Brought to you by:
Paper

On extracting design principles from biology: I. Method–General answers to high-level design questions for bioinspired robots

and

Published 2 February 2015 © 2015 IOP Publishing Ltd
, , Citation M Haberland and S Kim 2015 Bioinspir. Biomim. 10 016010 DOI 10.1088/1748-3190/10/1/016010

1748-3190/10/1/016010

Abstract

When millions of years of evolution suggest a particular design solution, we may be tempted to abandon traditional design methods and copy the biological example. However, biological solutions do not often translate directly into the engineering domain, and even when they do, copying eliminates the opportunity to improve. A better approach is to extract design principles relevant to the task of interest, incorporate them in engineering designs, and vet these candidates against others. This paper presents the first general framework for determining whether biologically inspired relationships between design input variables and output objectives and constraints are applicable to a variety of engineering systems. Using optimization and statistics to generalize the results beyond a particular system, the framework overcomes shortcomings observed of ad hoc methods, particularly those used in the challenging study of legged locomotion. The utility of the framework is demonstrated in a case study of the relative running efficiency of rotary-kneed and telescoping-legged robots.

Export citation and abstract BibTeX RIS

1. Introduction

Nature provides a wealth of design ideas valuable to engineers because it holds the only examples of design other than our own [1]. Biology has lent the pattern for many original and successful materials, devices, and manufacturing methods [2, 3], and so 'biomimetics' has become a popular field of engineering research. It promises that by harvesting biology's secrets, we can apply our understanding of the natural world to improve human quality of life. But even though humans have looked to nature for technological inspiration for over 3000 years [4], development of biomimetic design methodology is ongoing. Indeed, [4] states that no general approach has been developed for biomimetics, [5] admits that while there are many case studies of biologically inspired design, these are only a first step toward developing a design methodology, and [6] notes that only recently have researchers begun to develop formal bioinspired design methods and tools. Several of these approaches are outlined in [3, 79], and all of these address the importance of extracting principles from biology, often for the purpose of tabulating these principles for subsequent automated lookup. However, these methods do not fully address how correct principles can be isolated from alternative hypotheses, even after they are inspired by observation of biological examples.

Although experimental biologists extract scientific principles using well-developed experimental practices and careful inductive reasoning, often their objective is to better understand purely biological systems, and so the results are not necessarily suitable for immediate application to engineering design due to differences between the biological and engineering domains. Biological systems may rely on unique capabilities currently beyond the performance limits of their engineering counterparts. For instance, in perhaps the most seminal publication regarding the mechanism of gecko adhesion [10], the authors admitted that production of 'small, closely packed arrays mimicking setae' was beyond the manufacturing technology of the time, and years of additional research were necessary to extend the abilities of component technologies before bioinspired prototypes became possible [1116]. Conversely, biological solutions are often limited by constraints not shared in engineering. A biological example may indeed perform a function of interest to engineers, but the particular approach taken by biology may not truly be optimized for the particular function [17]. Rather, it is the result of evolution, which must consider many other complex biological objectives and constraints [1] that are not often required in engineering, such as the needs to extract energy from food, to grow, and to reproduce. Consequently, engineers must assess whether a hypothesized bioinspired principle actually applies in the engineering domain and whether the biomimetic approach has the potential to outperform existing solutions before applying it to design.

The differences between the biological and engineering domains prompt the need to extract principles only as they apply to engineering design, as illustrated in figure 1. Although we have not found any general discussion of a procedure for extracting bioinspired principles for application in the engineering domain, we might attempt to generalize one of the ad hoc methods used by relevant studies. For instance, researchers have considered the following questions for legged robots:

  • What is the effect of spine flexibility on running speed and efficiency [1921]?
  • What is the effect of tails on quadruped maneuverability [22, 23]?
  • What is the effect of foot shape on energetic efficiency [24, 25]?
  • What is the effect of having toes on locomotion stability [26]?
  • What is the effect of the upper body on locomotion stability and efficiency [27]?
  • What is the effect of leg segmentation on robustness to disturbances [28, 29]?
  • What is the effect of leg configuration on energy loss of walking and running [30]?

General answers to these questions could serve as principles to guide robot design. Unfortunately, it would be difficult to generalize the conclusions of any of these studies to a broad class of realistic robots for one or both of the following reasons:

  • the use of a model with fixed parameters limits the applicability of the results to the particular system studied [1921, 24, 26, 27], and/or
  • the model used is too simple to provide convincing evidence of direct applicability to physical robots [22, 2830].

Because of these limitations, none of the ad hoc methods used in these studies (or any others we are aware of ) provides a complete basis for extracting principles from biology for application to engineering design.

Figure 1.

Figure 1. The biological and engineering domains have different characteristic elements and constraints. We seek to extract only the principles that are useful in engineering designs. (Cheetah photo from [18].)

Standard image High-resolution image

Surprisingly, extensions of the rigorous methods used by experimental biologists are not commonly found in the robotics literature. For example, biologists can carefully test physiological hypotheses by performing experiments on several specimens of a given type of organism. Based on their observation of multiple samples from a population they can induce the truth or falsity of a hypothesis for the entire species. We argue that engineers can similarly test design hypotheses by performing experiments on multiple specimens from a 'species' of robot.

In this paper, we present a methodology that employs rigid body dynamic simulations and optimization methods to answer high level design questions regarding the truth of a hypothesis for particular robot/controller designs and uses statistical techniques to generalize the results for a wide class of robots. In section 2, we will carefully define 'design principles' and 'principle extraction', and outline a framework for systematically and rigorously verifying hypothesized principles. Section 3 will present an example, section 4 will discuss the merits and shortcomings of the framework, and section 5 will conclude.

2. Methods: a general framework for bioinspired design principle extraction

Consider the design process as a procedure that begins with initial conceptual designs and refines design parameters iteratively until the final solution achieves certain design requirements including objectives and constraints. Compare this to the numerical process of solving a system of inequalities: a computer typically begins with an initial guess for the solution, and iteratively changes the independent variables until equality and inequality conditions are satisfied. While computer algorithms perform best with complete, quantitative relationships or partial derivatives, human designers typically evaluate these only sparsely. Human designers excel at achieving a final solution based on 'design principles': relationships between input design parameters and output objectives and constraints applicable to many systems. 'Biologically-inspired design principles' are, in the opinion of the authors, design principles derived in any part from the observation of biological systems1 , and 'principle extraction' is any process used to turn observation of biological systems into design principles. For example, the MIT Cheetah will employ a tail to assist in aerial maneuvers because the principle that tails are more effective than reaction wheels has been confirmed under the constraints of a typical quadrupedal robot [22]. Figures 2 and 3 illustrate our approach to the discovery of such biomimetic design principles, which consists of:

  • (i)  
    observation of nature to inspire biological design questions,
  • (ii)  
    transfer of the question from the biological domain to the engineering domain to prompt an engineering model,
  • (iii)  
    computer experiments with the engineering model to provide evidence for engineering principles, and
  • (iv)  
    analysis of the evidence to yield engineering principles.

Figure 2.

Figure 2. We present a process to filter the hypothesized principles from biology that are applicable to engineering design from those that are not.

Standard image High-resolution image
Figure 3.

Figure 3. Our approach to providing general answers to high-level design questions. (Running man graphic from [31].)

Standard image High-resolution image

Once confirmed, these principles can be used to inform robot design and, in the spirit of biomutualism, inspire additional hypotheses about the natural world. The remainder of the section details these steps.

2.1. Observation of nature inspires biological design questions

Nature provides an abundant supply of profound biological examples, but our ability to extract design principles from them relies to some extent on innate human curiosity and talent at recognizing patterns. Even when a biological example is particularly well adapted for performing a task similar to one desired of machines, humans must recognize the comparison before the extraction process can begin. Once this link is established, the researcher can observe and study the biological example further, often in collaboration with experimental biologists, in an attempt to form a question and a hypothesis about the relationship between the organism's features and its performance of the task. The researcher may first pose the question in the biological domain, 'What is the effect of this feature on the performance of the organism?'

2.2. Transfer of the question from the biological domain to the engineering domain prompts an engineering model

Although a physiological principle obtained by observation of biology may not be true in the engineering domain, a question inspired by observation of biology can be transferred to the engineering domain directly. The question becomes 'What is the effect of an engineering adaptation of this biological feature on the performance of an engineered system?' To perform this transfer, the researcher must define the high-level architecture of the the engineered system, including the adaptation of the bioinspired feature. In some cases, such as legged locomotion, there is already some consensus on the architecture of the engineered system: the bones of an animal are often replaced with plastic or metal links, the joints with bearings or flexures, and the muscles with electromagnetic or pressurized fluid actuators. In others systems, this transfer relies more on the researcher's innovation. For instance, when researchers have studied the effects of geometry and material on the adhesion performance of gecko-inspired adhesives, engineering adaptations of the adhesive gecko hairs have ranged from millimeter-scale elastomer pillars [32] to carbon nanotubes [33]. As with the solution of a new differential equation, there may not necessarily be a formulaic approach to defining the architecture of the engineering system; instead the researcher must make an educated guess to be tested. Any guess is admissible, but the utility of the results will depend on the choice of the engineering system that is studied.

Once the architecture has been established, the researcher has several options for answering the question: analysis, simulation, and physical experiment. Typically, analytical studies are impeded by the complexity of the system; principles can be deduced only in the most heavily simplified cases. Physical testing of engineering systems offer the greatest guarantees for the accuracy of conclusions, but constructing and testing multiple subject machines is usually prohibitively time-consuming and expensive. When systems governed by relatively well-understood physics can be modeled to simulate behavior quickly and inexpensively, computer experiments are often the method of choice for testing hypotheses. We will focus on in silico testing for our framework because of its ability to treat complex systems with relatively low resource requirements.

Computer experiments begin with a model of the relationship between the system's inputs and outputs. Simple models with few inputs are general in the sense that by avoiding representation of any particular system, they represent many different embodiments of a system simultaneously [34]. This feature would seem to make them particularly suitable for principle extraction studies. For instance, the SLIP model and variations have been used to study the effects of certain biologically-inspired features, such swing-leg retraction, on running performance metrics, such as efficiency [3537] and stability [36, 3842]. However, [43] compares some of these results for simple models, a detailed model, and a physical robot, and finds that the simple SLIP model is unreliable at predicting such complex behaviors. Therefore we suggest that unless simple models have been carefully verified against more detailed models and physical systems, they should only be used for explanation of observed phenomena and not be relied on for principle extraction. Instead, the model used to extract principles for a class of robots should be as detailed as that used to develop controllers for and/or simulate the behavior of a physical robot of that class.

The definition of the output of interest is also important. In the study of legged robots, there are many different ways of quantifying concepts like 'stability' and 'efficiency' [44, 45], and they may not be interchangeable. For example, the rate of recovery from infinitesimal disturbances may not at all be indicative of the magnitude of disturbances a robot can recover from [43]. Likewise, although a portion of a robot's energy expenditure may be due to impact with the ground, collisional losses and energy use per distance traveled may not always be strongly correlated [43]. Selection of the method used to quantify the output of interest cannot be based solely on computational efficiency, but also on its direct relevance to the performance or ability desired by robot designers.

2.3. Experiments with the engineering model provide data supporting engineering principles

Once the model can be used to calculate the system outputs given values for the inputs, an experiment must be performed to determine the relationship between selected inputs of interest and an output of interest. If this relationship is to be used as a design principle, it must be true for many systems, so the the system inputs other than the ones of interest cannot simply be left fixed at a single set of values chosen by the researcher.

For example, experimental biologists have shown that for both humans [46] and quadrupeds [47], adding distal mass to the limbs increases the energetic cost of locomotion more than adding the same mass proximally. This may inspire us to question whether the same is true of robots, and to answer this question we would build a model that calculates a metric of energy efficiency given the leg center of mass location and all other mechanical and controller parameters [48]. Although we seek the the effect of a single input of interest on a single output of interest, the legged robot has many mechanical and control parameters other than leg center of mass location, such as the leg segment lengths and the running gait produced by the controller. Over the wide range of possibilities that a designer might select for these other system inputs, it is uncertain a priori how the relationship we seek is affected.

Consequently, the experimental design must account for the possible variety in the system under study. This presents significant challenges for complex systems; here we propose solutions.

2.3.1. Experimental design

In the field of experimental design, it is common to select a value of the input as a 'control' and refer to other values of the input as different 'treatments'. The set of all possible combinations of the other inputs may be referred to as the 'population'. Then it may be possible to learn enough about the relationship of interest for design purposes by answering either of the following questions:

Question 1. What is the effect of the treatment on the percentage of the population with improved $\lt output\gt $, relative to the control?

Question 2. What is the effect of the treatment on the average (across the population) change in $\lt output\gt $, relative to the control?

While answering these questions does not provide detailed information about the coupling relationships among the inputs, the conclusions can still be applicable to the full class of system under study. Furthermore, these questions can be answered using Monte Carlo methods [49], that is, sampling at random from the population, measuring the effect of the treatment for these samples, and using the results to make an inference about the entire population. The computational cost of Monte Carlo methods can be far lower than that of testing all samples generated by varying parameters on a grid. Critically, we can also quantify our uncertainty in the conclusion with statistical tests, such as the binomial test [50] or a resampling method [49], and be sure to collect sufficient data to build confidence in the result.

Design of experiments [50, 51] is a vast field of applied statistics. Working knowledge of the fundamentals is imperative for performing principle extraction experiments, and many advanced topics might be beneficial as well. The purpose of this subsection is not to summarize statistics for experimenters, but to outline steps for answering questions of the two forms above, chosen because they require only simple techniques and relatively low computational resources to answer, yet their answers may be very insightful for design purposes. In section 4 we will mention other questions and techniques that may also be useful.

Use of statistics. An important difference between computer experiments with deterministic simulations and physical experiments is the absence of random error. Therefore, statistical techniques are not needed to quantify our uncertainty due to random error [52], but primarily the uncertainty in our conclusions due to the limited number of population members sampled.

We seek to understand a relationship f between input of interest x and scalar output $y=f(x)$ to inform a design. Specifically, we seek to answer questions 1 and 2 numerically, which are expressed mathematically as:

  • (1)  
    what is the percentage of the population $g({{x}_{{\rm t}}})=\frac{{{\int }_{P}}H\left( f({{x}_{{\rm t}}},{\bf p})-f({{x}_{{\rm c}}},{\bf p}) \right){\rm d}P}{{{\int }_{P}}{\rm d}P}$ for which output $f({{x}_{{\rm t}}},{\bf p})\gt f({{x}_{{\rm c}}},{\bf p})$, where H is the Heaviside step function, and
  • (2)  
    what is the average across the population $h({{x}_{{\rm t}}})=\frac{{{\int }_{P}}f({{x}_{{\rm t}}},{\bf p})-f({{x}_{{\rm c}}},{\bf p}){\rm d}P}{{{\int }_{P}}{\rm d}P}$ of the difference in outputs $f({{x}_{{\rm t}}},{\bf p})-f({{x}_{{\rm c}}},{\bf p})$

for treatment xt and control xc as other input parameters ${\bf p}$ vary across the population P. We cannot evaluate the integrals exactly, but we can use Monte Carlo methods to provide, at reasonable computational cost, an estimate for this value and a confidence interval in which the the answer is believed to lie.

Answer to question 1 . Using N random samples ${{{\bf p}}_{i}}$ from P, we can approximate the true percentage $g({{x}_{{\rm t}}})$ with the observed percentage

Equation (1)

a random variable. Then $\tilde{g}({{x}_{{\rm t}}})N$ follows the binomial distribution of N experiments and unknown proportion of success $q=g({{x}_{{\rm t}}})$, i.e. $\tilde{g}({{x}_{{\rm t}}})N\sim B(N,g({{x}_{{\rm t}}}))$. We can use this information to test a null hypothesis, that is, the likelihood of the observed $\tilde{g}({{x}_{{\rm t}}})$ being produced with an assumed q0, and likewise we can produce a binomial proportion confidence interval by any one of several methods [53, 54].

Answer to question 2 . Using N random samples ${{{\bf p}}_{i}}$ from P, we can approximate the true mean $h({{x}_{{\rm t}}})$ as the sample mean

Equation (2)

a random variable that follows an unknown distribution. Nonetheless, we can estimate a confidence interval for $h({{x}_{{\rm t}}})$ using resampling methods, such as the bootstrap [49].

Note that questions 1 and 2 both call for the practice of blocking: assigning similar, or in this case identical, test subjects ${{{\bf p}}_{i}}$ to all treatments and the control group [50] in a given trial. This improves the sensitivity of the experiment because it allows the effect of the treatments to be measured independently of any confounding factors. By performing trials with many different 'blocks', each with a randomly selected test subject ${{{\bf p}}_{i}}$, we collect data to make an inference about the population.

Variable partitioning. Once the form of the question is understood, the researcher must decide how to treat each of the input variables, and in doing so determines the scope of the study's applicability.

Fixed variables. Variables that assume fixed values for the duration of the experiment (e.g. the architecture of the model) we call 'fixed variables'. While sometimes necessary, these typically limit the generality of the study. On the other hand, a certain number of parameters may be held fixed without loss of generality as a consequence of the Buckingham Π theorem [55]. Note that we do not list fixed variables as inputs to the function of interest $f(x,{\bf p})$ because they do not change; they are instead considered to be part of f.

Sampled variables. Those variables ${\bf p}$ which vary only between trials, e.g. model parameters, we call 'sampled variables'. Choosing different values for different trials, as by random sampling from a population, broadens the scope of the study.

Independent variables. Those variables x which are varied and take on values xc and xt within a trial, e.g. the inputs of interest, are referred to as 'independent variables'. Multiple variables independently varied on a grid usually simplify interpretation of results, but can require extensive data collection. More efficient experiment designs for multiple variables exist (e.g. Latin hypercube or factorial designs [50]), but we limit our discussion here to a single, scalar independent variable.

Decision variables. Variables that vary depending on others within a trial, e.g. to optimize an objective and/or satisfy constraints, we call 'decision variables'. We discuss the use of such variables in 2.3.2. Although these limit the applicability of a study somewhat, sometimes it is necessary to treat variables as such when handling them as sampled variables or independent variables is impractical. We do not list decision variables as inputs to the function of interest $f(x,{\bf p})$ because they are not independent of, but rather determined by, the other inputs ${\bf p}$ and x.

Variable level selection. After the inputs have been partitioned into different types, the designer of a computer experiment has the task of defining the levels each will assume, the population from which they will be sampled, or the space over which they will be optimized. When doing so, the researcher must bear in mind that the purpose of the experiment is to provide evidence for or against the applicability of a principle for design purposes. The ranges of values should be large enough to include all the potential system designs for which the researcher wishes conclusions to be applicable. The ranges should be no larger than necessary, however, because inclusion of irrelevant members in a population can skew the results. Choice of the populations of sampled variables and bounds on decision variables deserve particular attention.

2.3.2. Experiment implementation

The use of computers dramatically simplifies the process of collecting data. Once the experiment has been designed, there is code to be written, but computers do most of the processing.

While the code to simulate the system of interest will be problem specific, and code to select sampled variables from the population is relatively straightforward to write, the use of optimization deserves special attention here. Inputs can be treated as decision variables when treating them as fixed variables limits the generality of the experiment, treating them as independent variables would require excessive computation, and treating them as sampled variables is impractical because defining the distribution of input values (or functions) is difficult. It is also justified when reasonable levels of the input are unknown, but is likely that the system designer would choose the level of the input to improve the output of interest. In our experience, inputs that must respect nonlinear constraints and/or inputs that are functions, rather than scalars or vectors, can often be treated only as either fixed variables or decision variables. Of the two options, treatment as decision variables seems more satisfying because rather than fixing the inputs at arbitrary values that may not be of interest, they assume optimal, bounding values that are relevant to researchers.

Optimization. In the study of legged robots, for example, the robot controller is an input that will affect the output of interest. The controller is a function that must satisfy certain constraints (e.g. the robot cannot fall), or else the output of interest cannot be evaluated. If there were a standard parametrization for a robot controller with accepted ranges for the parameters, perhaps those parameters could be treated as sampled variables, but this is not typical. We would prefer not to choose an arbitrary controller that satisfies the constraints because this limits the applicability of the study to robots using that particular controller. It is preferable, then, to treat the controller as a decision variable; we use the controller which optimizes the output of interest for the given fixed variables, sampled variables, and independent variables. Fortunately, it is often reasonable to expect that regardless of the physical robot design, the designer will typically choose a controller that will nearly optimize the output of interest (e.g. stability, efficiency, etc...), and so extraction of a principle regarding the optimal value of the output of interest is of interest to many robot designers. This motivates a brief discussion of optimal control, which we use to determine this optimal value of the output of interest.

Analytical optimal control is usually impossible for all but the simplest models, and thus computational optimal control techniques must be employed [56]. As is common in numerical algorithms, computational optimal control schemes typically begin with a guess of the solution, a 'seed'. We believe that it is important for this seed to be generated automatically, as hand-selecting the seed for many test subjects would be time-consuming, and more importantly, it is undesirable for human intervention to affect the results (i.e. which local optimum is found). We prefer that the seed be selected at random from the space of all possible seeds to minimize the effect of human choice on the output of the experiment, and we perform optimization from many such seeds ('multistart' [57]) to reduce the effect of chance occurrences on experimental results.

Computational optimal control is typically divided into direct methods, which discretize the optimal control problem in order to formulate a discrete nonlinear programming problem, and indirect methods, which use optimal control theory to generate necessary conditions for optimality in the form of a boundary value problem that can be discretized and solved numerically [56, 58]. Similar discretization ideas can be employed in either direct or indirect methods; single shooting, multiple shooting, and collocation are common discretization methods [56, 58]. We prefer the direct approach, as it is easier to use, and we discretize using collocation methods because they result in a numerical problem that is easier to solve without a good initial guess [56]. In particular, we find pseudospectral methods [59, 60] to be remarkably robust to the initial guess; regularly converging from a random seed to a solution confirmed by Runge–Kutta integration [61] to satisfy the dynamics and constraints. Potentially important to their success is use of accurate partial derivatives; complex step approximation [62], algorithmic differentiation [63], and analytical differentiation are all more accurate and potentially faster than the typical first-order finite difference approximation. When partial derivatives are not available, derivative-free optimization [64] techniques and/or metaheuristics [65] might be used.

2.4. Analysis of the data yields engineering principles

The statistical tests mentioned in 2.3.1 are the primary analyses as they answer the questions 1 and 2 the experiment was designed to investigate, and the answers to those questions are the primary conclusions of the study. However the experiment yields a tremendous amount of data which can be used for exploratory data analysis [66]. For instance, it may be informative to plot histograms of the output of interest, as a distribution with an unusual structure may be worthy of deeper investigation. Also, the output of interest and possibly other outputs can be plotted against each input, or more generally functions of inputs and outputs can be plotted against one another. Any notable trends may prompt conclusions or suggest ideas for further study.

Although these experiments reveal the relationship between inputs and outputs, they do not always help us understand why the observed relationship between inputs and outputs exists. Knowing the overall relationship may be sufficient for design purposes, but intuitively understanding the relationship may also be desirable. In this case, the studies can be refined to explore the chain of causality that causes the inputs to affect the outputs as observed in the original experiment. For instance, if the output of interest can be subdivided into several parts, and similar experiments are performed with each part as the output of interest, then analysis of the new results may make it easier to relate the results of the original experiment to physical intuition.

3. A case study in principle extraction: can a bioinspired knee improve the efficiency of a legged robot?

The general framework presented in section 2 is illustrated with a brief case study in legged robots.

3.1. Introduction

Several early, influential legged robots from the MIT Leg Lab [6771] and more recent examples from other labs [72, 73] have used legs with telescoping, or prismatic joints. However, all animals and many other influential legged robots [7477] are equipped with articulated legs, that is, legs with a revolute joint. While several advantages of revolute-jointed legs have been suggested, including simpler mechanical design and construction [74], improved dynamic stability and robustness [28], and variable mechanical advantage [78, 79], the relative advantages of these two joint options are still not entirely understood [28]. To elucidate the performance differences between these two leg types, we will use our framework to guide an investigation of the effect of leg joint type on the energetic efficiency of running robots.

3.2. Method

3.2.1. Question.

All running animals have legs with revolute joints. While nature's unique constraints may have discouraged the evolution of prismatic joints [78], revolute joints may also have offered biological systems particular advantages. We ask: What is the effect of revolute joints on the energetic efficiency of running?

3.2.2. Model

The question can be transferred to the engineering domain as: What is the relative efficiency of revolute joints compared to telescoping joints for running robots? We begin with the running robot model used in [34] and illustrated in figure 4: a point body of mass m, massless point foot, and massless leg actuated to apply a force F between the foot and body.

Figure 4.

Figure 4. The model used in this study consists of a point mass m acting under the influence of gravity g and leg force F. Both telescoping and kneed versions have the same electric motor which can produce maximum power P*. Consequently, in the telescoping model, the maximum force depends on leg contraction/extension speed but is independent of leg length; for the kneed leg, however, the maximum force is a function of both leg length and speed.

Standard image High-resolution image

The body accelerates according to gravity g during the 'flight' phase and is influenced by both gravity and the leg force during 'stance'. During stance, the state of the model is parametrized by the length of the leg l, the speed of leg extension $\dot{l}$, the angle of the leg with respect to vertical θ, and the angular velocity of the leg $\dot{\theta }$. The controls are the angle θtd and length ltd of the leg at touchdown, when the foot touches the ground, the flight phase ends, and the stance phase begins; the length of the leg at takeoff lto, when the foot leaves the ground, the stance phase ends, and the flight phase begins; and the force trajectory F(t) of the leg during stance.

We modify the model to make it more representative of running robots: the source of actuation is an electric motor whose maximum torque is constrained by an affine function of its speed, as required by the equilibrium equations of a direct current motor [80]. In the 'telescoping leg' version of the model, the massless leg is endowed with a prismatic joint and a massless, ideal rack and pinion to convert the motor's torque and rotation to a force and translation. In the 'kneed leg' model, the motor has a massless, ideal gearbox which acts on a revolute joint positioned in the center of the leg.

The maximum positive or negative voltage that can be applied to the motor, and thus the maximum mechanical power P* the motor can produce, is the same for the two models. However, the effective force limits of the two legs are different, although both can be written in the same form,

Equation (3)

where ${{\Delta }_{l}}=\sqrt{l_{0}^{2}-{{l}^{2}}}$ is the compression of the leg from its maximum length, ${{F}_{0}}({{\Delta }_{l}})$ is the stall force, and $\dot{l}$ is the instantaneous rate of change in leg length. The difference is that for the telescoping leg ${{F}_{0}}({{\Delta }_{l}})={{F}_{0}}$ is a constant (independent of displacement) while for the kneed leg ${{F}_{0}}({{\Delta }_{l}})=\frac{{{T}_{0}}}{{{\Delta }_{l}}/2}$, where T0 is the stall torque of its actuator (motor and gearbox).

The inefficiency is quantified using the absolute mechanical cost of transport

Equation (4)

the integral of the absolute mechanical power $|F\dot{l}|$ over the step time tstep normalized by the product of the robot's weight $mg$ and distance traveled Δx in one cycle (step) of periodic motion.

We have chosen a simple model against our recommendations in section 2.2 only to demonstrate the effectiveness of the approach presented in section 2 without being distracted by the complexity of a very high-dimensional problem. Using a model and output similar to [34] has the added advantage of allowing us to compare our results to those of the reference. The framework itself scales to higher-dimensional systems with modest increase in computational complexity, and indeed this was the motivation for the choice of statistics and optimization techniques we have presented. A more realistic, higher dimensional case study is presented in our companion paper [81].

3.2.3. Experiment

The questions can now be formulated more specifically as:

  • (i)  
    For what percentage of the design space of running robots would a 'kneed leg' robot have lower optimal mechanical cost of transport than a 'telescoping leg' robot?
  • (ii)  
    Across the design space of running robots, what is the average improvement in the optimal mechanical cost of transport for a 'kneed leg' over a 'telescoping leg'?

The results of the experiment are intended to be applicable to all running robots that are well-approximated by our model, but there are some limitations:

  • (i)  
    The experiment will measure the optimal cost of transport cmt over all possible limit cycles (periodic motions) of $l(t),\dot{l}(t),\theta (t),\dot{\theta }(t)$ that can be achieved with force trajectories F(t) under the dynamics. Because these variables are treated as decision variables, the results are most applicable to robots following efficient trajectories.
  • (ii)  
    Additionally, rack and pinion stall force F0 is optimized for the telescoping leg, and gearbox stall torque T0 is optimized (independently) for the kneed leg. Therefore, the results are applicable to robots with properly tuned transmissions.
  • (iii)  
    As explained in [34], the dimensionless Froude number $Fr=\frac{{{v}_{x,{\rm avg}}}}{\sqrt{gl}}$ must be constrained or the optimal solution would tend toward zero average horizontal speed ${{v}_{x,{\rm avg}}}$, which is not useful. We limit the experiment to $Fr$ in the range [1, 5], which encompasses much of the range of $Fr$ normally observed in nature [82].
  • (iv)  
    Likewise, the dimensionless step length $Sl=\frac{{{\Delta }_{x}}}{{{l}_{0}}}$ must be constrained or the optimal solution would tend toward zero-length steps, which is not realistic [34]. We limit the experiment to $Sl$ in the range [1,5], which encompasses the range of step lengths normally observed in nature [82].
  • (v)  
    The dimensionless duty factor D.F. = $\frac{{{t}_{{\rm stance}}}}{{{t}_{{\rm step}}}}$ must be constrained or the optimal solution would converge toward impulsive running, that is, infinitely high forces $F\to \infty $ over a vanishingly short stance ${{t}_{{\rm stance}}}\to 0$. However, cmt approaches its optimal value asymptotically, so D.F. is bounded below by a value (0.05) considered small enough to be representative of the optimal, impulsive gait.
  • (vi)  
    Only four parameters remain: $m,g,{{l}_{0}}$, and P*. According to the Buckingham Π theorem, the relationship among the N variables can be expressed as a relationship among $N-M$ dimensionless parameters, where M is the number of fundamental physical units. In this case M = 3 corresponding with the physical units of mass, length, and time. Consequently, we fix $m,g$, and l0 without loss of generality and constrain dimensionless power ratio ${\rm Pr} =\frac{{{P}^{*}}}{mg{{v}_{x,{\rm avg}}}}$. We limit the experiment to ${\rm Pr} $ in the range [0.25, 1.25]. At the lower limit, the robot inefficiency is guaranteed to be less than a typical human cost of transport [83], while the upper limit would provide sufficient power for the machine to climb against gravity at a rate 25% greater than the average horizontal speed.

The different types of variables and their ranges are tabulated in tables 1–3.

Table 1.  Fixed variables, i.e., variables that assume fixed values for the duration of the experiment. These variables are only used to keep the equations in dimensional, physically-intuitive form; they do not affect the optimal cmt as a consequence of the Buckingham Π theorem. Therefore, the experiment remains valid for robots with any values of these variables.

Symbol Description Range Units
m Body mass 80 kg
l0 Max leg length 1 m
g Gravity 9.81 m s−1

Table 2.  Sampled variables, i.e., variables for which trials are performed at randomly sampled values to broaden the applicability of the experiment to a wide class of robots and gaits.

Symbol Description Range Units
$Fr$ Froude number [$1,5$]
$Sl$ Step length [$1,5$]
${\rm Pr} $ Power ratio [$0.25,1.25$]

Table 3.  Decision variables, i.e., those which are optimized to achieve minimal cmt. Values in gray are de-emphasized because they do not affect the experiment; they are sufficiently generous that they are rarely observed to be active. When a bound happens to be active, the corresponding results should be discarded to prevent the choice of bound value from affecting results. Values in black are 'hard' bounds set by physics or by the intent of the experiment, except for the lower bound on duty factor, which is sufficiently large to prevent numerical difficulties associated with a vanishing stance phase but sufficiently small to permit approximation of the optimal cmt.

Symbol Description Range Units
F0 Stall force [$0,100\;mg$] N
T0 Stall torque [$0,100\;mg{{l}_{0}}$] Nm
l(t) Leg length [$0.75\;{{l}_{0}},{{l}_{0}}$] m
$\dot{l}(t)$ Leg length rate [${{\dot{l}}_{l}},{{\dot{l}}_{u}}$] m s−1
$\theta (t)$ Leg angle [$\frac{\pi }{3},\frac{2\pi }{3}$] rad
$\dot{\theta }(t)$ Leg angle rate [${{\dot{\theta }}_{l}},{{\dot{\theta }}_{u}}$] rad s−1
F(t) Leg force [$0,100\;mgl$] N
D.F. Duty factor [$0.05,0.49$]

For each of many sets of sampled variables, we choose the decision variables to minimize the mechanical cost of transport for a robot with a 'kneed leg' and again, independently, for a robot with a 'telescoping leg'. For each point in the design space, this optimization is performed using General Pseudospectral Optimal Control Software [60, 8486] from many seeds randomly generated from within the decision variables' bounds to avoid the potential for single local optima to skew the results.

3.3. Analysis

3.3.1. Results and discussion

As expected from the results of [34], optimizations consistently converged to an impulsive running gate for $Fr\geqslant 1.5$: a bouncing gait characterized by a single, short burst of high force from the actuator during stance followed by a long ballistic trajectory. Results reported in [34] for running at $Fr=1,Sl=1$ also match those of the present study:

  • (i)  
    The optimal gait is a 'pendular run' characterized by bursts of high force at the beginning and end of a relatively long stance followed by a moderate-length ballistic flight.
  • (ii)  
    The CoT reported in [34] of $\approx 0.1$ is comparable to that of $\approx 0.3$ found in the present study after considering that we include not only positive but also negative work in the CoT definition and that our addition of actuator constraints and a duty factor lower bound can only increase cost.

Finally, all optimizations from randomly selected seeds resulted in tightly clustered CoT values (as shown in figure 7), which suggests that the local minima found are actually close approximations to global optima. Together, these are taken as evidence that the dynamics are correctly implemented and that the optimal control algorithm is working properly. Also, optimization performed with randomly selected fixed variables, instead of those reported in table 1, resulted in nearly identical CoT values so long as the dimensionless parameters remained constant. This confirms that the results of the study are applicable to robots without dependence on m, l0, and g as expected according to the Buckingham Π theorem.

Figure 7.

Figure 7. The effect of the ratio of normalized step length ($Sl$) to Froude number squared ($Sl/F{{r}^{2}}$) on the minimal mechanical cost of transport. All data points, that is, all local minima from the multiple random seeds, are included to show how tightly the local minima cluster.

Standard image High-resolution image

Figure 5 presents the relative efficiency of the kneed leg with respect to the telescoping leg for a range of Froude number $Fr$ and normalized step length $Sl$ at fixed power ratio Pr = 1. It graphically answers question 1 for the Pr = 1 cross-section of the population by shading regions according to whether the kneed leg or telescoping leg are more efficient. Data from the 17 × 17 grid suggests that the kneed leg is more efficient for approximately 11.2% of 'successful samples', that is, points in the space for which both kneed and telescoping legs had running solutions.

Figure 5.

Figure 5. Design space shaded by efficiency of kneed leg relative to telescoping leg. 'Null hypothesis', that there is no efficiency difference between telescoping and kneed leg, applies where a t-test on the data for that point did not indicate significance at the p = 0.01 level. 'No data' means that optimization did not converge to a feasible solution from any seed.

Standard image High-resolution image

The Monte Carlo approximation with only 100 samples agrees surprisingly well; the kneed leg outperformed the telescoping leg in 10 (11.1%) of the N = 90 successful samples. Referring to the binomial distribution, the 95% confidence interval for the true percentage of the space that is more efficient with a telescoping leg is 5.5%–19.5%.

Figure 6 shows the difference in minimal CoT of the kneed leg and the minimal CoT of the telescoping leg as as a function of Froude number $Fr$ and normalized step length $Sl$ at fixed power ratio Pr = 1. The sample mean of the height of the gridpoints, 0.0055, answers question 2: the minimal CoT of the kneed leg is about 4.6% higher than the minimal CoT of the telescoping leg, on average.

Figure 6.

Figure 6. The effect of Froude number ($Fr$) and normalized step length ($Sl$) on difference in minimal mechanical cost of transport (kneed leg minus telescoping leg). Where local minima resulting from the multiple random seeds were not identical (but tightly clustered), the surface represents the difference in the sample means.

Standard image High-resolution image

The Monte Carlo approximation agrees, with an average difference between the two CoTs of 0.0068. The bootstrap reference distribution is generated by randomly selecting N = 90 'resamples' with replacement from the observed N = 90 sample differences, taking the mean of the resamples, and repeating this process 10 000 times. Comparing the sample mean of the observed samples to this reference distribution permits the generation of an approximate 95% confidence interval of the true mean difference between 0.0045 and 0.0105.

The close agreement between estimates from data evaluated on the grid and from the Monte Carlo data demonstrate that Monte Carlo approximation can be a useful tool to estimate such statistics with acceptable confidence intervals, but with far lower computation. For this problem, evaluating the minimal cost of transport on our coarse grid (17 × 17) in two dimensions ($Fr,Sl$) as in figure 5 was reasonable on a 24-core server (about 43 h), but producing a reasonably fine grid (say, $41\times 41\times 21$) in all three dimensions (${\rm Pr} ,Fr,Sl$) would be very time-consuming (about 31 weeks). For realistic models in higher dimensions, use of grids is entirely infeasible, but we can approximate statistics using the Monte Carlo method with no additional computation. The number of samples needed depends only on the desired width or confidence of the interval; it is independent of the dimensionality of the population space. Table 4 presents the answers to questions 1 and 2 using a Monte Carlo approximation with 100 sample points from the full $Fr$-$Sl$-${\rm Pr} $ population space.

Table 4.  Summary of results for the full population. As the population now includes samples with lower ${\rm Pr} $, the dominance of the telescoping leg is even more clear. This might be expected, as a difference between the two configurations is only possible for low ${\rm Pr} $; for high ${\rm Pr} $ the constraints distinguishing the two joint types (equation (3)) are less likely to become active.

% Population kneed leg is more efficient
Sample % C.I. lower C.I. upper
3 % 1% 10 %
Mean additional CoT of kneed leg
Sample mean C.I. lower C.I. upper
0.0060 0.0044 0.0081

While the Monte Carlo method presented does not yield detailed spatial information, neither is analysis restricted to the calculation of scalar statistics. Exploratory analysis of the data can produce additional interesting results: for instance, inspection of figure 6 suggests that perhaps $Fr$ and $Sl$ do not affect minimal CoT independently, rather it is a ratio between the two that is most important. Figure 7 shows the minimal mechanical cost of transport as a function of ratio $\frac{Sl}{F{{r}^{2}}}$. The trend appears quite linear with little scatter, suggesting that the mechanical cost of transport does not depend strongly on the speed and step length independently, but only on a function of the two dimensionless parameters. Both the telescoping-leg and kneed-leg runner are efficient when running fast with small steps, but less efficient when running slowly with large steps. Furthermore, even at the magnification level in the figure it is possible to see that none of the points where the kneed leg is more efficient occur with a CoT greater than 0.1 or a ratio $Sl/F{{r}^{2}}$ greater than about 0.7. This tells us that the kneed legs are only more efficient for relatively low stride lengths or high speed running, as is apparent in figure 5.

3.3.2. Conclusion

The results support the following design principle, applicable to robots of all sizes and weights running according to the model studied: the optimal efficiency of telescoping legs tends to match or exceed that of kneed legs. Although this is only a strong statistical tendency and may not hold for all possible robots within the range studied, it serves as a useful guideline to improve performance when resources do not permit more detailed study for design of a particular machine.

4. Discussion

The motivation of this work is not to make particular robot design recommendations based on the simple running model above but to encourage greater rigor in future biomimetic principle extraction research. We hope the considerations discussed find use in legged robotics and other fields. In fact, the latter half of the framework may be used to test hypotheses for any system model, biologically inspired or not. Before concluding, there are a number of strengths and limitations of the current work and directions for further research we wish to discuss.

Recall that the answer to question 1 is that for q percent of the researcher-defined design space a particular treatment improves the output of interest relative to the control. This does not necessarily imply that an engineer who has designed a machine that falls within the design space has a q percent chance of improving the output of interest by applying the treatment rather than the control. This is because designing a machine is not the same as random sampling of a machine from the researcher's design space; we cannot assume that a design process will converge to all designs with the same probability density of random sampling. Nonetheless, when q is found to be especially high ($\approx 100\%$) or low ($\approx 0\%$), the designer might still decide that in the absence of other information and lack of resources to perform more focused study that the better design decision would be to implement the treatment or not as recommended by the general study. When q is neither high nor low ($\approx 50\%$), this suggests to the designer that more information is needed to make a good decision. The answer to question 2, the average improvement due to the treatment, should be used in conjunction: it will suggest whether the treatment is very important or not. Together, these results can be sufficient to proceed with design, or to suggest the importance of performing a more focused study.

We recognize that the answers to questions 1 and 2 do not provide all the information a designer might need to optimize a design. More detailed understanding of the coupled effect of several input variables on several output variables can be ascertained by performing more extensive experimentation using a spatially structured experimental design (e.g. factorial, latin square, etc...) and regression analysis, albeit at greater computational cost. Also, ANOVA or other techniques should be utilized when comparing multiple treatments or performing post-hoc testing on the data to avoid 'false alarms', that is, Type I error. Bayesian statistics and Gaussian processes may be useful for making inferences about the location of global minima given many local minima. New research on sampling from manifolds [87] may allow some variables currently classified by necessity as decision variables to be treated as sampled variables instead, improving the generality of results.

Previously we defined design principles as 'relationships between input design parameters and output objectives and constraints applicable to many systems', and the framework presented here is formulated with such a definition in mind. We recognize that this may not be a universal definition of the phrase 'design principle', but would argue that there are benefits in attempting to view the principle extraction process from this perspective. Thinking about biological and engineering design in terms of functions, inputs, and outputs is quite general, and is also very useful because it helps frame them in the construct of mathematics, which provides a vast array of tools for the researcher and designer.

Also, the framework is not a complete, automatic procedure to extract engineering design principles directly from biological data. Rather, it is a process for generating rich sets of evidence in support of (or against) a hypothesized principle that can be analyzed to generate statistically significant conclusions. Just as there is no substitute for creativity in proposing conceptual designs for a machine, so human understanding and ingenuity are needed to hypothesize a design principle to test. The importance of scientific understanding of the biological system through experimental biology has already been highlighted as essential to investigating the right question and proposing a fruitful model.

5. Conclusion

In this article we have demonstrated the importance of principle extraction in the biomimetic design process, outlined some of the difficulties of isolating design principles from biological examples, and suggested a framework under which to approach principle extraction that overcomes these difficulties. Although it is not a strictly linear, algorithmic process from biological observation to engineering conclusions, it highlights some of the important decisions that the researcher must make and suggests tools for verifying the researcher's hypotheses using Monte Carlo computer experiments and optimization. Overall, the framework guards against logical fallacies which might otherwise plague principle extraction studies, namely false analogy and hasty generalization, in the hope that future principle extraction research will produce reliable conclusions to advance the capabilities of biologically-inspired design.

Acknowledgments

This work was supported by the Defense Advanced Research Projects Agency through M3 Program Award W91CRB-11-C-0048. We thank Kim Vandiver and Peko Hosoi for their insightful suggestions for improving this work.

Footnotes

  • This does not require that the design principle actually be observed in biological systems.

Please wait… references are loading.
10.1088/1748-3190/10/1/016010