Top Banner
AbstractFunctional electrical stimulation (FES) has become a popular and successful form of rehabilitation for people with neurological conditions such spinal cord injury or stroke. FES is the direct application of electrical current across the motor neurons of a muscle to generate artificial muscle contractions to perform functional tasks. Arm-cycling is a beneficial rehabilitation activity and motivation exists to use FES in arm- cycling. It has previously been shown that closed-loop FES control can be designed to achieve accurate, repetitive motion. In a previous study, a robust sliding-mode controller was used to track a desired crank cycle cadence (crank velocity) on a decoupled hand cycle. For this thesis the hand-cycle was modified to yield an improved tracking performance. Using a combination of FES and volition, experimental results from three able-bodied participants are presented for the developed control system. Two protocols were ran, one of which activated the motor for all time and demonstrated an average cadence tracking error of -0.04 ± 2.83 revolutions per minute (RPM) for a desired cadence of 40 RPM. The second protocol deactivated the motor when the bicep was stimulated and demonstrated an average cadence tracking error of -0.02 ± 5.35 RPM. Index TermsFunctional electrical stimulation (FES), Lyapunov, rehabilitation robotics, switched systems, stability I. INTRODUCTION UNCTIONAL electrical stimulation (FES) has been used for decades as a form of rehabilitation for individuals with upper motor neuron lesions, caused by an event such as a stroke or spinal cord injury (SCI) [1]. Neuromuscular electrical stimulation (NMES) is called FES when it is used to perform functional tasks [2]. NMES is the application of electrical current across muscle fibers to yield artificial muscle contractions [2]. FES has commonly been used for rehabilitation because it helps restore voluntary muscle function, increases bone mineral density and muscle mass, improves cardiovascular health, and decreases spasticity [1]. This can also lead to psychological benefits such as improved independence and self-esteem [1]. Enhancing a person’s physiological (skeletal, cardiopulmonary and muscular systems) and psychological well-being may enable them to improve their ability to perform activities of daily living [3]. In research, FES applied to the lower extremities is generally divided into three categories: cycling, standing, and walking. FES-cycling has an advantage over standing and walking exercises since people with paralysis can perform cycling [3]. Regular FES cycling has been shown to improve general health, decrease the possibility of cardiovascular disease, and diminish the effects of secondary complications of SCIs [4]. The FES stimulation intensity, in past studies, has been controlled using proportional-derivative feedback or open-loop control to obtain a desired cycling cadence (crank velocity) [1]. Cadence tracking was used as the primary control objective because it is one of the most important factors in cycling rehabilitation [1]. Furthermore, the stimulation was switched between multiple muscle groups to achieve a predetermined stimulation pattern throughout the crank cycle [1]. Thus, FES cycling systems can be seen as switched control systems with autonomous, state-dependent switching. In addition, FES control input is commonly not applied during kinematic dead zones in the crank cycle, where only a small portion of the torque produced by the cyclist’s muscles is transferred as torque about the crank axis [5]. In a previous study, electric motors were used throughout the entire crank cycle to help maintain the desired cadence and account for the kinematic dead zones [4]. Recent studies have also implemented Lyapunov-based nonlinear control techniques to improve the performance and effectiveness of FES cycling [1]. A recent study successfully developed a controller that switches the control input between an electric motor and FES of various muscle groups [5]. Using the electric motor only during kinematic dead zones when FES control input is not used, can maximize the contribution of the cyclist’s muscles [5]. However, there are several challenges with using FES control. One challenge is that stimulated muscles fatigue at a higher rate than muscles that contract voluntarily [4]. The muscles fatigue at a high rate because the stimulation only engages small portions of each muscle, which also makes controlling the muscles difficult [6]. Electric motors can be used to compensate for muscle fatigue so that the person can continue performing the exercise even while fatigued and without stimulation [4]. Another challenge that exists relates to the repeatability of the stimulation because the effectiveness of the stimulation is influenced by many factors such as muscle fatigue, body fat, body hydration, and the correlation between stimulation and muscle reaction [6]. Electrode placement is another factor because there is no guarantee the electrode will be placed at the same location every time or in the best position for stimulation [2]. Based on the success of FES cycling for the lower extremities, there is evidence that FES cycling can also be applied to the upper extremities. Individuals with SCIs above Functional Electrical Stimulation Induced Cadence Control of a Hand Cycle Michael Woc, Christian Cousin, Brendon Allen, Warren E. Dixon F
8

Functional Electrical Stimulation Induced Cadence Control ... · neurological conditions such spinal cord injury or stroke. FES is the direct application of electrical current across

Oct 20, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • Abstract—Functional electrical stimulation (FES) has become a

    popular and successful form of rehabilitation for people with

    neurological conditions such spinal cord injury or stroke. FES is

    the direct application of electrical current across the motor

    neurons of a muscle to generate artificial muscle contractions to

    perform functional tasks. Arm-cycling is a beneficial

    rehabilitation activity and motivation exists to use FES in arm-

    cycling. It has previously been shown that closed-loop FES control

    can be designed to achieve accurate, repetitive motion. In a

    previous study, a robust sliding-mode controller was used to track

    a desired crank cycle cadence (crank velocity) on a decoupled hand

    cycle. For this thesis the hand-cycle was modified to yield an

    improved tracking performance. Using a combination of FES and

    volition, experimental results from three able-bodied participants

    are presented for the developed control system. Two protocols

    were ran, one of which activated the motor for all time and

    demonstrated an average cadence tracking error of -0.04 ± 2.83 revolutions per minute (RPM) for a desired cadence of 40 RPM.

    The second protocol deactivated the motor when the bicep was

    stimulated and demonstrated an average cadence tracking error

    of -0.02 ± 5.35 RPM.

    Index Terms—Functional electrical stimulation (FES),

    Lyapunov, rehabilitation robotics, switched systems, stability

    I. INTRODUCTION

    UNCTIONAL electrical stimulation (FES) has been used

    for decades as a form of rehabilitation for individuals with

    upper motor neuron lesions, caused by an event such as a stroke

    or spinal cord injury (SCI) [1]. Neuromuscular electrical

    stimulation (NMES) is called FES when it is used to perform

    functional tasks [2]. NMES is the application of electrical

    current across muscle fibers to yield artificial muscle

    contractions [2]. FES has commonly been used for

    rehabilitation because it helps restore voluntary muscle

    function, increases bone mineral density and muscle mass,

    improves cardiovascular health, and decreases spasticity [1].

    This can also lead to psychological benefits such as improved

    independence and self-esteem [1]. Enhancing a person’s

    physiological (skeletal, cardiopulmonary and muscular

    systems) and psychological well-being may enable them to

    improve their ability to perform activities of daily living [3].

    In research, FES applied to the lower extremities is generally

    divided into three categories: cycling, standing, and walking.

    FES-cycling has an advantage over standing and walking

    exercises since people with paralysis can perform cycling [3].

    Regular FES cycling has been shown to improve general health,

    decrease the possibility of cardiovascular disease, and diminish

    the effects of secondary complications of SCIs [4].

    The FES stimulation intensity, in past studies, has been

    controlled using proportional-derivative feedback or open-loop

    control to obtain a desired cycling cadence (crank velocity) [1].

    Cadence tracking was used as the primary control objective

    because it is one of the most important factors in cycling

    rehabilitation [1]. Furthermore, the stimulation was switched

    between multiple muscle groups to achieve a predetermined

    stimulation pattern throughout the crank cycle [1]. Thus, FES

    cycling systems can be seen as switched control systems with

    autonomous, state-dependent switching.

    In addition, FES control input is commonly not applied

    during kinematic dead zones in the crank cycle, where only a

    small portion of the torque produced by the cyclist’s muscles is

    transferred as torque about the crank axis [5]. In a previous

    study, electric motors were used throughout the entire crank

    cycle to help maintain the desired cadence and account for the

    kinematic dead zones [4]. Recent studies have also

    implemented Lyapunov-based nonlinear control techniques to

    improve the performance and effectiveness of FES cycling [1].

    A recent study successfully developed a controller that switches

    the control input between an electric motor and FES of various

    muscle groups [5]. Using the electric motor only during

    kinematic dead zones when FES control input is not used, can

    maximize the contribution of the cyclist’s muscles [5].

    However, there are several challenges with using FES

    control. One challenge is that stimulated muscles fatigue at a

    higher rate than muscles that contract voluntarily [4]. The

    muscles fatigue at a high rate because the stimulation only

    engages small portions of each muscle, which also makes

    controlling the muscles difficult [6]. Electric motors can be

    used to compensate for muscle fatigue so that the person can

    continue performing the exercise even while fatigued and

    without stimulation [4]. Another challenge that exists relates to

    the repeatability of the stimulation because the effectiveness of

    the stimulation is influenced by many factors such as muscle

    fatigue, body fat, body hydration, and the correlation between

    stimulation and muscle reaction [6]. Electrode placement is

    another factor because there is no guarantee the electrode will

    be placed at the same location every time or in the best position

    for stimulation [2].

    Based on the success of FES cycling for the lower

    extremities, there is evidence that FES cycling can also be

    applied to the upper extremities. Individuals with SCIs above

    Functional Electrical Stimulation Induced

    Cadence Control of a Hand Cycle Michael Woc, Christian Cousin, Brendon Allen, Warren E. Dixon

    F

  • 2

    2

    Th1 (thoracic vertebrae 1) suffer from upper extremity

    impairments and would benefit from upper limb rehabilitation

    [7]. The aforementioned methods for FES cycling in the lower

    extremities can be applied to FES cycling with the upper limbs.

    Preliminary work such as characterizing the dynamics of an

    arm-cycle system and defining the arm stimulation regions have

    been performed in [6], [8]. This thesis seeks to improve the

    performance (crank cycle cadence tracking) of the controller

    developed in [6] by improving the cycle hardware (i.e., the

    electric motor).

    In this thesis, a switched-systems sliding mode controller

    was developed and used on a FES arm-cycle system with

    electric motor assistance. The cadence was tracked during

    experiments and the cadence tracking error was used to quantify

    the performance of the controller. Preliminary experiments

    were conducted on three able-bodied individuals to evaluate the

    performance of the controller. FES cycling rehabilitation is

    intended for people with neurological conditions (NCs), but for

    their safety the arm-cycle system was first tested on able-bodied

    individuals. Two different protocols were implemented for the

    experiments, Protocol 1 and Protocol 2. Both protocols had

    each participant cycle at a specific RPM (revolutions per

    minute). In Protocol 1, the motor was on during the entire crank

    cycle and the participant only added volition if the cadence error

    exceeded ±5 RPM. In Protocol 2, the motor was deactivated during bicep stimulation and the participant was never allowed

    to add volition. In the subsequent sections the arm-cycle

    dynamic model, the developed switched-systems sliding mode

    controller, and the resulting stability analysis are provided.

    II. MODEL

    A. Arm-Cycle Dynamic Model

    A person pedaling a motor assisted arm-cycle can be

    modeled as single degree-of-freedom system [6], [8], which can

    be expressed as:

    𝑀(𝑞)�̈� + 𝑉𝑚(𝑞, �̇�)�̇� + 𝐺(𝑞) + 𝑃(𝑞, �̇�) + 𝑐𝑑�̇� + 𝑑(𝑡)

    = 𝜏𝑚(𝑞, �̇�, 𝑡) + 𝜏𝑒(𝑞, �̇�, 𝑡) (1)

    where 𝑞 and �̇� denote the measured crank angle and velocity of the cycle, respectively; �̈� denotes the unmeasured crank acceleration; 𝑀 is the inertia matrix; 𝑉𝑚 represents the centripetal and Coriolis effects; 𝐺 denotes the gravitational effects; 𝑃 accounts for the viscoelastic tissue forces; 𝑐𝑑 is the coefficient of viscous damping, and 𝑑 accounts for time-varying disturbances. The torque produced by the upper limb

    muscle contractions of the individual is expressed as:

    𝜏𝑚(𝑞, �̇�, 𝑡) = ∑ 𝐵𝑚𝑢𝑚𝑚∈ℳ

    (2)

    where 𝐵𝑚 denotes the uncertain, nonlinear control effectiveness term that relates the stimulation input to the torque output of a

    muscle group. The term 𝑢𝑚 is the input stimulation intensity applied to each muscle group. The subscript 𝑚 indicates the muscle group and is defined as 𝑚 ∈ ℳ ≜{𝑅𝐵𝑖, 𝐿𝐵𝑖, 𝑅𝑇𝑟𝑖, 𝐿𝑇𝑅𝑖}. 𝐵𝑖 and 𝑇𝑟𝑖 represent the biceps brachii and triceps brachii muscle groups, respectively; 𝑅 and 𝐿 correspond to right and left arm, respectively. The torque

    applied about the crank cycle axis by the electric motor is

    represented by:

    𝜏𝑒(𝑡) = 𝐵𝑒𝑢𝑒 (3) where 𝐵𝑒 is a known positive constant term that relates the motor’s applied current to the resultant torque. 𝐵𝑒 can be bounded as 0 < 𝐵𝑒 ≤ 𝑐𝑒. The current applied to the motor is represented by 𝑢𝑒.

    B. Switched System Model

    As previously mentioned, the stimulation needs to switch

    between different muscle groups in many FES applications to

    produce coordinated motion, such as cycling. Switched systems

    have additional control challenges since switching between

    subsystems, such as muscles and electric motors, can generate

    instabilities in the overall system and it is necessary to prove

    the stability of the overall switched system, even if the

    subsystems are stable. The stability analysis for the arm-cycle

    system used in this thesis was developed similar to that of

    previous studies done on a recumbent stationary cycle [1], [5].

    In both studies, the stimulation was not applied during

    kinematic dead zones. In [5], it was shown that coupling a

    motor to the cycle increased the controllability of the dead

    zones. As a result, the arm-cycle system was analyzed as a

    switched system with control input switching between

    stimulation of the biceps and triceps muscle groups and the

    electric motor.

    The stimulation regions for each muscle were defined by a

    range of crank angles in the crank cycle that allow for that

    muscle to produce positive torque about the crank axis. The

    motor-controlled regions are the kinematic dead zones or

    kinematically inefficient regions for any of the muscles to

    produce positive torque. Let the set of crank angles (ℚ ) vary from [0, 2π), the crank cycle regions where stimulation is

    applied to each muscle be represented by ℚ𝑚 ⊂ ℚ, and the regions of the crank cycle where the electric motor is used be

    ℚ𝑒 ⊂ ℚ. The union of crank regions where FES is applied is denotes as 𝑄𝐹𝐸𝑆 is defined as:

    ℚ𝐹𝐸𝑆 ≜ ⋃ ℚ𝑚𝑚∈ℳ

    (4)

    The 𝑄𝑒 regions can be expressed as: ℚ𝑒 ≜ ℚ\QFES (5) A graphical representation of the kinematic dead zones (also

    known as ℚ𝑒 regions in this thesis) and the stimulation regions, ℚ𝑚, using crank angles over the crank cycle is in Fig. 1.

    Fig. 1. Sample schematic showing the arm muscle stimulation regions for the biceps (red region) and the triceps (green region). The kinematic dead zones

    (KDZ) are also displayed with the gray areas [6].

  • 3

    3

    The control inputs for the muscle stimulation (𝑢𝑚) and current input (𝑢𝑒) for the electric motor are represented as:

    𝑢𝑚 ≜ 𝑘𝑚𝜎𝑚𝑢𝑠 (6)

    𝑢𝑒 ≜ 𝑘𝑒𝜎𝑒𝑢𝑐 (7)

    where 𝑢𝑠, 𝑢𝑐 are subsequently designed control inputs and 𝑘𝑚, 𝑘𝑒 are positive, constant control gains. A switching signal for the muscle groups (𝜎𝑚) and the motor (𝜎𝑒) are defined as:

    𝜎𝑚 ≜ {

    1,0,

    𝑖𝑓 𝑞 ∈ ℚ𝑚 𝑖𝑓 𝑞 ∉ ℚ𝑚

    (8)

    𝜎𝑒 ≜ {

    1,0,

    𝑖𝑓 𝑞 ∈ ℚ𝑒 𝑖𝑓 𝑞 ∉ ℚ𝑒

    (9)

    Switched control effectiveness terms 𝐵𝑀 and 𝐵𝐸 are introduced to rewrite the dynamics of the overall arm-cycle

    system. 𝐵𝑀 is defined as:

    𝐵𝑀 = ∑ 𝐵𝑚𝑘𝑚𝜎𝑚𝑚∈ℳ

    (10)

    and 𝐵𝐸 is defined as:

    𝐵𝐸 = ∑ 𝐵𝑒𝑘𝑒𝜎𝑒𝑚∈ℳ

    (11)

    The arm-cycle system dynamics can be rewritten, accounting

    for the switching signals, as:

    𝑀�̈� + 𝑉𝑚�̇� + 𝐺 + 𝑃 + 𝑐𝑑�̇� + 𝑑 = 𝐵𝑀𝑢𝑠 + 𝐵𝐸𝑢𝑐 (12)

    Finally, the overall switched system adheres to the following

    properties:

    Property 1: The inertia matrix (𝑀) is positive-definite and can be bounded as 𝑐𝑚 ≤ 𝑀 ≤ 𝑐𝑀, where 𝑐𝑚 and 𝑐𝑀 are known constants.

    Property 2: The centripetal and Coriolis matrix (𝑉𝑚) can be upper bounded as |𝑉𝑚| ≤ 𝑐𝑣|�̇�|, where 𝑐𝑣 is a known constant.

    Property 3: The gravitational effects matrix (𝐺) can be upper bounded as |𝐺| ≤ 𝑐𝐺, where 𝑐𝐺 is a known constant.

    Property 4: The passive and viscoelastic tissue effects (𝑃) can be bounded as |𝑃| ≤ 𝑐𝑃1 + 𝑐𝑃2|�̇�|, where 𝑐𝑃1 and 𝑐𝑃2 are known constants.

    Property 5: The time-varying disturbance term (𝑑) can be upper bounded as |𝑑| ≤ 𝑐𝑑, where 𝑐𝑑 is a known constant.

    Property 6: The lumped muscle control effectiveness term

    (𝐵𝑀) can be upper and lower bounded as 𝑐𝐵1 ≤ 𝐵𝑀 ≤ 𝑐𝐵2, where 𝑐𝐵1 and 𝑐𝐵2 are known constants.

    Property 7: The motor control effectiveness term (𝐵𝐸) can be upper and lower bounded as 𝑐𝐸1 ≤ 𝐵𝐸 ≤ 𝑐𝐸2, where 𝑐𝐸1 and 𝑐𝐸2 are known constants.

    Property 8: The inertia and centripetal and Coriolis matrices

    follow the skew-symmetry relationship: 1

    2�̇� − 𝑉𝑚 = 0.

    III. CONTROL DEVELOPMENT

    The control objective of the arm-cycle system is to track a

    desired crank cadence. Two tracking errors are measured and

    the error signals are expressed as:

    𝑒1 ≜ 𝑞𝑑 − 𝑞 (13)

    𝑒2 ≜ �̇�1 + 𝛼𝑒1 (14)

    where 𝑒1 is the position tracking error, 𝑒2 is an auxiliary filtered tracking error that considers errors in position and cadence

    tracking, and 𝛼 is a user-defined positive gain. The desired crank position (𝑞𝑑) is designed so that its derivatives exist and are bounded as �̇�𝑑, �̈�𝑑 ∈ ℒ∞ . Additionally, the tracking errors are combined to generate an error vector (𝑧):

    𝑧 ≜ [𝑒1 𝑒2]𝑇 (15)

    The open-loop error system is calculated by taking the time

    derivative of (14), multiplying the derivative by the inertia

    matrix (𝑀), and using (12)-(14) to yield:

    𝑀�̇�2 = 𝜒 − 𝑒1 − 𝑉𝑚𝑒2 − 𝐵𝑀𝑢𝑠 − 𝐵𝐸𝑢𝑐 (16)

    where 𝜒 represents a group of constant and state-dependent terms that can be bounded using Properties 1-8 by a known

    function of the states. Equation 16 and the Lyapunov-based

    stability analysis from the subsequent section are used to design

    the robust sliding mode controllers for the stimulation (𝑢𝑠):

    𝑢𝑠 ≜ 𝑘1𝑒2 + (𝑘2 + 𝑘3‖𝑧‖ + 𝑘4(‖𝑧‖)2)𝑠𝑔𝑛(𝑒2) (17)

    and the current supplied to the motor (𝑢𝑐):

    𝑢𝑐 ≜ 𝑘5𝑒2 + (𝑘6 + 𝑘7‖𝑧‖ + 𝑘8(‖𝑧‖)2)𝑠𝑔𝑛(𝑒2) (18)

    where 𝑘1, 𝑘2, 𝑘3, 𝑘4, 𝑘5, 𝑘6, 𝑘7, and 𝑘8 are selectable positive control gains and sgn( ̇ ∙) is the signum function. The control inputs in (17)-(18) are substituted into the open-loop error

    system in (16) to generate the closed-loop system dynamics:

    𝑀�̇�2 = 𝜒 − 𝑒1 − 𝑉𝑚𝑒2 − 𝐵𝑀[𝑘1𝑒2 + (𝑘2 + 𝑘3‖𝑧‖

    +𝑘4(‖𝑧‖)2)𝑠𝑔𝑛(𝑒2)]- 𝐵𝐸[𝑘5𝑒2 + (𝑘6 + 𝑘7‖𝑧‖ +

    𝑘8(‖𝑧‖)2)𝑠𝑔𝑛(𝑒2)]

    (19)

    IV. STABILITY ANALYSIS

    The stability analysis begins by defining 𝑉𝐿, a positive-definite, continuously differentiable common Lyapunov

    function candidate, denoted as:

    𝑉𝐿 ≜

    1

    2𝑀𝑒2

    2 +1

    2𝑒1

    2 (20)

  • 4

    4

    where 𝑉𝐿 can be bounded as:

    𝜆1(‖𝑧‖)2 ≤ 𝑉𝐿 ≤ 𝜆2(‖𝑧‖)

    2 (21)

    where 𝜆1 and 𝜆2 are known constants. Taking the derivative of (20) and substituting in (14) and (16) results in:

    �̇�𝐿 = 𝑒2(𝜒 − 𝑒1 − 𝑉𝑚𝑒2 − 𝐵𝑀𝑢𝑠 − 𝐵𝐸𝑢𝑐)

    +1

    2�̇�𝑒2

    2 + 𝑒1(𝑒2 − 𝛼𝑒1) (22)

    For the case when 𝜎𝑚 = 1 and 𝜎𝑒 = 0, (22) can be simplified by using Properties 6-8, (17)-(18), cancelling terms, upper

    bounding, and assuming certain gain conditions are met to

    produce:

    �̇�𝐿 ≤ −𝑐𝐵1𝑘1𝑒22 − 𝛼𝑒1

    2 (23)

    For the case when 𝜎𝑚 = 0 and 𝜎𝑒 = 1, (22) can be simplified by using Properties 6 and 8, (17)-(18), cancelling terms, upper

    bounding, and assuming certain gain conditions are met to

    produce:

    �̇�𝐿 ≤ −𝑐𝐸1𝑘5𝑒22 − 𝛼𝑒1

    2 (24)

    The bounds in (23) and (24) can be expressed in terms of the

    error vector (𝑧) and a known constant (𝜆3) for all time as:

    �̇�𝐿 ≤ −𝜆3(‖𝑧‖)2 (25)

    Knowing that 𝑉𝐿 is bounded by (21), then �̇�𝐿 can be expressed as a first order differential equation:

    �̇�𝐿 ≤ −

    𝜆3𝜆2

    𝑉𝐿 (26)

    The bound in (26) was solved for 𝑉𝐿 in terms of the initial condition of 𝑉𝐿 (𝑉𝐿,0), 𝜆2, 𝜆3, and 𝑡 to yield:

    𝑉𝐿 ≤ 𝑉𝐿,0exp (−𝜆3(𝑡−𝑡0)

    𝜆2) (27)

    Using (21) and the initial condition of the error vector (𝑧0), (27) can be rewritten as:

    ‖𝑧‖ ≤ √𝜆2𝜆1

    ‖𝑧0‖exp (−𝜆3(𝑡 − 𝑡0)

    𝜆2) (28)

    Hence, the error system is bounded by an exponentially

    decaying envelope. As a result, the stability analysis shows that

    the closed-loop error system in (19) is globally, exponentially

    stable for all 𝑡 ∈ [𝑡0, ∞), where 𝑡0 is initial time.

    V. EXPERIMENTS

    For simplicity, the performance of the controllers in (17)-(18)

    were assessed using preliminary experiments on the right arm.

    The goal of the experiments was to track a desired crank

    velocity of 40 RPM. In the experiments, only the biceps and

    triceps muscles of the right arm were stimulated. Experiments

    were performed on three able-bodied individuals (three male)

    23-27 years old. Prior to the experiment, each individual gave

    written informed consent approved by the University of Florida

    Institutional Review Board.

    A. Arm-Cycle Testbed Setup

    The arm-cycle system, shown in Fig. 2, consists of two

    independently controlled arm-cycles. Each arm-cycle has its

    own electric motor, shaft, handle, torque sensor, and encoder.

    The motor is a 250 Watt, brushed, 24 VDC electric motor

    (Unite Motor Co. Ltd. MY1016Z) mounted to the system frame

    and coupled to the handle. The torque sensor is located on the

    same shaft as the handle. The encoder (US Digital H1) is used

    to provide crank position feedback. The handle has straps to

    ensure the user’s arm remains fixed to the cycle. There is also

    an emergency stop button that enables the user to stop the

    experiment immediately, if necessary.

    Fig. 2. Right arm-cycle testbed setup. (A) Electric motor. (B) Handle. (C)

    Encoder. (D) Torque Sensor.

    The current supplied to the motor was controlled with data

    acquisition hardware (Quanser QPIDe) and an Advanced

    Motion Controls motor driver. A desktop computer was used to

    run real-time control software (QUARC 2.5,

    MATLAB/Simulink, Windows 10) at a sampling rate of 500

    HZ to implement the controller. A current-controlled stimulator

    (Hasomed RehaStim) and PALS electrodes were used to deliver

    stimulation to the biceps and triceps; the same computer and

    control software were used to control the stimulation.

    B. Experimental Setup

    Each participant was seated in a chair in front of the arm-

    cycle testbed at a distance that allowed the entire crank cycle to

    be completed without fully extending the right arm. Electrodes

    were placed on the individual’s right biceps and triceps, as seen

    in Fig. 3. The top biceps and triceps electrodes were placed high

  • 5

    5

    enough on the arm to ensure there was enough space for the

    bottom electrodes, but low enough that the shoulder would not

    be stimulated. The individuals were instructed to remain

    passive throughout the entire exercise, unless otherwise

    instructed such as in Protocol 1, and to stay fixed in the chair;

    this was done to minimize the volitional effort in the

    experiment.

    Fig. 3. Arm-cycle testbed with an individual. (A) PALS electrodes for Biceps.

    (B) PALS electrodes for Triceps.

    Two different protocols were performed on the arm-cycle

    testbed, identified as Protocol 1 and Protocol 2. In Protocol 1,

    the motor was always on and muscle stimulation was applied in

    the muscle stimulation regions. Individuals were shown a

    monitor with the actual and desired cadence; they were also

    instructed to briefly apply volitional effort if the cadence error

    began to exceed ±5 RPM. Protocol 1 was conducted to evaluate the performance of the modified arm-cycle testbed. In

    Protocol 2, the motor was on in all regions except the bicep

    stimulation region. Protocol 2 was conducted without the motor

    during the bicep stimulation region to explore the possibility of

    only having the motor on during kinematic dead zones.

    Each experiment was designed to last 180 seconds. The

    experiment began by using the motor to bring the arm-cycle to

    40 RPM before starting the switching between muscle

    stimulation and motor. The arm cycle speed was increased

    exponentially to 40 RPM before leveling out. The switching

    signal was designed to start at 20 seconds and from that point

    on the participant received stimulation until the end of the

    experiment when in FES regions of the crank cycle.

    Furthermore, the motor had a current offset of 0.05 A to

    ensure smooth cycling in the muscle stimulation regions, since

    the motor was never completely turned off in Protocol 1. The

    stimulation regions were found experimentally and identified

    as:

    5.06 𝑟𝑎𝑑 < 𝑄𝑏𝑖 < 5.76 𝑟𝑎𝑑 (28)

    1.75 𝑟𝑎𝑑 < 𝑄𝑡𝑟𝑖 < 2.44 𝑟𝑎𝑑 (29)

    where 𝑄𝑏𝑖 and 𝑄𝑡𝑟𝑖 represent the regions where the biceps and triceps were stimulated, respectively. Multiple experiments

    were performed to adjust the gains in (14) and (17)-(18) to

    ensure that the controller tracked the desired cadence

    effectively. Once the controller was tuned, a final experiment

    was performed to collect data.

    C. Results

    Protocol 1

    The cadence, cadence error, motor current input, and pulse

    width stimulation for each subject for Protocol 1 are seen in

    Fig. 4-9.

    Fig. 4. Plots of Subject 1 (S1) results. The top graph compares the actual

    cadence (blue) to the desired cadence (red). The bottom graph shows the

    cadence and position errors, where 𝑒1̇ (blue) is the cadence error in RPM and 𝑒1 (red) is the position error in rad. The plots start at 20 seconds since the rider starts at rest and ramps up to the desired cadence (40 RPM) in 20 seconds, after

    which the controllers in (17)-(18) turn on.

    Fig. 5. Plots of Subject 1 (S1) results. The top graph shows the amount of

    current sent to the motor. The bottom graph shows the pulse width stimulation

    over the course of the experiment. The vertical black line indicates when the controllers in (17)-(18) were activated.

    Fig. 6. Plots of Subject 2 (S2) results. The top graph compares the actual

    cadence (blue) to the desired cadence (red). The bottom graph shows the

    cadence and position errors, where 𝑒1̇ (blue) is the cadence error in RPM and 𝑒1 (red) is the position error in rad. The plots start at 20 seconds since the rider starts at rest and ramps up to the desired cadence (40 RPM) in 20 seconds, after which the controllers in (17)-(18) turn on.

  • 6

    6

    Fig. 7. Plots of Subject 2 (S2) results. The top graph shows the amount of current sent to the motor. The bottom graph shows the pulse width stimulation

    over the course of the experiment. The vertical black line indicates when the

    controllers in (17)-(18) were activated.

    Fig. 8. Plots of Subject 3 (S3) results. The top graph compares the actual

    cadence (blue) to the desired cadence (red). The bottom graph shows the

    cadence and position errors, where 𝑒1̇ (blue) is the cadence error in RPM and 𝑒1 (red) is the position error in rad. The plots start at 20 seconds since the rider starts at rest and ramps up to the desired cadence (40 RPM) in 20 seconds, after

    which the controllers in (17)-(18) turn on.

    Fig. 9. Plots of Subject 3 (S3) results. The top graph shows the amount of

    current sent to the motor. The bottom graph shows the pulse width stimulation

    over the course of the experiment. The vertical black line indicates when the controllers in (17)-(18) were activated.

    Protocol 2

    The cadence, cadence error, motor current input, and

    pulse width stimulation for each subject for Protocol 2 are

    seen in Fig. 10-15.

    Fig. 10. Plots of Subject 1 (S1) results. The top graph compares the actual cadence (blue) to the desired cadence (red). The bottom graph shows the

    cadence and position errors, where 𝑒1̇ (blue) is the cadence error in RPM and 𝑒1 (red) is the position error in rad. The plots start at 20 seconds since the rider starts at rest and ramps up to the desired cadence (40 RPM) in 20 seconds, after

    which the controllers in (17)-(18) turn on.

    Fig. 11. Plots of Subject 1 (S1) results. The top graph shows the amount of current sent to the motor. The bottom graph shows the pulse width stimulation

    over the course of the experiment. The vertical black line indicates when the

    controllers in (17)-(18) were activated.

    Fig. 12. Plots of Subject 2 (S2) results. The top graph compares the actual

    cadence (blue) to the desired cadence (red). The bottom graph shows the

    cadence and position errors, where 𝑒1̇ (blue) is the cadence error in RPM and 𝑒1 (red) is the position error in rad. The plots start at 20 seconds since the rider starts at rest and ramps up to the desired cadence (40 RPM) in 20 seconds, after

    which the controllers in (17)-(18) turn on.

  • 7

    7

    Fig. 13. Plots of Subject 2 (S2) results. The top graph shows the amount of current sent to the motor. The bottom graph shows the pulse width stimulation

    over the course of the experiment. The vertical black line indicates when the

    controllers in (17)-(18) were activated.

    Fig. 14. Plots of Subject 3 (S3) results. The top graph compares the actual cadence (blue) to the desired cadence (red). The bottom graph shows the

    cadence and position errors, where 𝑒1̇ (blue) is the cadence error in RPM and 𝑒1 (red) is the position error in rad. The plots start at 20 seconds since the rider starts at rest and ramps up to the desired cadence (40 RPM) in 20 seconds, after

    which the controllers in (17)-(18) turn on.

    Fig. 15. Plots of Subject 3 (S3) results. The top graph shows the amount of current sent to the motor. The bottom graph shows the pulse width stimulation

    over the course of the experiment. The vertical black line indicates when the

    controllers in (17)-(18) were activated.

    The mean and standard deviation of the cadence error (𝑒1̇) and cadence were calculated for each subject in Protocols 1-2 and

    the results are shown in Table I-II.

    TABLE I

    PROTOCOL 1 ANALYSIS

    Subject 1 Subject 2 Subject 3 Averages

    𝑒1̇ Mean (RPM) -0.04 -0.06 -0.04 -0.04 𝑒1̇ STD (RPM) 2.69 3.43

    2.35 2.83

    Cadence Mean (RPM)

    40.04 40.06 40.04 40.04

    Cadence STD

    (RPM) 2.69 3.43 2.35 2.83

    TABLE II

    PROTOCOL 2 ANALYSIS

    Subject 1 Subject 2 Subject 3 Averages

    𝑒1̇ Mean (RPM) -0.02 -0.03 -0.04 -0.03 𝑒1̇ STD (RPM) 4.74 5.37

    5.94 5.35

    Cadence Mean

    (RPM) 40.02 40.03 40.04 40.03

    Cadence STD

    (RPM) 4.74 5.37 5.94 5.35

    VI. DISCUSSION

    A. Experimental Results

    The experimental results from Protocol 1 demonstrate the

    ability of the controllers in (17)-(18) to distribute current to the

    electric motor and apply FES to the rider’s muscles on the

    modified arm-cycle. The experiment in [6] had the same

    protocol as Protocol 1 except for the cadence tracked. In [6], a

    cadence of 65 RPM was tracked, while the experiments

    performed in this thesis had a cadence of 40 RPM. A cadence

    of 65 RPM was deemed too fast to be an effective arm cycling

    cadence and was reduced to 40 RPM. In addition, participants

    were not allowed to apply volition in [6] and Protocol 2 was

    also not performed in the previous study [6]. The experiment in

    [6] had a cadence error of -0.06 ± 7.96 RPM and the experiments in this thesis had an average cadence error of -0.04

    ± 2.83 RPM. The standard deviations of the cadence error (𝑒1̇) and cadence

    for each subject in Protocol 1 were all under 4 RPM. However,

    the standard deviations approximately double in Protocol 2

    compared to Protocol 1. The average cadence error in Protocol

    2 was -0.03 ± 5.35 RPM, which is also smaller than the error in [6]. Thus, these results display the arm-cycle testbed’s

    potential for future experiments.

    A challenge associated with decoupled cycling, such as the

    arm-cycle in this thesis, is that for a portion of the crank cycle

    the participant must fight against gravity. With a coupled cycle,

    when one side is being resisted by gravity, the other is being

    aided to assist in offsetting the gravitational effect. However,

    this is not the case with a decoupled cycle. This results in a

    portion of the cycle having an increased difficulty, which is

    partially responsible for the worse cadence error when the

    participant provides no voluntary input. To improve this result,

    a more advanced controller will be necessary, such as an

    adaptive controller to learn these periodic dynamics that occur

    throughout the crank cycle.

  • 8

    8

    B. Arm-cycle Testbed Design

    The arm-cycle testbed used for this thesis is similar to the one

    used in [6]; Fig. 16 shows the left arm-cycle testbed used in [6].

    The arm-cycle testbed in this thesis used the same arm-cycle

    frame, handle, shaft couplers, encoder, and torque sensor. The

    direct drive electric motor was replaced with the chain driven

    electric motor mentioned before. The existing direct drive

    motor lacked resolution resulting in an inability to effectively

    control the system using a robust sliding-mode controller.

    Fig. 16 Original left arm-cycle testbed.

    To accommodate the new motor, a motor mount was designed

    and machined so that the motor was mounted to the existing

    arm-cycle testbed structure. A new shaft was machined to

    couple the handle, sprocket driven by the motor, and the gear

    connected to the encoder. The chain used to drive the sprocket

    was fed through a non-spring-loaded chain tensioner to

    maintain tension in the chain.

    Furthermore, a sleeve bearing was mounted near the encoder

    to provide stability at the end of the shaft and to ensure accurate

    measurements were recorded by the encoder. The previous

    sleeve bearing near the handle was replaced with one that only

    had a 5° shaft misalignment capability; the existing bearing had

    a 23° misalignment capability. This sleeve bearing was replaced

    to reduce disturbances seen by the system during cycling.

    C. Future Design Improvements

    The arm-cycle testbed can be improved for future

    experiments by increasing the rigidity of the system. The

    system currently flexes a visible amount along the coupling

    shaft. Flexible couplers were initially chosen to help account

    for misalignment but have proven to be detrimental. The

    flexible couplers can be replaced with rigid couplers. In

    addition, an aluminum shaft was used to couple the sprocket

    and encoder gear which can be replaced with a steel shaft.

    Lastly, the coupling shaft portion of the system can be

    shortened.

    VII. CONCLUSION

    This thesis discussed the challenges and benefits of

    controlling FES systems, particularly for rehabilitation

    applications. The application discussed in this thesis was for an

    arm-cycle system and the thesis sought to improve the

    performance (cadence tracking) of the arm-cycle system in a

    previous study. The changes included implementing new

    hardware (i.e. electric motor) and software (i.e. modified robust

    sliding-mode controller). The results from the preliminary

    experiments indicate the controllers’ ability to sufficiently track

    a desired cadence; the results also showed that the new system

    is an improvement over the previous system.

    The improved arm-cycle testbed will enable future research

    in FES rehabilitation for upper extremities. There is potential

    for restoring muscle function in individuals with NCs using

    arm-cycling; however, experimental results from experiments

    conducted on able-bodied individuals do not represent how

    individuals with NCs would perform. Moreover, additional

    experiments, especially on individuals with NCs, will need to

    be performed to further refine the controller and investigate

    how this system would affect people with NCs. Future research

    will involve characterizing the stimulation regions analytically

    using upper arm kinematics and implementing adaptive control

    schemes to account for unknown disturbances. Finally, future

    experiments will involve simultaneous, independent cycling of

    the left and right arms.

    ACKNOWLEDGMENT

    I would like to acknowledge my advisor, Dr. Warren E.

    Dixon for allowing me to work in his lab the last three years,

    my committee members and colleagues at the Nonlinear

    Controls and Robotics Laboratory. I would also like to

    acknowledge my mentors: Christian Cousin, Courtney Rouse,

    and Brendon Allen for all the help and knowledge they have

    provided me throughout the years. I would not be the engineer

    I am today without everyone’s help and for that I will be

    eternally grateful.

    REFERENCES

    [1] M. J. Bellman, T. -H. Cheng, R. J. Downey, C. J. Hass, and W. E. Dixon, "Switched Control of Cadence During Stationary Cycling Induced by Functional Electrical Stimulation," IEEE Transactions on Neural Systems

    and Rehabilitation Engineering, vol. 24, no. 12, pp. 1373-1383 (2016).

    [2] W. E. Dixon and M. J. Bellman, "Cycling Induced by Functional Electrical Stimulation: A Control Systems Perspective," ASME Dynamic

    Systems & Control Magazine, vol. 4, no. 3, pp. 3-7 (2016).

    [3] C.-W. Peng, S.-C. Chen, C.-H. Lai, C.-J. Chen, C.-C. Chen, J. Mizrahi, and Y. Handa, “Review: Clinical benefits of functional electrical

    stimulation cycling exercise for subjects with central neurological

    impairments,” J. Med. Biol. Eng., vol. 31, pp. 1-11, 2011. [4] K. Hunt, B. Stone, N.-O. Negard, T. Schauer, M. Fraser, A. Cathcart, C.

    Ferrario, S. Ward, and S. Grant, “Control Strategies for Integration of

    Electric Motor Assist and Functional Electrical Stimulation in Paraplegic

    Cycling: Utility for Exercise Testing and Mobile Cycling,” IEEE

    Transactions on Neural Systems and Rehabilitation Engineering, vol. 12,

    no. 1, pp. 89–101, Mar. 2004. [5] M. J. Bellman, R. J. Downey, A. Parikh and W. E. Dixon, "Automatic

    Control of Cycling Induced by Functional Electrical Stimulation with

    Electric Motor Assistance," IEEE Transactions on Automation Science and Engineering, vol. 14, no. 2, pp. 1225-1234 (2017).

    [6] M. H. Cohen, V. H. Duenas, C. A. Cousin, and W. E. Dixon, “Control of the Upper LImbs for Rehabilitative Arm-Cycling via Functional Electrical Stimulation,” Journal of Undergraduate Research, University

    of Florida, vol. 20, no. 1, 2018.

    [7] G. J. Snoek, M. J. Ijzerman, H. J. Hermens, D. Maxwell, and F. Biering-Sorensen, “Survey of the needs of patients with spinal cord injury: impact

    and priority for improvement in hand function in tetraplegics,” Spinal

    Cord, vol. 42, no. 9, pp. 526–532, 2004. [8] A. Carraha, C. Cousin, V. Duenas, C. Rouse, and W. Dixon, “Stationary

    Arm Cycling Induced by Functional Electrical Stimulation – Muscle Zone

    Characterization,” Journal of Undergraduate Research, University of Florida, vol. 18, no. 3, 2017.

    [9] V. H. Duenas, C. A. Cousin, A. Parikh, P. Freeborn, E. J. Fox, and W. E. Dixon, "Motorized and Functional Electrical Stimulation Induced Cycling via Switched Repetitive Learning Control," IEEE Transactions on

    Control Systems Technology, to appear.