Dynamic Digital Human Model for ergonomic assessment based on human-like behaviour and requiring a reduced set of data for a simulation Giovanni De Magistris*„ , Alain Micaelli„ , Jonathan Savin… , Clarisse Gaudez… , Jacques Marsot… „ CEA, LIST, Interactive Simulation Laboratory, 18 route du panorama, BP6, Fontenay aux Roses, F- 92265 France … Institut national de recherche et de s´ ecurit´ e (INRS), rue du Morvan, CS 60027, Vandœuvre-l` es-Nancy, F- 54519 France

Abstract Biomechanical risk factors assessment of a work activity is usually based on the study of a human operator’s postures and forces while performing the work task. Hence, assessing the ergonomics of a future workstation at the design stage requires that an operator performs the work on a prototype or a similar equipment. An alternative solution has emerged through the use of digital human models (DHM) for ergonomics analysis. Yet, using industrial DHM software packages available for ergonomic assessment is usually a complex and time-consuming task. A challenging aim therefore consists in developing an easy-to-use DHM capable of computing dynamic, realistic movements and internal characteristics (position, velocities, accelerations and torques) in quasi-real time, based on a simple description of the future work task, in order to achieve reliable ergonomics assessments of various work task scenarii at an early stage of the design process. We have developed such a dynamic DHM automatically controlled in force and acceleration, inspired by human motor control and based on robotics and physics simulation. In our simulation framework, the DHM motion is dictated by real-world Newtonian physical and mechanical simulation, along with automatic control of applied forces and torques. Our controller handles multiple simultaneous tasks (balance, contacts, manipulation) in real time along with human-like feedforward force and impedance control. An experimental insert-fitting activity has been simulated and assessed based on the OCRA ergonomic index. A comparison with experimental human data showed consistent results: joint torques, DHM movements and their related OCRA assessment were realistic and coherent with human-like behaviour and performance. The main interest of our DHM is that it requires minimal information for a simulation: a starting point, an intermediate point for obstacle avoidance and an end point, along with the applied force for insert clipping. Moreover, changing the subject’s anthropometry and the scenario does not require new trajectory specification nor additional tuning. Keywords: Posture and motion, Motor Behaviour , Design Integrated Prevention

1. Introduction Biomechanical risk factors assessment of a work activity is usually based on the study of a human operator’s postures and forces while performing the work task. A basic analysis of the task can rely on questionnaires, interviews and video analysis, but a more accurate and comprehensive analysis requires collection of exertion (force sensor and/or electromyography) and posture data (e.g. motion capture technique). Such an analysis entails complex and expensive instrumentation that may hamper task performance. Moreover, it is not easily suitable at early design stage, when no prototype or similar equipment exists yet. In recent years, an alternative solution appeared with the use of digital human models (DHM) for ergonomics analysis. Several software packages have been developed to facilitate ergonomic assessment, such as SAMMIE (Porter et al., 2004), JACK (Badler, 1997), Ergoman (Schaub et al., 1997) and SANTOSHuman (Group, 2004)(Vignes, Corresponding author. Email address: giovanni [email protected]

2004). Nevertheless, using industrial DHM software packages available for ergonomic assessment is usually a complex and time-consuming task. For instance, simulation may be built up like a cartoon through interactive positioning of the DHM with mouse, menus and keyboard, which requires expert skills in ergonomics and human motion in order to avoid coarse or unrealistic behaviours. One can also use a tracking system or motion capture to record realistic movements but this requires extensive instrumentation, a full scale mock-up of the future workstation, or a similar one, and tricky motion capture data processing. Simulations can eventually be based on pre-defined human motion libraries (reach, grasp...), but they usually look quite unnatural. Yet, besides postures, ergonomics assessments also need at least an estimate of exertions, which may not be reliable (Savin, 2011). For these reasons, using industrial DHM software for biomechanical risk factors assessment based on simulations of industrial or experimental

work task situations may lead to significantly misleading stress estimation (Lamkull et al., 2009). A challenging aim therefore consists in developing an easy-to-use DHM capable of computing dynamic, realistic movements and internal characteristics (position, velocities, accelerations and torques) in quasi-real time, based on a simple description of the future work task, in order to achieve reliable ergonomics assessments of various work task scenarii at an early stage of the design process. We have developed such a dynamic DHM automatically controlled in force and acceleration (DeMagistris et al., 2011), inspired by human motor control (Todorov and Jordan, 1998) and based on robotics and physics simulation. In our simulation framework, the DHM motion is dictated by real-world Newtonian physical and mechanical simulation, along with automatic control of applied forces and torques. Our controller handles multiple simultaneous tasks (balance, contacts, manipulation) in real time along with human-like feedforward force and impedance control (Burdet et al., 2001). An experimental insert-fitting activity has been simulated and assessed based on the OCRA ergonomic index. A comparison with experimental human data showed consistent results. The main interest of our DHM is that it requires minimal information for a simulation: a starting point, an intermediate point for obstacle avoidance and an end point, along with the applied force for insert clipping. Moreover, changing the subject’s anthropometry and the scenario does not require new trajectory specification nor additional tuning. Joint torques, DHM movements and trajectory and their related OCRA assessment are realistic and consistent with human-like behaviour and performance. The first part of this paper describes the industrial insert-fitting activity used as an experimental and validation frame. The second part outlines the principles of our dynamic DHM. The third part describes the trajectory and velocity analysis method to compare real human data and simulations. The fourth part describes results (trajectory and velocity analysis, ergonomic assessment and torque analysis). Finally, we discuss the issues raised by the approach and the conclusions.

Figure 1: Simulation and validation processes

Institut national de recherche et de s´ecurit´e (INRS), Laboratory for Biomechanics and Ergonomics (Gaudez, 2008). This team subsequently adapted it to laboratory experimentation in accordance with biomedical research requirements. Eleven healthy right-handed subjects (nine males and two females) took part in the study [age = 29.4 ± 9.2 yrs (mean ± standard deviation), height = 177.7 ± 10.3 cm, body mass = 75.9 ± 9.3 kg]. The subjects gave their informed consent to the experiments, which were approved by the institutional ethics committee, and completed a health questionnaire. The subjects’ anthropometry was measured to build human subjects models based on the Hanavan model (Hanavan, 1964) (size, weight and 41 body segment measurements). The workstation comprised a force platform assembled on a lift table fitted with a row of ten insert supports arranged at 45° from back to front in the sagittal plane. Subjects were asked to perform the experimental task ac-

2. Material and Methods Fig. 1 illustrates simulation and validation processes. In this section, we explain the principles of both the simulation process and our model. 2.1. Experimental protocol Our experiment deals with an industrial assembly task consisting in clipping small metal parts to the plastic instrument panel of a vehicle prior to screwing components to it. This activity is common in the automotive industry and presents risk of upper limb work-related musculoskeletal disorders. Our activity was first studied by France’s

Figure 2: Human subject experiment

2

cording to the two methods used in the workshop: either using only fingers or using a hand-held tool meeting specific ergonomic criteria (NST-n168, 1998). Each time, they fitted ten inserts into the ten parallel supports at constant pace (one insert every 4 seconds). The two methods were performed in random order. The height of the table was adapted to the subject’s anthropometry. It was set to 90% of elbow-ground height, in accordance with European standards for a standing work activity requiring normal vision and precision (CEN, 2008). To analyze the task and make an ergonomic assessment, the following instruments have been used:

3. biceps: subjects were seated with the elbow flexed at 90°and arms horizontal, the hand in a neutral position (midway between pronation and supination - thumb facing the rear). Having a strap blocking their wrist, subjects performed an elbow flexion as strong as possible. 4. triceps: subjects were in the same posture as for the biceps test and performed an elbow extension as strong as possible. 5. deltoid : subjects were seated with the arm forward flexed to 45°and the elbow flexed to 90°. A strap blocking their elbow, subjects elevated the arm as strong as possible. 6. trapezius: subjects were standing upright with arms horizontal at shoulder height, elbow blocked by a strap, subject elevated arm as strong as possible.

ˆ A ten cameras MotionAnalysis system to record the

whole body positions and postures ˆ A single top view camera, synchronized with the mo-

tion capture system, to record the subject’s activity so as to be able to subsequently correct optical markers losses or cross-over.

2.1.1. Experimental Analysis The insertion of the first and last two inserts of the row were recorded but not processed: only the insertion of the six central clips has been studied. Motion capture data were post-processed to retrieve the subject anatomical joint angles (shoulder flexion/extension, abduction/adduction and internal/external rotation, elbow pronation/supination, etc.) and to make ergonomic assessment of the experimental task. Motion capture markers positions were first filtered (25 Hz 4th order Butterworth low pass) then processed to calculate joint angles using the Calcium solver module (MarquardtLevenberg optimization algorithm) of the Cortex suite (3.0.0 software version). The EMG signal was filtered and then sampled at a frequency of 1024 Hz; it was then quantified by RMS (Root Mean Square) values calculated using constant consecutive periods of 500 ms. Because the activity is regular and repetitive, the mean RMS for each insertion sequence (RM Sm ) was calculated. The mean value was then normalized using the MVC value. It is expressed as a percentage: RMSm (1) RMSnorm = RMSmvc

ˆ A force platform attached to the lift table (AMTI,

model BP600900-1000) to measure external forces and moments exerted during the activity. ˆ An electromyography system, to record flexor and ex-

tensor muscles activities of the fingers and wrist, biceps, triceps, deltoid and trapezius. Once the skin had been prepared, the Blue Sensor electrodes were placed on the muscle. The detection mode was bipolar, the distance between electrodes was 20 mm and the maximum inter-electrode resistance was 5 kΩ. Data from the MotionAnalysis system and the force platform were recorded simultaneously at a 100 Hz sampling rate. To assess the force exerted during the experiment, maximum isometric contraction (MVC) tests were performed before the experiment. They helped to normalize the amount of electro-muscular activity recorded, thereby enabling the comparison of exertion between subjects. To find the maximum isometric contraction, subjects had to maintain their maximum effort for a few seconds. The operation was performed twice with a 2-minute rest period between each test. The retained MVC reference value was the mean of the two average values recorded during the two tests (Matthiassen et al., 1995). Tests for the different muscles were:

2.1.2. Experimental observations Subjects implemented various strategies when using only their fingers: 1. Four subjects picked up the ten inserts one by one and clipped them onto the support using only the right hand 2. Four subjects picked up the ten inserts one by one from the table with the left hand, then transferred them to the right hand, which was then used to clip them onto the support 3. Two subject picked up the ten inserts all at once from the table with the left hand, then picked up the inserts from the left hand with the right hand, which was then used to clip them onto the support

1. fingers and wrist flexors: subjects were seated with the elbow flexed at 90°, hand supinated (palm up). Having a strap blocking their palm, subjects performed a wrist and fingers flexion as strong as possible. 2. fingers and wrist extensors: subjects were seated with the elbow flexed at 90°, hand pronated (palm down). Having a strap passing over the back of their hand, subjects performed a wrist and fingers extension as strong as possible. 3

[Vroot q˙1 · · · q˙ndof ]t are respectively the acceleration and velocity vectors in generalized coordinates. Gr is the gravity force, τ = [τ1 · · · τndof ]t is the joint torque vector, L = [0(ndof,6) Indof ]t is the matrix to select the actuated degrees of freedom, W = [Γ F ]t denotes all the external wrenches (Fig. 4) where Γ is the moment and F is the force. In Eq. (2), the superscript r stands for “real” wrench values in a simulation. The subscripts c stands for non-sliding contacts at known fixed locations such as the contacts between the feet and the ground, the subscripts e stands for unknown contacts with environment.

4. One subject picked up the inserts one by one from the table with the right hand, transferred them to the left to position them properly, then transferred them back to the right, which was then used to clip them onto the support. In this article, we analyze only the first two strategies. When using the hand-held tool, all subjects implemented the same strategy: they all picked up the ten inserts one by one with the left hand and placed them with this hand on the top of the tool, which was held in the right hand. They then pushed the inserts in place using only the tool in their right hand. Using the hand-held tool, the average applied force during insertion was F x = −15N , F y = −50N , F z = 40N . For the two fingered insertion variants, this force was F x = −25N , F y = −60N , F z = 55N . These values have been used as input parameters of our simulations.

2.2.2. Human-like Dynamic DHM control To control DHM’s movement and postures, we implement a multi-objective DHM controller. Common control techniques are based on pure stiffness compensation of internal and external disturbances. In this article we used a controller that combines both feedback and feedforward techniques (DeMagistris et al., 2013). When a multibody system get in touch with an object, it is important to make the limb more compliant to avoid “contact instability” (Hogan, 1990). An important conclusion, which consistently emerges from these theoretical analyses, is that mechanics needs a feedforward control to be regulated. A number of studies have shown that the nervous system uses internal representations to anticipate the consequences of dynamic interaction forces. In particular, Lackner and Dizio (Lackner and Dizio, 1994) demonstrated that the central nervous system (CNS) is able to predict the centripetal and Coriolis forces; Gribble and Ostry (Gribble and Ostry, 1999) demonstrated the compensation of interaction torques during multijoint limb movement. These studies suggest that the nervous system has sophisticated anticipatory capabilities. We therefore need an accurate internal representation or an inverse model of controlled body dynamics and environment. Based on the notion underlying the acceleration-based control method (Abe et al., 2007)(Colette et al., 2008) and the JacobianTranspose (JT) control method (Pratt et al., 1996)(Liu et al., 2011)(DeMagistris et al., 2011), we developed a combined anticipatory feedforward and feedback control systems. This controller is formulated as two successive Quadratic Programming (QP) problems involving multiple degrees of freedom for simultaneously solving all the constraint equations. The first problem is feedforward control and second is feedback. This computational optimization framework is detailed in (DeMagistris et al., 2013). To simulate the task described in Section 2.1, several objectives have been identified and prioritized:

2.2. DHM simulation According to the overall simulation development process detailed in Fig. 1, we modeled the different components of our DHM. 2.2.1. Human Model Body and Dynamics In our study, the human body was kinematically modelled as a set of articulated rigid bodies branches (Fig. 3), organized into a redundant tree structure, which is characterized by its degrees of freedom (DoF). Depending on the function of the corresponding human segments, each articulation of the tree can be modelled into a number of revolute joints. Our DHM therefore comprises 39 joint DoF and 6 root DoF, with 8 DoF for each leg and 7 for each arm. The root is not controlled. Our digital human models were dimensioned based on subjects’ anthropometry (Hanavan, 1964).

Figure 3: DHM with skinning and collision geometry (left). Right hand model with skinning and collision geometry (right)

The dynamics of the robot is described as a second order system as: X X M T˙ + N T + Gr = Lτ + JcTj Wcrj + JeTk Werk (2) j

1. Centre of Mass (com). The dynamic controller maintains the DHM balance by imposing that the horizontal plane projection of the centre of mass (com) lies within a convex support region (Bretl and Lall, 2008). 2. Thorax. During the experimental task, we observed that thorax orientation varied very little. We there-

k

in which M is the generalized inertia matrix, N T is the centripetal and Coriolis forces, T˙ and T = 4

3.

4.

5.

6.

7.

fore set the desired thorax orientation equal to its initial orientation. Posture. To obtain more realistic movements and to help the controller to avoid any singularity, we specified one DHM reference joint position for the whole simulation. End effectors (EE). This objective deals with the hand movements required to perform the specific manipulation task. Head. When studying the real work task, we noticed that the head follows the movement of the end effector performing the predominant task. This is the head objective. Contact force. This objective is not a target tracking objective, but it is used to minimize contact force and, as its desired value is unknown a priori, we set the desired contact force to zero. Gravity compensation. This objective is supposed to make target tracking control independent of gravity compensation.

More complex trajectories can be divided into overlapping basic trajectory similar to the reaching movements. Such spatio-temporal invariant features of normal movements can be explained by a variety of criteria of maximum smoothness, such as the minimum jerk criterion (Flash and Hogan, 1985) or the minimum torque-change criterion (Uno et al., 1989). We implemented a modified minimum jerk criterion with via-points to perform trajectories. This trajectory is developed in (DeMagistris et al., 2013) and inspired by (Todorov and Jordan, 1998). Thanks to minimum jerk, in order to calculate the trajectory for the rotations and the translations, we only need to provide the starting, intermediates and end 6D points X, the starting and end velocities V and the starting and end accelerations A (the intermediate times TP are found using a nonlinear simplex method to minimize the optimal jerk over all possible passage times). 2.2.4. Contacts Simulations were based on the XDE physics simulation module developed at the CEA-LIST. This module manages the whole physics simulation in real time, including accurate and robust contact detection. Advanced friction effects were modelled in compliance with Coulomb’s friction law, which can be formulated as:

2.2.3. Human-like Movement A movement of our DHM is characterized by the initial and final points of the trajectory (position and orientation), potential obstacle positions (via-points of the trajectory because our controller does not manage collision avoidance for now) and duration. Experimental study of human movements has shown that voluntary movements obey the following three major psychophysical principles:

kfxy k ≤ µ kfz k

with kfxy k the tangential contact force, µ the dry friction factor and kfz k the normal contact force. 2.2.5. Hands The hand model, illustrated in Fig. 3, has 20 DoF. We use a simple Proportional-Derivative controller to control joint position θ where a set of desired position θd corresponds to open/close hand and different preset grasps.

ˆ Hick-Hyman’s law: the average reaction time TRave

required to choose among n probable choices depends on the logarithm of them (Hyman, 1953): TRave = d log2 (n + 1)

(3)

ˆ Fitts’ law: the movement time depends on the log-

2.2.6. Digital mock-up The DMU (Digital mock-up) scenario (Fig. 4) reproduces the experimental environment by ensuring geometric similarity. The inputs used to build the DMU scenario are the workplace spatial organization (x, y and z dimensions), inserts and tool descriptions (x, y, z positions and weight) and the DHM position.

arithm of the relative accuracy (the ratio between movement amplitude and target dimension) (Fitts, 1954): D = g + z log2 (2ΥP ) (4) where D is the duration time, Υ is the amplitude, P is the accuracy, g and z are empirically determined constants.

2.2.7. Task Models It should firstly be recalled that work tasks are usually broken up into sequences of a few elementary motions or postures. For this reason, we modelled the clipping tasks with Finite State Machines (FSM) in Fig. 5 as in the basic principles underlying many design methods (for instance Method Time Measurement (Maynard et al., 1948)) or ergonomic assessment index (OCcupational Repetitive Actions (OCRA) (Occhipinti, 1998)). The different states of one-hand-task in Fig. 5 are:

ˆ Kinematics invariance: hand movements have a

bell-shaped speed profile in straight reaching movements (Morasso, 1981). For more complex trajectories (i.e. handwriting) a 2/3 power law predicts a correlation between speed and curvature (Morasso and Mussa-Ivaldi, 1982) described as: 2

s(t) ˙ = Zs R1− 3

(6)

(5)

where s(t) ˙ is the tangential velocity, R is the radius of curvature and Zs is a proportionality constant, also termed ”velocity gain factor”.

1. Idle: at the start of the simulation, the DHM body is upright and its arms are along the body. 5

State REACH

POSITION

GRASP PUSH Figure 4: DMU scenario with the desired wrenches: com (center of mass) for balance, head following end effector (EE) movement, thorax avoiding large movement, c (contacts) for no sliding contacts, lhand (left hand) and rhand (right hand) are end effector tasks when performing handling action

RELEASE

Input - duration time - in presence of obstacle, intermediate 6D point (via point to obstacle avoidance) - final 6D point - duration time - in presence of obstacle, intermediate 6D point (via point to obstacle avoidance) - final 6D point - duration time - hand fingers position - duration time - push force - duration time - hand fingers position

Table 1: Inputs for the different states of FSM

3. Trajectory Analysis To compare trajectories measured from real human data X h and obtained from simulations X s , we analyze the trajectory of ”position” step of the finite state machine for four subjects. We chose to compare the simulated trajectory to the six medial clip insertion trajectories - insertions of the first and last two clips were ignored. A set of elementary affine transformations was to be used to facilitate comparison between simulated and real trajectories, because they have different duration times, start and final points. An affine transformation (also called an affinity) is any transformation that preserves distance ratios (e.g., the midpoint of a curve remains the midpoint after transformation),angles and colinearity (all points lying on a single line). In general, an affine transformation is a composition of rotations, translations and homothetic transformations. We designed affine transformations so that the recorded and simulated start point (respectively end point) match together.

Figure 5: Task models for the clipping task. One-hand-task at the top and two-hands-task at the bottom

2. Reach: the right hand adopts a grasping position and the head follows the right hand movement. 3. Grasp: the right hand closes the finger and picks up the insert. 4. Position: the right hand moves to preset insertion point and the head follows the right hand movement. 5. Push: the DHM pushes the insert into the appropriate support with the same force as that measured on the force connected to the table. 6. Release: the right hand fingers open. 7. Idle: the DHM returns to its initial position.

3.1. Spatial Transformations 1. Translation. The first step of the needed affine transformation consists in matching recorded and simulated start point. Trajectories are translated to the zero point from the start position by using a simple transformation:

In the same way, we can define states of the two-handtask and the tool-task. When the right and left upper limbs are working, we have two simultaneous tasks (Fig. 5). When the task is performed with the hand-held tool, the FSM is the same as the one for the task performed with two hands, but the RELEASE LHAND GRASP RHAND state is replaced by the RELEASE LHAND POSITION RHAND state. Table 1 shows the different inputs for the different states.

  h  x (t) − xh (0) xs (t) − xs (0) X s =  y s (t) − y s (0)  , X h =  y h (t) − y h (0)  z s (t) − z s (0) z h (t) − z h (0) 

(7)

with t ∈ [0, ..., tf ]. 2. Rotation between vectors. The second step of the needed transformation consists in a rotation. Hence, we define two spatial vectors that have the same origin (zero point) and whose end point is the last point of 6

the recorded and the simulated trajectories. Then we need to rotate the one of these two vectors to match with other. To find the relevant rotation, we calculate the cross product and the angle between the two normalized vectors X s (tf ) = [ xsf yfs zfs ] and X h (tf ) = [ xhf yfh zfh ]. We define the cross product by the determinant of a formal matrix: i j k (8) Xs (tf ) × Xh (tf ) = xsf yfs zfs xh y h z h f

f

3.2. Distances between trajectories Various definitions exist in the literature to characterize the distance between two trajectories defined as non-empty sets. One common definition for the distance between two non-empty subsets of a given set is the Hausdorff distance. This measures the ”closeness” of two sets of points that are subsets of a metric space. Such a measure may be used to assign a scalar score to the similarity between two trajectories (Chen et al., 2011), data clouds or any sets of points. This distance is defined as:

f

dH (X h , X s ) = max{ sup inf s d(x, y),

Using Sarrus’ rule, it expands to s

h

X (tf ) × X (tf )

=

x∈X h y∈X

i(yfs zfh − zfs yfh ) + j(zfs xhf −xsf zfh ) + k(xsf yfh − yfs xhf )

sup inf d(x, y)}

y∈X s x∈X h

(9)

= iux + juy + kuz

where sup represents the supremum and inf the infimum. To give a distance between two trajectories, we measure the Hausdorff distance and the average distance defined as: inf d(x, y)} (14) dave (X h , X s ) = mean{

where ux = (yfs zfh − uz = (xsf yfh − yfs xhf )

zfs yfh ), uy = (zfs xhf − xsf zfh ) and are the components of the unit vector u = (ux , uy , uz ). The angle between vectors is calculated as: θ = arccos(Xs (tf ) · Xh (tf ))

x∈X h ,y∈X s

(10)

where · is the scalar product. 3. Homothetic transformation. Now that the simulated and recorded end points are aligned in direction, the third step consists in matching their position. Hence, we do a homothetic transformation between the one trajectory with the origin O as homothetic center and the scale factor λ equal to: λ=

kX s (tf )k kX h (tf )k

(13)

4. Velocities Analysis To compare motions executed with different velocities it is interesting to make speed independent of the time scale. To obtain this, we consider the dimensionless normalized time:

(11)

t=

4. Rotation to obtain coplanarity of three keypoints. We calculate the midpoint C = (Cx , Cy , Cz ) −−−→ of the segment OX h identified by origin point and the final point of the trajectory. To obtain the coplanarity

t tf

(15)

where tf is the duration of the rigid body motion and is referred to as the time scale. Multiplying V (t) by tf and substituting t by t results in invariants that are independent of the time scale (Schutter, 2008). 5. Results 5.1. Ergonomic Assessment The experimental insert-fitting activity and DHM simulations were assessed based on OCRA ergonomic index (Occhipinti, 1998)(EU, 2006). In Tables 2 and 3 we compare these assessments. Detailed results are reported in (DeMagistris et al., 2013).

Figure 6: Rotation to obtain coplanarity of three key-points

of three curve points (initial points, final points and one point for each curve), we make a rotation around −−−→ the segment OX h (see Fig. 6). Therefore, we find the two points A on the trajectory X h and B on the trajectory X s intersecting the plane perpendicular to −−−→ the segment OX h and we calculate the rotation angle γ as: −→ −−→ ! AC · BC γ = arccos (12) −→ −−→ ||AC||||BC||

OCRA Index Human DHM LUL RUL LUL RUL One-hand 7.25 ± 3.70 7.19 ± 3.39 Two-hands 1.21 ± 0.34 7.46 ± 2.62 1.15 ± 0.32 6.93 ± 2.78 Tool 2.49 ± 0.61 3.30 ± 0.67 2.40 ± 0.57 3.23 ± 0.61

Table 2: OCRA Index Values. Mean ± standard deviation. LUL and RUL stand for Left and Right Upper Limb respectively.

7

Risk Human DHM LUL RUL LUL RUL One-hand Risk Risk Two-hands No Risk Risk No Risk Risk Tool Very Low Very Low Very Low Very Low

Table 3: Risk level. LUL and RUL stand for Left and Right Upper Limb respectively.

5.2. Trajectory In Table 4, we show trajectory distances between the simulated trajectory and the human trajectory of “position” step of the finite state machine (see Fig. 5) for all insertions (six central insertions) during one-hand-task. We compare the simulated trajectory and the human trajectory for the four subjects using only the right hand (see Section 2.1.2). Subject 1 2 3 4

Hausdorff (1.1 ± 0.3) (1.5 ± 0.4) (1.7 ± 0.5) (2.6 ± 1.2)

distance cm cm cm cm

Figure 8: Right wrist trajectories for the second subject

Average distance (0.5 ± 0.1) cm (0.7 ± 0.2) cm (0.9 ± 0.3) cm (1.5 ± 0.6) cm

Table 4: Right Wrist Trajectories - Distances between trajectories for all insertions

Figure 9: Right wrist trajectories for the third subject

Figure 7: Right wrist trajectories for the first subject

5.3. Velocities Figure 10: Right wrist trajectories for the fourth subject

In Figs. 11, 12, 13 and 14 we compare velocities obtained by real human data and DHM simulations. In particular, we analyze human velocities of ”position” step of the finite state machine (see Fig. 5) for four subjects during one-hand-task. We compare the simulated trajectory and the human trajectory for the four subjects using only the right hand (see Section 2.1.2).

5.4. Torque Analysis It is noticeable that our DHM obtains torques that are compatible with human performance. For example, the maximum simulated value of the right elbow flexion torque 8

is about 22 N ·m (respectively 5 N ·m for the wrist torque, see Fig. 15). These values are always smaller than maximum admissible torque at the elbow and wrist joints (this maximum torque is approximately 70 N · m for men and 35 N · m for women (Askew et al., 1981) at the elbow and approximately 8.05 N · m in flexion and 6.53 N · m in extension (Ciriello et al., 2001) at the wrist). 6. Discussion The ergonomic assessment based on real human data and simulations are consistent (see Tables 2 and 3). Our DHM computes torques compatible with human performance. This is particularly noticeable since common DHM software may compute joint torques and/or working posture that are sometimes not coherent with human performances, thus leading to erroneous ergonomic assessments (Lamkull et al., 2009)(Savin, 2011). Although we obtain similar trajectories and speed profiles, the difference is more noticeable for small subjects (Figs. 11 and 14). This could be due to the fact that the horizontal distance (distance between the subject and the raw insert sockets, or range between the extreme sockets) are not adapted according to subjects’ size (on the contrary, the height of the table was set to 90% of elbowground height, in accordance with European standards for a standing work activity requiring normal vision and precision (CEN, 2008)).

Figure 11: Right wrist velocities for the first subject

Figure 12: Right wrist velocities for the secondt subject

7. Conclusion In this paper, we introduced a dynamic digital human model controlled in relation to force and acceleration. We used this model to simulate an experimental insert clipping activity in quasi-real-time and applied the simulated postures, time and exertions to an OCRA index-based ergonomic assessment. Given only scant information on the scenario (typically initial and final operator-positions and clipping force), the simulated ergonomic evaluations were in the same risk area as human data. In addition, DHM trajectories are similar to real trajectories.

Figure 13: Right wrist velocities for the third subject

(a) Right elbow torque

(b) Right wrist torque

Figure 14: Right wrist velocities for the fourth subject Figure 15: Simulated torques are compatible with human performance

9

These encouraging results show that a DHM controller may soon overcome the limits of currently common DHM software.

Hyman, R., 1953. Stimulus information as a determinant of reaction time. Journal of Experimental Psychology 45, 188–196. Lackner, J. R., Dizio, P., 1994. Rapid adaptation to coriolis force perturbations of arm trajectory. Journal of Neurophysiology 72, 299–313. Lamkull, D., Hanson, L., Ortengren, R., 2009. A comparative study of digital human modelling simulation results and their outcomes in reality: A case study within manual assembly of automobiles. International Journal of Industrial Ergonomics 39, 428–441. Liu, M., Micaelli, A., Evrard, P., Escande, A., Andriot, C., 2011. Interactive dynamics and balance of a virtual character during manipulation tasks. In: IEEE International Conference on Robotics and Automation. Shanghai, China, pp. 1676–1682. Matthiassen, S. E., Winkel, J., Hgg, G. M., 1995. Normalization of surface emg amplitude from the upper trapezius muscle in ergonomic studies. a review. Journal of Electromyography and Kinesiology 5, 197–226. Maynard, H. B., Stegemerten, G. J., Schwab, J. L., 1948. Methodstime measurement. McGraw Hill book company. Morasso, P., 1981. Spatial control of arm movements. Experimental Brain Research 42, 223–227. Morasso, P., Mussa-Ivaldi, F. A., 1982. Trajectory formation and handwriting: a computational model. Biological Cybernetics 45, 131–142. NST-n168, 1998. Ergonomie des outils ` a main - Probl´ ematique et ´ etat de l’art. Note scientifique et technique de l’INRS. Occhipinti, E., 1998. Ocra, a concise index for the assessment of exposure to repetitive movements of the upper limbs. Ergonomics 41, 1290–1311. Porter, J. M., Case, K., Marshall, R., Gyi, D., Sims, R., 2004. Beyond jack and jill: designing for individuals using hadrian. International Journal of Industrial Ergonomics 33, 249–264. Pratt, J., Torres, A., Dilworth, P., Pratt, G., 1996. Virtual actuator control. In: IEEE International Conference on Intelligent Robots and Systems. pp. 1219–1226. Savin, J., August 2011. Digital human manikins for work-task ergonomic assessment: which degree of confidence and using constraints? Proceedings of the Institution of Mechanical Engineers, Part B: Journal of Engineering Manufacture 225, 1401–1409. Schaub, K., Landau, K., Menges, R., Grossmann, K., 1997. A computer-aided tool for ergonomic workplace design and preventive health care. Human Factors and Ergonomics in Manufacturing 7, 269–304. Schutter, J. D., 2008. Invariant description of rigid body motion trajectories. ASME Journal of Mechanisms and Robotics. Todorov, E., Jordan, M. I., 1998. Smoothness maximization along a predefined path accurately predicts the speed profiles of complex arm movements. Journal of Neurophysiology 80, 697–714. Uno, Y., Kawato, M., Suzuki, R., 1989. Formation and control of optimal trajectory in human multijoint arm movement: Minimum torque-change model. Biological Cybernetics 61, 89–101. Vignes, R. M., 2004. Modeling Muscle Fatigue in Digital Humans. Center for Computer-Aided Design, The University of IOWA. Tech. rep.

References Abe, Y., da Silva, M., Popovi´ c, J., 2007. Multiobjective control with frictional contacts. In: Proc. ACM SIGGRAPH/EG Symposium on Computer Animation. Aire-la-Ville, Switzerland, pp. 249–258. Askew, L. J., An, K. N., Morrey, B. F., Chao, E. Y., 1981. Functional evaluation of the elbow. normal motion requirements and strength determinations. Orthop. Trans., 5–304. Badler, N. I., 1997. Virtual humans for animation, ergonomics, and simulation. In: Proceedings of the IEEE workshop on non-rigid and articulates motion. pp. 28–36. Bretl, T., Lall, S., 2008. Testing static equilibrium for legged robots. IEEE Transactions on Robotics 24, 794–807. Burdet, E., Osu, R., Franklin, D. W., Milner, T. E., Kawato, M., 2001. The central nervous system stabilizes unstable dynamics by learning optimal impedance. Nature 414, 446–449. CEN, 2008. Safety of machinery - anthropometric requirements for the design of workstations at machinery. European standard EN ISO 14738:2008. Chen, J., Wang, R., Liu, L., Song, J., 2011. Clustering of trajectories based on hausdorff distance. In: International Conference on Electronics, Communications and Control (ICECC). pp. 1940–1944. Ciriello, V. M., Snook, S. H., Webster, B. S., Dempsey, P., 2001. Psychophysical study of six hand movements. Ergonomics 44, 922– 936. Colette, C., Micaelli, A., Andriot, C., Lemerle, P., 2008. Robust balance optimization control of humanoid robots with multiple non coplanar grasps and frictional contacts. In: Proceedings of the IEEE International Conference on Robotics and Automation. Pasadena, USA, pp. 3187–3193. DeMagistris, G., Micaelli, A., Andriot, C., Savin, J., Marsot, J., 2011. Dynamic virtual manikin control design for the assessment of the workstation ergonomy. In: First International Symposium on Digital Human Modeling. Lyon. DeMagistris, G., Micaelli, A., Evrard, P., Andriot, C., Savin, J., Gaudez, C., Marsot, J., 2013. Dynamic control of dhm for ergonomic assessments. International Journal of Industrial Ergonomics 43, 170–180. EU, 2006. Eu directive on machinery - directive 2006/42/ec of the european parliament and of the council of 17 may 2006 on machinery. Official Journal of the European Union. URL http://eur-lex.europa.eu/LexUriServ/LexUriServ.do? uri=OJ:L:2006:157:0024:0086:EN:PDF Fitts, P. M., 1954. The information capacity of the human motor system in controlling the amplitude of movement. Journal of Experimental Pshychology 47, 381–391. Flash, T., Hogan, N., 1985. The coordination of arm movements: an experimentally confirmed mathematical model. Journal of Neuroscience 7, 1688–1703. Gaudez, C., 2008. Upper limb musculo-skeletal disorders and insert fitting activity in automobile sector: impact on muscular stresses of fitting method and insert position on part. In: Computer Methods in Biomechanics and Biomedical Engineering. Vol. 11. pp. 101–102. Gribble, P. L., Ostry, D. J., 1999. Compensation for interaction torques during single and multijoint limb movement. Journal of Neurophysiology 82, 2310–2326. Group, V. R., 2004. Technical report for project virtual soldier research. Center for Computer-Aided Design, The University of IOWA. Tech. rep. Hanavan, E. P., 1964. A mathematical model of the human body. Wright-Patterson Air Force Base Report No. AMRL-TR-102, 64– 102. Hogan, N., 1990. Mechanical impedance of single- and multi-articular systems. j. m. winters & s.l. woo. springer-verlag. Multiple muscle systems: Biomechanics and movement organization.

10

Dynamic Digital Human Model for ergonomic ...

Virtual actuator control. In: IEEE International Conference on Intelligent Robots and Systems. pp. 1219–1226. Savin, J., August 2011. Digital human manikins for ...

1MB Sizes 1 Downloads 227 Views

Recommend Documents

Dynamic workflow model fragmentation for distributed execution
... technology for the coordination of various business processes, such as loan ... Workflow model is the basis for workflow execution. In ...... One way to deal with ...

GlobalMobileSearchShare - DYNAMIC DIGITAL CONSULTING
Mobile usage is rapidly taking off in every corner of the globe as a key way consumers stay connected and informed. ISRAEL. UNITED KINGDOM. 37%. 46%.

Dynamic Model Selection for Hierarchical Deep ... - Research at Google
Figure 2: An illustration of the equivalence between single layers ... assignments as Bernoulli random variables and draw a dif- ..... lowed by 50% Dropout.

A dynamic causal model for evoked and induced ...
Jul 30, 2011 - changes in spectral activity induced by other sources and exogenous. (e.g., stimulus) ..... of model alternatives, all one can do is “motivate model space carefully” .... The free-energy principle: a unified brain theory? Nat. Rev.

A Three-dimensional Dynamic Posture Prediction Model for ...
A three-dimensional dynamic posture prediction model for simulating in-vehicle seated reaching movements is presented. The model employs a four-segment ...

Appendices - A Novel Dynamic Pricing Model for the ...
Appendices - A Novel Dynamic Pricing Model for the Telecommunications Industry.pdf. Appendices - A Novel Dynamic Pricing Model for the ...

A Self-Similar Traffic Prediction Model for Dynamic ...
known about the traffic characteristics of wireless networks. It was shown in [1] that wireless traffic traces do indeed exhibit a certain degree of self-similarity and ...

A dynamic stochastic general equilibrium model for a small open ...
the current account balance and the real exchange rate. ... a number of real frictions, such as habit formation in consumption, investment adjustment costs ...... also define the following equations: Real imports. (. ) m t t t t m Q c im. = +. (A30).

Dynamic Sensorimotor Model for open-ended ...
sensors and motors : dS = M. - Uses motor and sensor simulations, for motor control and creation of ... Tool Use : from low to high level skills. A - Context :.

A Self-Similar Traffic Prediction Model for Dynamic ...
The availability of precise high-quality and high-volume data sets of traffic ... to forecast real-time traffic workload could make dynamic resource allocation more ...

Dynamic Sensorimotor Model for open-ended acquisition of tool-use
reach the target directly, fails, then make a detour to grasp a tool perceived as an extension of ... (e.g., action planing, tool-use) share common learning mech- anisms. A continuous .... The target is in the top-right corner, and, in the ... the em

adaptive model combination for dynamic speaker ...
as MAP [7]) and speaker space family (such as eigenvoice. [6]). .... a global weight vector is learned for all phone classes of test ..... Signal Processing, vol. 9, pp.

Model Dynamic RAD GRK.pdf
Average Data from. International Cities. Data from BPS, and. Recent studies. Method Multiple linear variables Energy Modelling. Analysis Growth, VKT (veh.kilo ...

a model for generating learning objects from digital ...
In e-Learning and CSCL there is the necessity to develop technological tools that promote .... generating flexible, adaptable, open and personalized learning objects based on digital ... The languages for the structuring of data based on the Web. ...

Statistical surface shape adjustment for a posable human hand model
statistical analysis of the shape difference from the measured actual human hand data in many different postures. To increase the number of precise shape.

a model for generating learning objects from digital ...
7.2.9 Tools for generating Learning Objects. ....................................................... ... 9 Schedule of activities . ..... ones: Collaborative Notebook (Edelson et. al. 1995) ...

Statistical surface shape adjustment for a posable human hand model
statistical analysis of the shape difference from the measured actual human hand data in many different postures. To increase the number of precise shape.

Hemicraniectomy: A New Model for Human ...
Berger improved his EEG recording technology, he was able to acquire EEG from .... approximate underlying Brodmannʼs area (BA; Koessler et al., 2009; Okamoto ...... (Canolty et al., 2006) that information obtained through interfrequency ...

Hemicraniectomy: A New Model for Human ...
rhythms, we hypothesized that that data from the elec- trodes placed over the ... patient with at least 4 weeks recovery time following their ... was 3 in the field.

Digital Learning Model-2016.pdf
Mapping A STEAM. Pathway. Expanding. Computer Science. & Technology. Course. Opportunities. Change Team. Student Help Desk. Preparing Students.