine locked with the shedding frequency of the flow. As such, the lock-in phenomenon is established at a wider range of wind speeds. In conventional VBWTs, the magnet or the coil usually moves in an arched profile relative to each other. In this proposed configuration, an extra degree of freedom is added. The magnet is allowed to oscillate in the longitudinal direction of the mast body. The principal motive of this research work is to explore the feasibility of utilizing a passive tunable progressive-rate spring. To quantify the harvested power characteristics of the turbine, a mathematical model describing the nonlinear coupled dynamics of the turbine is presented in
Section 3. The proposed design structure is introduced next.
2. Turbine Design and Structure
In general, bladeless wind turbines are based on piezoelectric [
5] and/or electromagnetic [
9] harvesting units. Salvador et al. [
11] experimentally investigated two configurations of magnets in a harvester, namely disc-shaped and arc-shaped. In this design, the magnets were attached to the base of the mast body that was pivoted to swing as a pendulum. Results show that the arc-shaped magnet yields a higher voltage output. This might be attributed to the larger size of the arc-shaped magnet tested in the study compared with the disc-shaped magnet. Villarreal [
12] proposed a set of magnets arranged around the circular circumference of the turbine base. Each magnet is placed between two coils, which are attached to the mast body. As the mast oscillates, the coils move relatively to the magnets to generate an induced electrical power. Sets of magnets and coils are stacked above each other to generate more power. Gautam et al. [
13] compared two configurations of magnet field arrangements. The first is to arrange the magnets in a circular pattern, and the other is to arrange them radially. It was observed that the magnetic flux change is enhanced in the radial configuration for small vibrations. This research also investigated stacking the magnets in layers. It was concluded that two opposite radial magnets per layer stacked vertically in three layers would further improve the flux change rate. The layers were oriented 120° in-plane relative to each other. Moreover, five different coil configurations were proposed in [
14] to improve the field gradient for small vibrations. Each design showed a maximum field density at a different location from the mean or center position of the oscillating mast body. The coil configurations were mainly made of four Golay coils that are generally used to create a magnetic field gradient perpendicular to the main magnetic field. A flux density variation of 900 mT was achieved by the optimum design of the coil configuration.
The structure of the proposed vortex bladeless wind turbine (VBWT) is presented in
Figure 1. The turbine comprises three main components: a cantilever beam, a cylindrical mast body, and a harvesting unit. The underlying concept of the VBWT is to capture the vortex-induced vibration of the body mast attached to a vertical flexible cantilever beam using a permanent magnet (PM) oscillating inside a multi-layer coil. Among others, this method has been proven to provide high electromechanical energy conversion efficiency, energy density, and high voltage [
13]. The electromagnetic harvesting unit is hosted inside the turbine to harness the mechanical energy of the vibrating mast into electrical energy based on Faraday’s law of induction. The mechanical vibration of the mast is harnessed to electrical energy due to the relative motion between the magnet and the coil. In order to provide an efficient harvesting mechanism, the coil is fixed to the stationary body of the turbine, and the magnet is attached to the vibrating body mast. Using this configuration, the permanent magnet will have a circular arc (2D) motion relative to the coil. Moreover, in this unique design, the magnet is connected to a spring, which allows it to slide in the longitudinal direction of the body mast and the coil. A rigid guide rod is used to confine the lateral and rotary motion of the magnet not to introduce instability during the operation of the turbine.
The tuning mechanism that changes the fundamental frequency of the turbine as the wind speed changes can be achieved by altering the intrinsic stiffness of the spring attached to the sliding permanent magnet of the harvester. It will be shown later in the following sections that the fundamental frequency of the turbine is a function of the spring stiffness. As such, the lock-in phenomenon can be initiated theoretically at any wind speed by tuning the fundamental frequency of the turbine (i.e., by changing the spring stiffness) to match the shedding frequency of the flow at a given wind speed. Progressive-rate springs such as Belleville [
15,
16], wave [
17], or variable-pitch [
18] springs can be selectively engineered to offer the displacement–force characteristics needed for the tunable mechanism of the turbine. The dynamic model of this turbine is presented in the next section.
3. Dynamic Modeling
In this section, the nonlinear dynamic modeling of the proposed design of the tunable bladeless turbine is developed. Assuming the fundamental frequency of the turbine dominates the dynamics during the operation, the turbine can be schematically presented by the lumped-mass configuration depicted in
Figure 2. The permanent magnet (PM) of the harvester, which translates and swings simultaneously, is considered a point mass (
) oscillating on an inverted pendulum mounted on a cart. The cart with mass (
), on the other hand, represents the equivalent oscillation motion of the free end of the cantilever beam.
and
symbolize the equivalent stiffness of the beam and the tunable spring stiffness, respectively. The turbine is driven by a vortex-induced force that is represented by
. An electromagnetic coil is introduced to harness the PM motion to electrical power. The resistor (
) connected to the coil imitates the equivalent external load or circuit attached to the turbine.
Without loss of generality, the mast body is assumed to be made of a lightweight material, so its weight and rotary inertia are negligible. The magnet is confined to slide over a guide rod to maintain its motion in the longitudinal direction of the mast body while it is oscillating. As such, the rotational angle (
) of the oscillating magnet is equal to the rotational angle of the mast body which, in turn, is identical to the free-end angle of the beam. Moreover, the free-end angle (
) of the beam is related to its free-end deflection (
). The relationship is derived following the assumptions of the linear transverse vibrations of the undamped Euler–Bernoulli beam. The governing equation of motion of the cantilever beam carrying a tip mass is obtained as
where
is the transverse displacement of the beam at position
y along the cantilever beam.
and
are the bending stiffness and the mass per unit length of the beam, respectively. The boundary condition of forces at the free end of the beam carrying a tip mass of
(the permanent magnet) is expressed as
Using the Galerkin procedure [
19] to discretize Equation (2), the transverse displacement (
) can be approximated by introducing spatial and temporal functions as
where
(
= 1, 2, 3 …,
n) are the mode shapes of the clamped-free cantilever beam. Following our assumption that the fundamental frequency is dominant, then only the first mode can be considered in this analysis. The first mode is expressed as [
20]
in which,
is the length of the beam and
Assuming the rotary inertia of the tip mass is negligible, the eigenvalue of the first vibration mode denoted by
is calculated from the characteristic equation
For a lumped-mass representation of the turbine, the free-end displacement (
) and rotation (
) of the beam are expressed as
and
, respectively. Defining the parameter
as the ratio of the rotation (
) over the displacement (
) results in
The superscript (
) represents the derivative with respect to
. As it is clearly seen, the parameter
can be determined at a given magnet mass (
), beam length (
), and mass per unit length of the beam (
) by first calculating the eigenvalue (
) from Equation (6) and then substituting it into Equation (5) to calculate the modal parameter (
). Once
and
are known, Equation (7) is then used to determine the ratio
. It is worth noting that using the guide rod imposes restrictions on the rotational motion of the magnet that is now prescribed or constrained by the motion of free-end displacement of the cantilever beam, or vice versa, as described by Equation (7). The constraint equation that restricts the rotational motion of the magnet is given by
The equations of motion of the turbine are developed using the Lagrange formulation. Lagrange’s multiplier (
) [
21,
22] is introduced to impose the constraint rotational motion of the magnet. The turbine equations of motion are formulated by implementing the Lagrange given equation [
22]
where
denotes the Lagrangian that is the difference between the kinetic (
) and potential (
) energies of the system. For
,
and
are the generalized coordinates and forces of the system, respectively.
represents Rayleigh’s dissipation function of the turbine. The last RHS term of Equation (9) is the constraint force applied to the rotational motion of the magnet. By defining four generalized coordinates, which are beam deflection (
), mast rotation angle (
), magnet sliding motion (
) relative to the mast body, and output charge (
) of the electromagnetic coil, the kinetic, potential, and dissipative energies of the turbine have the form:
Here,
and
are the electromagnetic coil inductance and resistance, respectively.
is the initial length of the tunable spring (
).
is the acceleration due to gravity (9.81
). The last term in Equation (10) represents the coupling energy between the mechanical and electrical domains [
23]. The electromagnetic coupling factors in the
- and
-directions are denoted by
and
, respectively.
In addition to the above, the generalized forces (
) are represented only by the lift force in the
-direction (i.e.,
). The lift force due to the vortex shedding acting on the body mast of the turbine given by [
24]
where,
is the density of the air.
and
are the diameter and length of the mast body.
is the lifting coefficient that on average equals 0.50 for Reynold’s numbers between 10
3 and 10
5 [
25].
denotes the wind speed, and
is the shedding frequency in Hz that is analytically presented with the following expression [
26]
where
is the Strouhal number (~0.20) [
26] that represents a measure of the ratio of the inertial forces of the flow to the inertial forces of the convective acceleration. In this preliminary study, increasing the output power of the turbine is the primary interest, so the dynamics within the lock-in region that is associated with the bandwidth of the turbine are not considered.
Substituting Equations (10)–(13) into Lagrange’s Equation (9), the equations of motion are readily obtained. However, the equations of motion can be further simplified. For small angles, the trigonometric functions are approximated as and . Further, for an infinitesimal oscillation of the beam, the centrifugal force term () is neglectable. By doing this, the system equations of motion become
The
-direction equation:
The
-direction equation:
The
-direction equation:
The electromagnetic circuit equation:
Moreover, Equation (8) yields the fifth EOM that is needed to describe the full dynamics of the turbine, including the four generalized coordinates and Lagrange’s multiplier. In general, if the constraint equation (
) is explicitly a function of time and the generalized coordinates vector (
) such that
, then taking the second time derivative of the constraint equation yields [
27]
where
is the Jacobian matrix of the constraint equation associated with the generalized coordinates that is
is the vector of the time-partial derivative of the Jacobian matrix such that
is the vector of partial derivative of the term
calculated as
and
is the second time-partial derivative of the constraint equation. In the preceding equation and for our system under the investigation, only the first term remains nonzero. Thus, Equation (19) reduces to
Next, the electromagnetic circuit model parameters, including the coil inductance, resistance, and coupling factors, can be estimated using closed-form approximations. For instance, the coil inductance (
) in Henry is calculated using Wheeler’s spiral equation [
28]
in which,
is the number of turns of the coil.
,
, and
are the height, width, and mean diameter of the multilayer coil. The equation is accurate for
. The coil resistance (
) in Ohms is estimated using [
29]:
is the copper wire resistivity (i.e., 1.68 × 10
−8 Ωm [
30]), and
is the wire diameter of the coil. On the other hand, the coupling factors
and
are calculated based on Faraday’s law of induction [XX15]. The induced electromotive force (
) is described by
where
is the total flux of the coil. In our analysis, the coupling factors are constant and evaluated at the maximum flux point. The coupling factor in
-direction is given by [
23]
is the residual magnetic flux density of the magnet.
is the volume of the magnet. It can be shown from the analysis conducted by Behtouei et. al. [
31] that the coupling factor in the
-direction is approximately twice that of the
-direction. For a sinusoidal output current at steady state, the output rms electrical power (
) of the turbine is calculated as
To perform the dynamic analysis of the turbine, the coupled equations of motion are numerically solved in the next section using the MATLAB
® ode45 solver [
32].