Keywords

1 Introduction

The geometry of a linear pentapod is given by the five base anchor points \(\mathsf{M}_i\) in the fixed system \(\varSigma _0\) and by the five collinear platform anchor points \(\mathsf{m}_i\) in the moving system \(\varSigma \) (for \(i=1,\ldots ,5\)). Each pair \((\mathsf{M}_i,\mathsf{m}_i)\) of corresponding anchor points is connected by a SPS-leg, where only the prismatic joint is active.

If the geometry of the linear pentapod is given as well as the lengths \(R_i\) of the five pairwise distinct legs, it has generically mobility 1. This degree of freedom corresponds to the rotational motion about the carrier line \(\mathsf{p}\) of the five platform anchor points. As this rotation is irrelevant for applications with axial symmetry (e.g. 5-axis milling, laser or water-jet engraving/cutting, spot-welding, spray-based painting, etc.), these manipulators are of great practical interest. Nevertheless configurations should be avoided where the linear pentapod gains an additional uncontrollable mobility, which is referred as self-motion.

1.1 Review on Self-motions of Linear Pentapods

The self-motions of linear pentapods represent interesting solutions to a problem posed 1904 by the French Academy of Science for the Prix Vaillant, which is also known as Borel-Bricard problem (cf. [2, 3]). This still unsolved kinematic challenge reads as follows: “Determine and study all displacements of a rigid body in which distinct points of the body move on spherical paths.”

For the special case of five collinear points the Borel-Bricard problem was studied by Darboux [5, p. 222], Mannheim [6, p. 180ff] and Duporcq [7] (see also Bricard [3, Chap. III]). A contemporary and accurate reexamination of these old results, which also takes the coincidence of platform anchor points into account, was done in [1] yielding a full classification of linear pentapods with self-motions.

Beside the architecturally singular linear pentapods [1, Corollary 1] and some trivial cases with pure rotational self-motions [1, Designs \(\alpha \), \(\beta \), \(\gamma \)] or pure translational ones [1, Theorem 1] there only remain the following three designs:

Under a self-motion each point of the line \(\mathsf{p}\) has a spherical (or planar) trajectory. The locus of the corresponding sphere centers is a cubic space curve \(\mathsf{P}\), where the mapping from \(\mathsf{p}\) to \(\mathsf{P}\) is named \(\sigma \). \(\mathsf{P}\) intersects the ideal plane in one real point \(\mathsf{W}\) and two conjugate complex ideal points, where the latter ones are the cyclic points \(\mathsf{I}\) and \(\mathsf{J}\) of a plane orthogonal to the direction of \(\mathsf{W}\). \(\mathsf{P}\) is therefore a so-called straight cubic circle. The following subcases can be distinguished:

  • \(\mathsf{P}\) is irreducible:

    • \(\sigma \) maps the ideal point \(\mathsf{U}\) of \(\mathsf{p}\) to \(\mathsf{W}\) (Type 5 according to [1]).

    • \(\sigma \) maps \(\mathsf{U}\) to a finite point of \(\mathsf{P}\) (Type 1 according to [1]).

  • \(\mathsf{P}\) splits up into a circle and a line, which is orthogonal to the carrier plane of the circle and intersects the circle in a point \(\mathsf{Q}\). Moreover \(\sigma \) maps \(\mathsf{U}\) to a point on the circle different from \(\mathsf{Q}\) (Type 2 according to [1]).

1.2 Basics on Line-Symmetric Motions

Krames (e.g. [4, 10]) studied special one-parametric motions (Symmetrische Schrotung in German), which are obtained by reflecting the moving system \(\varSigma \) in the generators of a ruled surface of the fixed system \(\varSigma _0\), which is the so called basic surface. These so-called line-symmetric motions were also studied by Bottema and Roth [8, Sect. 7 of Chap. 9], who gave an intuitive algebraic characterization in terms of Study parameters \((e_0:e_1:e_2:e_3:f_0:f_1:f_2:f_3)\), which are shortly repeated next.

All real points of the Study parameter space \(P^7\) (7-dimensional projective space), which are located on the so-called Study quadric \(\Psi :\,\sum _{i=0}^3e_if_i=0\), correspond to an Euclidean displacement with exception of the 3-dimensional subspace \(e_0=e_1=e_2=e_3=0\), as its points cannot fulfill the condition \(N\ne 0\) with \(N:=e_0^2+e_1^2+e_2^2+e_3^2\). The translation vector \({\mathbf s}:=(s_1,s_2,s_3)^T\) and the rotation matrix \({\mathbf R}\) of the corresponding Euclidean displacement \({\mathbf m}_i\mapsto {\mathbf R}{\mathbf m}_i + {\mathbf s}\) are given for \(N=1\) by:

$$\begin{aligned} {\begin{matrix} s_1&{}=-2(e_0f_1-e_1f_0+e_2f_3-e_3f_2), \quad s_2=-2(e_0f_2-e_2f_0+e_3f_1-e_1f_3), \\ s_3&{}=-2(e_0f_3-e_3f_0+e_1f_2-e_2f_1), \\ {\mathbf R} &{}= \begin{pmatrix} r_{11} &{} r_{12} &{} r_{13} \\ r_{21} &{} r_{22} &{} r_{23} \\ r_{31} &{} r_{32} &{} r_{33} \end{pmatrix}= \begin{pmatrix} e_0^2+e_1^2-e_2^2-e_3^2 &{} 2(e_1e_2-e_0e_3) &{} 2(e_1e_3+e_0e_2) \\ 2(e_1e_2+e_0e_3) &{} e_0^2-e_1^2+e_2^2-e_3^2 &{} 2(e_2e_3-e_0e_1) \\ 2(e_1e_3-e_0e_2) &{} 2(e_2e_3+e_0e_1) &{} e_0^2-e_1^2-e_2^2+e_3^2 \end{pmatrix}. \end{matrix}} \end{aligned}$$

There always exists a moving frame (in dependence of a given fixed frame) in a way that \(e_0=f_0=0\) holds for a line-symmetric motion. Then \((e_1:e_2:e_3:f_1:f_2:f_3)\) are the Plücker coordinates (according to the convention used in [8]) of the generators of the basic surface with respect to the fixed frame.

1.3 Line-Symmetric Self-motions of Linear Pentapods

It is well known (cf. [7, Sect. 15], [3, Sect. 12]) that the self-motions of Type 5 are obtained by restricting the Borel-Bricard motionsFootnote 1 (also known as BB-I motions) to a line. Note that Krames gave a detailed discussion of this special case in [4, Sect. 5], where he also pointed out the line-symmetry of BB-I motions.

Beside these BB-I motions, there also exist line-symmetric motions (so-called BB-II motions), where every point of a hyperboloid carrying two reguli of lines has a spherical path. It is known (cf. [9, p. 24] and [10, p. 188]) that the corresponding sphere centers of lines, belonging to one regulus,Footnote 2 constitute irreducible straight cubic circles, which imply examples of Type 1 self-motions. It should be noted that there also exist degenerated cases where the hyperboloid splits up into the union two orthogonal planes, which contain examples of Type 2 self-motions.

A simple count of free parameters shows that not all self-motions of Type 1 (5-parametric setFootnote 3 of motions where all points of a line have spherical paths) can be generated by BB-II motions (which produce only a 4-parametric setFootnote 4). The same argumentation holds for Type 2 self-motions and the mentioned degenerated case.

As a consequence the question arise whether all self-motions of linear pentapods of Type 1 and Type 2 can be generated by line-symmetric motions . If this is the case we can apply a construction proposed by Krames [4, p. 416], which is discussed in Sect. 4, yielding new solutions to the Borel-Bricard problem.

Finally it should be noted that a detailed review on line-symmetric motions with spherical trajectories is given in [11, Sect. 1].

2 On the Line-Symmetry of Type 1 and Type 2 Self-motions

For our calculations we do not select arbitrary pairs \((\mathsf{m}_i,\mathsf{M}_i)\) of \(\mathsf{p}\) and \(\mathsf{P}\), which are in correspondence with respect to \(\sigma \) (\(\Leftrightarrow \) \(\sigma (\mathsf{m}_i)=\mathsf{M}_i\)), but choose the following special ones:

\(\mathsf{M}_4\) equals \(\mathsf{W}\), \(\mathsf{M}_2\) coincides with \(\mathsf{I}\) and \(\mathsf{M}_3\) with \(\mathsf{J}\). The corresponding platform anchor points are denoted by \(\mathsf{m}_4\), \(\mathsf{m}_2\) and \(\mathsf{m}_3\), respectively. As \(\mathsf{M}_i\) are ideal points the corresponding points \(\mathsf{m}_i\) are not running on spheres but in planes orthogonal to the direction of \(\mathsf{M}_i\). Therefore these three point pairs imply three so-called Darboux conditions \(\varOmega _i\) for \(i=2,3,4\). Moreover we denote \(\mathsf{U}\) as \(\mathsf{m}_5\) and its corresponding finite point under \(\sigma \) by \(\mathsf{M}_5\). This point pair describes a so-called Mannheim condition \(\varPi _5\) (which is the inverse of a Darboux condition). The pentapod is completed by a sphere condition \(\varLambda _1\) of any pair of corresponding finite points \(\mathsf{m}_1\) and \(\mathsf{M}_1\).

In [1] we have chosen the fixed frame \(\mathscr {F}_0\) in a way that \(\mathsf{M}_1\) equals its origin and \(\mathsf{M}_4\) coincides with the ideal point of the z-axis. Moreover we located the moving frame \(\mathscr {F}\) in a way that \(\mathsf{p}\) coincides with the x-axis, where \(\mathsf{m}_1\) equals its origin.

For the study at hand it is advantageous to select a different set of fixed and moving frames \(\mathscr {F}_0^{\prime }\) and \(\mathscr {F}^{\prime }\), respectively:

  • As \(\mathsf{M}_2\) and \(\mathsf{M}_3\) coincides with the cyclic points, we can assume without loss of generality (w.l.o.g.) that \(\mathsf{M}_5\) is located in the xz-plane (as a rotation about the z-axis does not change the coordinates of \(\mathsf{M}_1,\ldots ,\mathsf{M}_4\)). Moreover we want to apply a translation in a way that \(\mathsf{M}_5\) is in the origin of the new fixed frame \(\mathscr {F}_0^{\prime }\). Summed up the coordinates with respect to \(\mathscr {F}_0^{\prime }\) read as:

    $$\begin{aligned} {\mathbf M}_5=(0,0,0),\quad {\mathbf M}_1=(A,0,C) \quad \text {with} \quad A\ne 0 \end{aligned}$$
    (1)

    as \(A=0\) implies a contradiction to the properties of \(\mathsf{P}\) for Type 1 and Type 2 pentapods given in Sect. 1.1. Moreover, \(\mathsf{M}_2\), \(\mathsf{M}_3\) and \(\mathsf{M}_4\) are the ideal points in direction \((1,i,0)^T\), \((1,-i,0)^T\) and \((0,0,1)^T\), respectively.

  • With respect to \(\mathscr {F}_0^{\prime }\) the location of \(\mathsf{p}\) is undefined, but the coordinates \({\mathbf m}_i\) of \(\mathsf{m}_i\) can be parametrized as follows for \(i=1,\ldots , 4\):

    $$\begin{aligned} {\mathbf m}_i= {\mathbf n}+ (a_i-a_r){\mathbf d} \quad \text {with}\quad a_1=0,\,\, a_2=a_r+ia_c,\,\, a_3=a_r-ia_c \end{aligned}$$
    (2)

    where \(a_r,a_c\in {\mathbb R}\) and \(a_c\ne 0\) holds. \(\mathsf{m}_5\) is the ideal point in direction of the unit-vector \({\mathbf d}=(d_1,d_2,d_3)^T\), which obtains the rational homogeneous parametrization of the unit-sphere, i.e.

    $$\begin{aligned} d_1=\tfrac{2h_0h_1}{h_0^2+h_1^2+h_2^2}, \quad d_2=\tfrac{2h_0h_2}{h_0^2+h_1^2+h_2^2}, \quad d_3=\tfrac{h_1^2+h_2^2-h_0^2}{h_0^2+h_1^2+h_2^2}. \end{aligned}$$
    (3)

Now we are looking for the point \({\mathbf n}=(n_1,n_2,n_3)^T\) and the direction \((h_0:h_1:h_2)\) in a way that for the self-motion of the pentapod \(e_0=f_0=0\) holds. We can discuss Type 1 and Type 2 at the same time, just having in mind that \(a_4\ne 0\ne C\) has to hold for Type 1 and \(a_4=0=C\) for Type 2 (according to [1]).

By setting \({\mathbf r}_i:=(r_{i1},r_{i2},r_{i3})^T\) for \(i=1,2,3\) the Darboux and Mannheim constraints with respect to \(\mathscr {F}_0^{\prime }\) and \(\mathscr {F}^{\prime }\) can be written as:

$$\begin{aligned} \varOmega _2:\,&(s_1+{\mathbf r}_1{\mathbf m}_2)-i(s_2+{\mathbf r}_2{\mathbf m}_2)-p_2N=0,&\quad \varOmega _4:\,&(s_3+{\mathbf r}_3{\mathbf m}_4)-p_4N=0, \end{aligned}$$
(4)
$$\begin{aligned} \varOmega _3:\,&(s_1+{\mathbf r}_1{\mathbf m}_3)+i(s_2+{\mathbf r}_2{\mathbf m}_3)-p_3N=0,&\quad \varPi _5:\,&({\mathbf R}{\mathbf d})({\mathbf s}+{\mathbf R}{\mathbf p}_5)N^{-1}=0, \end{aligned}$$
(5)

with \({\mathbf p}_5={\mathbf n}+(p_5-a_r){\mathbf d}\), which is the coordinate vector of the intersection point of the Mannheim plane and \(\mathsf{p}\) with respect to \(\mathscr {F}^{\prime }\). Moreover \((p_j,0,0)^T\) for \(j=2,3\) (resp. \((0,0,p_4)^T\)) are the coordinates of the intersection point of the Darboux plane and the x-axis (resp. z-axis) of \(\mathscr {F}_0^{\prime }\).

Remark 1

As from the Mannheim constraint \(\varPi _5\) of Eq. (5) the factor N cancels out, all four constraints \(\varOmega _2,\varOmega _3,\varOmega _4,\varPi _5\) are homogeneous quadratic in the Study parameters and especially linear in \(f_0,\ldots ,f_3\). \(\diamond \)

According to [1, Theorems 13 and 14] the leg-parameters \(p_2,\ldots ,p_5,R_1\) have to fulfill the following necessary and sufficient conditions for the self-mobility (over \({\mathbb C}\)) of a linear pentapod of Type 1 and Type 2, respectively:

$$\begin{aligned}&p_2=\tfrac{Aa_3v}{(a_3-a_4)^2},\qquad p_3=\tfrac{Aa_2v}{(a_2-a_4)^2},\qquad p_4=-\tfrac{Ca_4v}{(a_2-a_4)(a_3-a_4)}, \end{aligned}$$
(6)
$$\begin{aligned}&\quad (a_2-a_4)^2(a_3-a_4)^2\left[ 2wp_5-vR_1^2-(2w-va_4)a_4 \right] + vw^2(A^2+C^2)=0, \end{aligned}$$
(7)

with \(v:=a_2+a_3-2a_4\) and \(w:=a_2a_3-a_4^2\). Therefore if we set \(p_2,p_3,p_4\) as given in Eq. (6) then only one condition in \(p_5\) and \(R_1\) remains in Eq. (7). Therefore these pentapods have a 1-dimensional set of self-motions.

Theorem 1

Each self-motion of a linear pentapod of Type 1 and Type 2 can be generated by a 1-dimensional set of line-symmetric motions. For the special case \(p_5=a_4=a_r\) this set is even 2-dimensional.

Proof

W.l.o.g. we can set \(e_0=0\) as any two directions \({\mathbf d}\) of \(\mathsf{p}\) can be transformed into each other by a half-turn about their enclosed bisecting line. Note that this line is not uniquely determined if and only if the two directions are antipodal.

W.l.o.g. we can solve \(\Psi ,\varOmega _2,\varOmega _3,\varOmega _4\) for \(f_0,f_1,f_2,f_3\) and plug the obtained expressions into \(\varPi _5\), which yields in the numerator a homogeneous quartic polynomial G[1563] in \(e_1,e_2,e_3\), where the number in the brackets gives the number of terms. Moreover the numerator of the obtained expression for \(f_0\) is denoted by F[600], which is a homogeneous cubic polynomial in

\(e_1,e_2,e_3\).

General Case \((v\ne 0)\) : The condition \(G=0\) already expresses the self-motion as G equals \(\varLambda _1\) if we solve Eq. (7) for \(R_1\). Moreover \(F=0\) has to hold if the self-motion of the line \(\mathsf{p}\) can be generated by a line-symmetric motion. As for any solution \((e_1:e_2:e_3)\) of \(F=0\) also \(G=0\) has to hold, G has to split into F and a homogeneous linear factor L in \(e_1,e_2,e_3\).

Now \(L=0\) cannot correspond to a self-motion of the linear pentapod, but has to arise from the ambiguity in representing a direction of \(\mathsf{p}\) mentioned at the beginning of the proof. This can be argued indirectly as follows:

Assumed \(L=0\) implies a self-motion, then it has to be a Schönflies motion (with a certain direction \(\mathsf{v}\) of the rotation axis) due to \(e_0=0\). As under such a motion the angle enclosed by \(\mathsf{v}\) and \(\mathsf{p}\) remains constantFootnote 5 the ideal point \(\mathsf{U}\) of \(\mathsf{p}\) has to be mapped by \(\sigma \) to the ideal point \(\mathsf{V}\) of \(\mathsf{v}\). This implies that \(\mathsf{V}\) has to coincide with \(\mathsf{W}\), which can only be the case for pentapods of Type 5; a contradiction.

Therefore there has to exist a pose of \(\mathsf{p}\) during the self-motion, where it is oppositely oriented with respect to the fixed frame and moving frame, respectively. As a consequence we can set \(L=d_1e_1+d_2e_2+d_3e_3\) which yields the ansatz \(\varDelta :\) \(\lambda LF-G=0\). The resulting set of four equations arising from the coefficients of \(e_1^3e_2\), \(e_1^3e_3\), \(e_1e_3^3\) and \(e_2e_3^3\) of \(\varDelta \) has the unique solution:

$$\begin{aligned} n_1=a_cd_2,\quad n_2=-a_cd_1, \quad n_3=(a_r-a_4)d_3, \quad \lambda =2(h_0^2+h_1^2+h_2^2). \end{aligned}$$
(8)

Now \(\varDelta \) splits up into \((e_1^2+e_2^2+e_3^2)^2(h_0^2+h_1^2+h_2^2)H[177]\), where H is homogeneous of degree 4 in \(h_0,h_1,h_2\). For more details on \(H=0\) please see Remark 3, which is given right after this proof.

Remark 2

Note that all self-motions of the general case can be parametrized as the resultant of G and the normalizing condition \(N-1\) with respect to \(e_i\) yields a polynomial, which is only quadratic in \(e_j\) for pairwise distinct \(i, j\in \left\{ 1,2\right\} \).\(\diamond \)

Special Case \((v= 0)\) : If \(v=0\) holds, we cannot solve Eq. (7) for \(R_1\). The conditions \(v=0\) and Eq. (7) imply \(p_5=a_4=a_r\). Now G is fulfilled identically and the self-motion is given by \(\varLambda _1=0\), which is of degree 4 in \(e_1,e_2,e_3\). Moreover for this special case \(F=0\) already holds for \({\mathbf n}\) given in Eq. (8). Therefore any direction \((h_0:h_1:h_2)\) for \(\mathsf{p}\) can be chosen in order to fix the line-symmetric motion.    \(\square \)

Remark 3

\(H=0\) represents a planar quartic curve, which can be verified to be entirely circular. Moreover \(H=0\) can be solved linearly for \(p_5\). The corresponding graph is illustrated in Fig. 1.

If we reparametrize the \(h_0h_1h_2\)-plane in terms of homogenized polar coordinates by:

$$\begin{aligned} h_0=(\tau _1^2+\tau _0^2)\rho _0,\qquad h_1=(\tau _1^2-\tau _0^2)\rho _1,\qquad h_2=2\tau _0\tau _1\rho _1, \end{aligned}$$
(9)

where \((\tau _0,\tau _1)\ne (0,0)\ne (\rho _0,\rho _1)\) and \(\tau _0,\tau _1,\rho _0,\rho _1\in {\mathbb R}\) hold, then H factors into \((\tau _0^2+\tau _1^2)^3(H_2\tau _1^2 + H_1\tau _0\tau _1 +H_0\tau _0^2)\) with

(10)

Therefore this equation can be solved quadratically for the homogeneous parameter \(\tau _0:\tau _1\). Note that the value \(p_5\) is fixed during a self-motion.                                            \(\diamond \)

Fig. 1
figure 1

For a type 1 pentapod with self-motion given by the parameters \(a_4=2\), \(A=-1\), \(C=-5\), \(a_r=7\) and \(a_c=4\), the graph of \(p_5\) in dependency of \(h_1\) and \(h_2\) with \(h_0=1\) is displayed in the axonometric view on the left and in the front resp. top view on the right side. The highlighted point at height 6 corresponds to the values \(h_1=-\tfrac{489262}{226525}+\tfrac{488}{226525}\sqrt{675091}\) and \(h_2=\tfrac{535336}{226525}+\tfrac{446}{226525}\sqrt{675091}\)

3 On the Reality of Type 1 and Type 2 Self-motions

A similar computation to [1, Example 1] shows that for any real point \(\mathsf{p}_t\in \mathsf{p}\) with \(t\in {\mathbb R}\) and coordinate vector \({\mathbf p}_t={\mathbf n}+ (t-a_r){\mathbf d}\) with respect to \(\mathscr {F}^{\prime }\) the corresponding real point \(\mathsf{P}_t\in \mathsf{P}\) has the following coordinate vector \({\mathbf P}_t\) with respect to \(\mathscr {F}_0^{\prime }\):

$$\begin{aligned} {\mathbf P}_t=\left( \tfrac{A(a_r^2+a_c^2-ta_r)}{(t-a_r)^2+a_c^2}, -\tfrac{Aa_ct}{(t-a_r)^2+a_c^2}, \tfrac{Ca_4}{a_4-t}\right) ^T. \end{aligned}$$
(11)

As \(L=0\) corresponds with one configuration of the self-motion we can compute the locus \(\mathscr {E}_t\) of \(\mathsf{p}_t\) with respect to \(\mathscr {F}_0^{\prime }\) under the 1-parametric set of self-motions by the variation of \((h_0:h_1:h_2)\) within \(L=0\). Moreover due to the mentioned ambiguity we can select an arbitrary solution \((e_0:e_1:e_2)\) for \(L=0\) fulfilling the normalization condition \(N=1\); e.g.:

$$\begin{aligned} e_1=\tfrac{h_2}{\sqrt{h_1^2+h_2^2}}, \quad e_2=-\tfrac{h_1}{\sqrt{h_1^2+h_2^2}} \quad \text {and} \quad e_3=0. \end{aligned}$$
(12)

Now the computation of \({\mathbf R}{\mathbf p}_t+{\mathbf s}\) yields a rational quadratic parametrization of \(\mathscr {E}_t\) in dependency of \((h_0:h_1:h_2)\).

Note that this approach also includes the special case \((v= 0)\) as there always exists a value for \(R_1^2\) (in dependency of \((h_0:h_1:h_2)\)) in a way that \(\varLambda _1=0\) holds.

For \(t\ne a_4\) all \(\mathscr {E}_t\) are ellipsoids of rotation (see Fig. 2a), which have the same center point \(\mathsf{C}\) and axis of rotation \(\mathsf{c}\). In detail, \(\mathsf{C}\) is the point of the straight cubic circle (11) for the value \(t=c\) with \(c:=\tfrac{a_4^2-a_c^2-a_r^2}{2(a_4-a_r)}\) (for \(a_4=a_r\) we get \(c=\infty \) thus \(\mathsf{p}_{\infty }=\mathsf{U}=\mathsf{m}_5\) holds, which implies \(\mathsf{C}=\mathsf{M}_5\)) and \(\mathsf{c}\) is parallel to the z-axis of \(\mathscr {F}_0^{\prime }\). Moreover the vertices on \(\mathsf{c}\) have distance \(|a_4-t|\) from \(\mathsf{C}\) and the squared radius of the equator circle equals \((a_r-t)^2+a_c^2\). Note that for \(a_4\ne a_r\) the only sphere within the described set of ellipsoids is \(\mathscr {E}_c\). For \(a_4= a_r\) no such sphere exists.

\(\mathscr {E}_{a_4}\) is a circular disc in the Darboux plane \(z=p_4\) (w.r.t. \(\mathscr {F}_0^{\prime }\)) centered in \(\mathsf{C}\).

Remark 4

The existence of these ellipsoids was already known to Duporcq [7, Sect. 9], who used them to show that the spherical trajectories are algebraic curves of degree 4 (intersection curve of \(\mathscr {E}_t\) and the sphere \(\varPhi _t\) centered in \(\mathsf{P}_t\) illustrated in Fig. 2b). \(\diamond \)

Fig. 2
figure 2

Type 1 pentapod with self-motion given by \(a_4=2\), \(A=-1\), \(C=-5\), \(a_r=7\) and \(a_c=4\). a The loci \(\mathscr {E}_{a_4}\), \(\mathscr {E}_c\) and \(\mathscr {E}_{t}\) with \(t=\tfrac{69}{20}\) are sliced (along the not drawn axis of rotation \(\mathsf{c}\)) in order to visualize their positioning with respect to the cubic \(\mathsf{P}\) on which the points \(\mathsf{P}_{\infty }=\sigma (\mathsf{U})\), \(\mathsf{P}_c=\mathsf{C}\) and \(\mathsf{P}_t\) are highlighted. Note that \(\mathsf{P}_{a_4}=\mathsf{W}\) is the real ideal point of \(\mathsf{P}\). b By setting \(p_5=6\) a one-parametric self-motion \(\mu \) is fixed. The trajectory of \(\mathsf{p}_t\) under \(\mu \) is illustrated as the intersection curve of \(\mathscr {E}_t\) and the sphere \(\varPhi _t\) centered in \(\mathsf{P}_t\). c A strip of the basic surface of \(\mu \) is illustrated for the value highlighted in Fig. 1. In addition \(\mathsf{P}\) and \(\overline{\mathsf{p}}\) are visualized, where the latter denotes the pose of \(\mathsf{p}\) such that its half-turns about the generators of the basic surface yield the self-motion \(\mu \). d Krames’s construction is illustrated with respect to the generator \(\mathsf{g}\) of the basic surface: As \(\mathsf{P}_{a_4}\) (resp. \(\overline{\mathsf{p}}_{\infty }\)) is the real ideal point of \(\mathsf{P}\) (resp. \(\overline{\mathsf{p}}\)), the trajectory of \(\mathsf{p}_{a_4}\) (resp. \(\overline{\mathsf{P}}_{\infty }\)) under \(\mu \) is planar. The (Mannheim) plane \(\in \varSigma \), which contains the point \(\mathsf{P}_{\infty }\) (resp. \(\overline{\mathsf{p}}_{a_4}\)) and is orthogonal to the direction of the real ideal point \(\mathsf{p}_{\infty }\) (resp. \(\overline{\mathsf{P}}_{a_4}\)) of \(\mathsf{p}\) (resp. \(\overline{\mathsf{P}}\)) in the displayed pose, slides through the point \(\mathsf{P}_{\infty }\) (resp. \(\overline{\mathsf{p}}_{a_4}\)) during the complete motion \(\mu \)

Based on this geometric property, recovered by line-symmetric motions, we can formulate the condition for the self-motion to be real as follows:

  • \(w\ne 0\): We can reduce the problem to a planar one by intersecting the plane spanned by \(\mathsf{P}_0=\mathsf{M}_1\) and \(\mathsf{c}\) with \(\mathscr {E}_0\) and the sphere with radius \(R_1\) centered in \(\mathsf{P}_0\). Now there exists an interval \(I_0=]I_-,I_+[\) such that for \(R_1\in I_0\) the two resulting conics have at least two distinct real intersection points. It is well known (e.g. [14]) that the computation of the limits \(I_-\) and \(I_+\) of the reality interval \(I_0\) leads across an algebraic problem of degree 4 (explicitly solvable). Thus for a real self-motion we have to choose \(R_1\in I_0\) and solve Eq. (7) for \(p_5\).

  • \(w=0\): Now \(\mathsf{P}_0\) coincides with \(\mathsf{C}\) and the interval collapses to the single value \(R_1=|a_4|\), which can be seen from Eq. (7). Moreover \(p_5\) can be chosen arbitrarily.

These considerations also show that any pentapod of Type 1 and 2 has real self-motions if the leg-parameters are chosen properly. Note that this is e.g. not the case for some designs of Type 5 pentapods described in [1, Sect. 6], where it was also proven that pentapods with self-motions have a quartically solvable direct kinematics. It is possible to use this advantage (closed form solution) of pentapods with self-motions without any risk,Footnote 6 by designing linear pentapods of Type 1 and Type 2, which are guaranteed free of self-motions within their workspace.

A sufficient condition for that is that (at least) for one of the five legs \(\mathsf{p}_t\mathsf{P}_t\) of the pentapod the corresponding reality interval \(I_t\) is disjoint with the interval of the maximal and minimal leg length implied by the mechanical realization. This condition for a self-motion free workspace gets especially simple if \(\mathsf{p}_c\mathsf{P}_c\) is this leg.

Remark 5

Due to limitation of pages, we refer for detailed examples to the paper’s corresponding arXiv version [13], which also show that for the general case \((v\ne 0)\) the basic surface is of degree 5 (see Fig. 2c) and that a general point has a trajectory of degree 6 under the corresponding line-symmetric motion.Footnote 7 Note that the latter also holds for a general point of the cubic \(\overline{\mathsf{P}}\) explained in the next section.                                      \(\diamond \)

4 Conclusion and Open Problem

Krames [4, p. 416] outlined the following construction (see Fig. 2d): Assume that \(\mathsf{p}\) is in an arbitrary pose of the self-motion \(\mu \) with respect to \(\mathsf{P}\), where \(\mathsf{g}\) denotes the generator of the basic surface, which corresponds to this pose. Moreover \(\overline{\mathsf{p}}\) and \(\overline{\mathsf{P}}\) are obtained by the reflexion of \(\mathsf{p}\) and \(\mathsf{P}\), respectively, with respect to \(\mathsf{g}\), where \(\overline{\mathsf{p}}\) belongs to the fixed system \(\varSigma _0\) and \(\overline{\mathsf{P}}\) to the moving system \(\varSigma \). Then under the self-motion \(\mu \) also the points of \(\overline{\mathsf{P}}\) are located on spheres with centers on the line \(\overline{\mathsf{p}}\).

We can apply this construction for each line-symmetric motion of Theorem 1, which yields new solutions for the Borel Bricard problem, with the exception of one special case where \(\mathsf{W}\in \overline{\mathsf{p}}\) holds (i.e. \(h_1=h_2=0\) or \(h_0=0\)), which was already given by Borel in [2, Case Fa4]. Moreover for this case Borel noted that beside \(\mathsf{p}\) and \(\overline{\mathsf{P}}\) only two imaginary planar cubic curves (\(\in \) isotropic planes through \(\mathsf{p}\)) run on spheres. The example of [13] shows that this also holds true for the general case.

Thus the problem remains to determine all line-symmetric motions of Theorem 1 where additional real points (beside those of \(\mathsf{p}\) and \(\overline{\mathsf{P}}\)) run on spheres. Until now the only known examples with this property are the BB-II motions (cf. Sect. 1.3).