1. Introduction
A solution to the motion planning problem in robotics is given by a motion planning algorithm, which is a function that takes as inputs a pair of configurations of a mechanical system and produces a continuous path connecting these two configurations as an output. More specifically, consider X to be the configuration space of a mechanical system and XI to be the free path space of X with a compact open topology. The free path space fibration $\pi:X^I\to X\times X$ is defined by $\pi(\gamma)=(\gamma(0),\gamma(1))$. A motion planning algorithm is then a section of a free path space fibration.
Farber [Reference Farber22] introduced the concept of topological complexity to analyse the difficulty of computing a motion planning algorithm for the configuration space X of a mechanical system. The topological complexity of a space X, denoted by $\mathrm{TC}(X)$, is defined as the smallest positive integer k for which X × X can be covered by open sets $\{U_1,\dots, U_k\}$ such that each Ui admits a continuous local section of π. Farber showed that a motion planning algorithm in the configuration space cannot be continuous unless it is contractible. Therefore, for non-contractible spaces, the topological complexity is at least two. The numerical invariant topological complexity has been extensively studied over the past two decades.
To generalize topological complexity, Rudyak introduced its sequential analogue in [Reference Rudyak38]. This models the motion planning problem for robots that need to visit several places in between their initial and final stages while performing their tasks. We briefly recall the definition. Consider the generalized free path space fibration $\pi_n: X^I\to X^n$ of path-connected spaces X
The sequential topological complexity of X is the smallest positive integer k for which Xn is covered by open sets $\{U_1,\dots, U_k\}$, such that each Ui admits a continuous local section of πn. Note that by definition, it follows that $\mathrm{TC}_2(X)=\mathrm{TC}(X)$. Several properties of topological complexity have been generalized and introduced in their symmetrized version in [Reference Basabe, González, Rudyak and Tamaki2].
There is an old invariant called LS-category, a close relative of sequential topological complexity, which was introduced by Lusternik and Schnirelmann in [Reference Lyusternik and Šnirelmann34]. The LS-category of a space X is denoted by $\mathrm{cat}(X)$, which is the least number of open subsets which cover X such that the inclusion on each open set is nullhomotopic. The authors of [Reference Basabe, González, Rudyak and Tamaki2] prove the following famous inequality
The notion of parametrized topological complexity was introduced by Cohen, Farber and Weinberger in [Reference Cohen, Farber and Weinberger8], and the corresponding sequential analogue was introduced by Farber and Paul in [Reference Farber and Kumar Paul23]. These parametrized motion planning algorithms offer enhanced universality and flexibility, enabling effective operation across various scenarios incorporating external conditions. These conditions are considered parameters and integral components of the algorithm’s input. A parametrized motion planning algorithm takes as inputs a pair of configurations living under the same external conditions and produces a continuous motion of the system, which remains constant under the external conditions.
We now define sequential parametrized topological complexity in brief. For a fibration $p: E \to B$, consider the fibre product $E^n_B$, the space of all n-tuples of points in E all of which lie in a common fibre of p. The space $E^I_B$ denotes the space of all paths in E with image in a single fibre. Note that the fibration defined in (1) restricts to the subspace $E^I_B$ of EI. We denote this fibration by $\Pi_n:E^I_B\to E^n_B$. The sequential parametrized topological complexity of a fibration $p : E \to B$ denoted by $\mathrm{TC}_n[p:E\to B]$ is the smallest integer k such that there is an open cover $\{U_1,\dots, U_k\}$ of $E^n_B$, where each open set Ui admits a continuous section of $\Pi_n$. Note that $\mathrm{TC}_2[p:E\to B]$ is known as the parametrized topological complexity of Cohen, Farber and Weinberger. The reader is referred to [Reference Cohen, Farber and Weinberger8, Reference Cohen, Farber and Weinberger9, Reference Farber and Kumar Paul23, Reference Farber and Weinberger27] for several interesting results related to (sequential) parametrized topological complexity. We also mention that the notion of parametrized topological complexity of fibrations is extended to fibrewise spaces in [Reference García-Calcines28] by García-Calcines.
The invariants we have discussed so far are special cases of a more general notion: the Schwarz genus of a fibration. Schwarz introduced and studied this notion in [Reference Švarc42]. Later, Bernstein and Ganea [Reference Berstein and Ganea4] extended and studied this notion for any map. Let’s now define this notion. The sectional category of a map $p:E\to B$, denoted $\mathrm{secat}(p)$, is the smallest integer k for which B can be covered by k open sets $W_1,\dots,W_k$, where each Wi admits a continuous homotopy section of p. If $p: E\to B$ is a fibration, then $\mathrm{secat}(p)$ coincides with another invariant called the Schwarz genus of a fibration, denoted $\mathrm{gen}(p)$ (see [Reference Švarc42]). For example, $\mathrm{TC}_n(X)=\mathrm{secat}(\pi_n)$, $\mathrm{cat}(X)=\mathrm{secat}(i: \{\ast\}\hookrightarrow{} X)$ and $\mathrm{TC}_n[p:E\to B]=\mathrm{secat}(\Pi_n)$. The reader is referred to [Reference James33] for a more general overview of the sectional category.
Generalizing the concept of sectional category, Colman and Grant [Reference Colman and Grant12] first introduced the corresponding equivariant analogue. The equivariant sectional category of a G-map $p:E\to B$ between G-spaces is denoted by $\mathrm{secat}_G(p)$ (see definition 3.1). Colman and Grant extended some of the classical results from [Reference Švarc42] to the equivariant settings. It can be observed that the (generalized) free path space fibration is a G-map (in fact, a G-fibration) (see §3). Colman and Grant introduced the notion of equivariant topological complexity as a consequence of the equivariant sectional category. The sequential analogue of this concept was introduced by Bayeh and Sarkar in [Reference Bayeh and Sarkar3]. They established several interesting relationships between the equivariant category [Reference Marzantowicz35] and sequential equivariant topological complexity. Additionally, they obtained new bounds on its corresponding non-equivariant counterpart.
By a (Serre) G-fibration we will mean a G-map $p : E\to B$ having the G-homotopy lifting property with respect to all G-spaces (G-CW complexes). One can observe that for a G-fibration $p:E\to B$, the corresponding fibre product $E^n_B$ admits a G-action. Moreover the space $E^I_B$ is also G-invariant and the fibration $\Pi_n:E^I_B\to E^n_B$ is a G-map (see §4 for details). Therefore, expecting an equivariant analogue of parametrized topological complexity is natural. In §4, we define the sequential equivariant parametrized topological complexity as the equivariant sectional category of a G-map $\Pi_n$. We denote this newly defined notion by $\mathrm{TC}_{G,n}[p:E\to B]:=\mathrm{secat}_G(\Pi_n)$. This notion generalizes sequential parametrized topological complexity. We generalize several results from the theory of sequential parametrized topological complexity. We define the fibrewise G-homotopy equivalent fibrations (see definition 4.13) and establish the fibrewise G-homotopy invariance of $\mathrm{TC}_{G,n}[p;E\to B]$ (see proposition 4.14). Later, in §4.2, we obtain several bounds on the sequential equivariant parametrized topological complexity involving the equivariant category of fibre product. We also obtain the cohomological lower bound theorem 4.25 and the dimension-connectivity upper bound in theorem 4.27.
2. Sequential parametrized topological complexity
Cohen, Farber and Weinberger introduced the notion of parametrized topological complexity in [Reference Cohen, Farber and Weinberger8] ,[Reference Cohen, Farber and Weinberger9]. This section briefly recalls some basic results related to this notion.
For a Hurewicz fibration $p:E\to B$, consider a subspace of the path space EI, defined as follows:
The fibre product corresponding to $p:E\to B$ is defined by
Define a map $\Pi_n: E^I_B\to E^n_B$ by
It follows from the appendix of [Reference Cohen, Farber and Weinberger9] that $\Pi_n$ is a Hurewicz fibration. A section of $\Pi_n$ is called a sequential parametrized motion planning algorithm.
In their recent paper [Reference Farber and Kumar Paul23], Farber and Paul have established a theory for the sequential parametrized topological complexity, generalizing several results of the original parametrized topological complexity.
Definition 2.1. Let $p:E\to B$ be a fibration. The n-th sequential parametrized topological complexity of p is denoted by $\mathrm{TC}_n[p:E\to B]$, and defined as
It is observed in [Reference Farber and Kumar Paul23] that for a pullback fibration $q:E'\to B'$ along a map $f:B'\to B$ of $p:E\to B$, we have
In particular, for a fibration $p:E\to B$ with fibre F, we have
Various bounds have been established to estimate the parametrized topological complexity in [Reference Cohen, Farber and Weinberger8], and later, these were generalized in the sequential setting in [Reference Farber and Kumar Paul23]. For example, the cohomological lower bound for the parametrized topological complexity is given in [Reference Cohen, Farber and Weinberger8], and its sequential analogue is stated in [Reference Farber and Kumar Paul23]. We state the sequential version of the cohomological lower bound.
Theorem 2.2 ([Reference Farber and Kumar Paul23, proposition 6.3])
Let $p:E\to B$ be a fibration, $\triangle: E\to E^n_B$ be the diagonal map and $\triangle^{\ast}:H^{\ast}(E^n_B;R)\to H^{\ast}(E;R)$ be the corresponding induced homomorphism. Suppose there exists cohomology classes $u_i\in \ker(\triangle^{\ast}) $ for $1\leq i\leq r$ such that their cup product $u_1\cup \dots \cup u_r$ does not vanish. Then
In [Reference Farber and Kumar Paul23], Farber and Paul established a dimension-connectivity upper bound on the sequential parametrized topological complexity, generalizing the base case proved in [Reference Cohen, Farber and Weinberger8, proposition 7.1].
Theorem 2.3 [Reference Farber and Kumar Paul23, proposition 6.1]
Let $p:E\to B$ be a fibration with fibre F such that spaces E, B and F are CW-complexes. Suppose that for $m\geq 0$, F is m-connected. Then
We now present the computations of the sequential parametrized topological complexity for some fibrations. Sarkar and Zvengrowski [Reference Sarkar and Zvengrowski39] introduced a class of topological spaces called generalized projective product spaces. We will now recall the definition of these spaces. Let M and N be CW-complexes with involutions $\tau \colon M \to M$ and $\sigma \colon N \to N$, where σ is fixed-point-free. The following identification space is called generalized projective product space
Note that this class of manifolds includes all projective product spaces [Reference Davis17] and Dold manifolds [Reference Dold18]. The author and Soumen Sarkar studied the LS-category and topological complexity of generalized projective product spaces in [Reference Daundkar and Sarkar15].
Example 2.4.
Observe that there is a fibre bundle
defined by $p([(x,y)])=[y]$, where $N/\left \lt \sigma \right \gt $ is the orbit space of the group $\left \lt \sigma \right \gt $ action on N induced by the involution σ.
It follows from theorem 2.3 that
In particular,
Also, from (3), we have the following lower bound:
Observe that if $N=S^0$ with free involution, then $N/\left \lt \sigma \right \gt $ is a singleton and $X(M,N)\cong M$. Thus, $\mathrm{TC}_n[p:X(M,S^0)\to S^0/\left \lt \sigma \right \gt ]=\mathrm{TC}_n(M)$.
Note that if $N=S^1$ with the antipodal involution and $\mathrm{TC}_n(M)=n\dim(M)+1$, then
One example of spaces with involutions having maximal sequential topological complexity are orientable surfaces of positive genus. An orientable surface of genus $g\geq 1$ can be embedded in $\mathbb{R}^3$ to admit antipodal involution. It was proved in [Reference González, Gutiérrez, Gutiérrez and Lara29] that $\mathrm{TC}_n(\Sigma_g)=n\dim(\Sigma_g)+1=2n+1$.
In some cases, we can now compute the exact value of sequential parametrized topological complexity. Suppose for $d\geq 2$, consider $M=S^d$ with an involution τ obtained by reflecting Sd across the X-axis. Observe that $p:X(S^d,N)\to N/\left \lt \sigma \right \gt $ is a sphere bundle with structure group $\mathbb{Z}_2=\left \lt \tau \right \gt $. One can generalize [Reference Colman and Grant12, example 5.9] to obtain $\mathrm{TC}_{\mathbb{Z}_2,n}(S^d)=n+1$. Then it follows from [Reference Farber and Oprea25, theorem 3.4] that
From (3) we have $\mathrm{TC}_n(S^d)\leq \mathrm{TC}_n[p:X(S^d,N)\to N/\left \lt \sigma \right \gt ]$. Therefore, if d is even, we have $ \mathrm{TC}_n[p:X(S^d,N)\to N/\left \lt \sigma \right \gt ]=n+1$. If d is odd, then $\mathrm{TC}_n[p:X(S^d,N)\to N/\left \lt \sigma \right \gt ]\in\{n, n+1\}$.
Example 2.5.
In [Reference Milnor36], Milnor introduced a class of submanifolds of the products of real and complex projective spaces to define generators for the unoriented cobordism algebra. We define these manifolds as follows: Let r and s be integers such that $0\leq s\leq r$. A real Milnor manifold, denoted by $\mathbb{R}M_{r,s}$, is the submanifold of the product $\mathbb{R}P^r \times \mathbb{R}P^s$ of real projective spaces. It is defined as follows:
The real Milnor manifold $\mathbb{R}M_{r,s}$ is an $(s+r-1)$-dimensional closed, smooth manifold. There is a fibre bundle:
A complex Milnor manifold, denoted by $\mathbb{C} M_{r,s}$, can be defined in a similar manner:
Note that $\mathbb{C} M_{r,s}$ is a closed, smooth manifold of dimension $2(s+r-1)$. Just as in the real case, $\mathbb{C} M_{r,s}$ is the total space of the fiber bundle:
The author and B. Singh [Reference Daundkar and Singh16] studied sequential topological complexity of $\mathbb{R} M_{r,s}$ and $\mathbb{C} M_{r,s}$, obtaining several sharp bounds and computing exact values in various cases.
Note that, using (3) and [Reference Farber and Kumar Paul23, proposition 6.1], we obtain
The topological complexity of projective spaces has been studied in [Reference Farber, Tabachnikov and Yuzvinsky26] and the corresponding sequential analogue is discussed in [Reference Cadavid-Aguilar, González, Gutiérrez, Guzmán-Sáenz and Lara6].
If $r=2^m+1$, then it is known that $\mathrm{TC}_n(\mathbb{R} P^{r-1})=n(r-1)$. Thus, for $r=2^m+1$, we have
Since $\mathbb{C} M_{r,s}$ is simply connected, applying theorem 2.3 gives us
Since $\mathrm{TC}_n(\mathbb{C} P^{r-1})=n(r-1)+1$, using (10) we obtain
Thus, if s = 1, then
3. Equivariant sectional category
The sectional category of a Hurewicz fibration was introduced by Švarc in [Reference Švarc42]. Colman and Grant [Reference Colman and Grant12] generalized this notion and introduced the equivariant sectional category. Let G be a compact Lie group and $E, B$ be G-spaces such that $p\colon E \to B$ is a G-map. The symbol $\simeq_G$ denotes the G-homotopy equivalence. In this paper, when we refer to a group G, we mean a compact Lie group.
Definition 3.1 ([Reference Colman and Grant12, definition 4.1])
Let $p:E\to B$ be a G-map. The equivariant sectional category of p, denoted by $\mathrm{secat}_G(p)$, is the least positive integer k such that there is a G-invariant open cover $\{U_1, \ldots, U_k\}$ of B and G-maps $s_i \colon U_i \to E$ for $i=1, \ldots, k$, such that $p\circ s_i \simeq_G i_{U_i}$, where $i_{U_i} : U_i \hookrightarrow B$ is the inclusion map.
If no such k exists, we say $\mathrm{secat}_G(p)=\infty$. If G is a trivial group, then $\mathrm{secat}_G(p)$ is called the sectional category of p, denoted by $\mathrm{secat}(p)$.
Fadell [Reference Fadell19] introduced the notion of G-equivariant LS-category for G-spaces, and it was later studied by Marzantowicz in [Reference Marzantowicz35], Clapp and Puppe in [Reference Clapp and Puppe7], Colman [Reference Colman11], and Angel, Colman, Grant, and Oprea in [Reference Angel, Colman, Grant and Oprea1]. This homotopy invariant of a G-space X is denoted by $\mathrm{cat}_G(X)$. Before defining this notion, we recall the notion of G-categorical sets. An invariant open set U of a G-space X is called G-categorical if the inclusion $i_{U}:U\hookrightarrow X$ is G-homotopic to the map that takes values in a single orbit.
Definition 3.2. For a G-space X, the G-equivariant category $\mathrm{cat}_G(X)$ is defined as the least positive integer r such that X can be covered by r G-categorical sets.
Let H be a closed subgroup of G and X be a G-space. The H-fixed point set of X is denoted by XH and defined as
Definition 3.3. A G-space X is said to be G-connected if, for any closed subgroup H of G, the H-fixed point set XH is path-connected.
It can be observed that if $x\in X$ is fixed point of a G-action and if X is a G-connected space, then for the inclusion map $i\colon \{x\} \to X$, $\mathrm{secat}_G(i\colon \{x\} \to X) = \mathrm{cat}_G(X)$; see [Reference Colman and Grant12, corollary 4.7]. Additionally, if G is trivial, then $\mathrm{cat}_G(X)=\mathrm{cat}(X)$.
We now recall the notion of sequential (higher) equivariant topological complexity introduced by Bayeh and Sarkar in [Reference Bayeh and Sarkar3]. This notion generalizes the equivariant topological complexity introduced by Colman and Grant in [Reference Colman and Grant12].
Let X be a G-space. Observe that the path space admits G-action via $(g\cdot \gamma)(t):=g\cdot \gamma(t)$. The product Xn is also a G-space with the diagonal G-action. One can check that the fibration (1) is a G-fibration.
Definition 3.4. Let X be a G-space. The sequential equivariant topological complexity is denoted by $\mathrm{TC}_{G,n}(X)$ is the least positive integer k such that the space Xn is covered by k-many G-invariant open sets with each of which having an G-equivariant section of πn. In other words, $\mathrm{TC}_{G,n}(X)=\mathrm{secat}_G(\pi_n)$.
In the next section, we introduce sequential equivariant parametrized topological complexity, a special case of which is sequential equivariant topological complexity.
Several results from [Reference Colman and Grant12] are generalized by Bayeh and Sarkar in [Reference Bayeh and Sarkar3]. The following result is a sequential analogue of [Reference Colman and Grant12, corollary 5.8]. We couldn’t find this result in Bayeh and Sarkar’s paper [Reference Bayeh and Sarkar3]. Since the proof is fairly straightforward, we only include the statement.
Proposition 3.5. Let X be a G-space and H be a stabilizer of some $a\in X$. Then
The following result is also missed out in [Reference Bayeh and Sarkar3], which is a sequential analogue of [Reference Colman and Grant12, proposition 5.12]. Since the proof is almost identical to that of [Reference Colman and Grant12, proposition 5.12], we omit it here.
Proposition 3.6. Let A be a topological group and G acts on A via topological group homomorphisms such that A is G-connected. Then $\mathrm{TC}_{G,n}(A)=\mathrm{cat}_G(A^{n-1})$.
4. Equivariant parametrized topological complexity
This section introduces the concept of sequential equivariant parametrized topological complexity and examines its properties. We define fibrewise G-homotopy equivalent fibrations and prove one of the important results of this section which states that the sequential G-equivariant parametrized topological complexity of fibrewise G-homotopy equivalent fibrations coincides. Then, we obtain various bounds on sequential equivariant parametrized topological complexity, which generalize bounds on the equivariant topological complexity obtained by Colman and Grant in [Reference Colman and Grant12].
Let $p:E\to B$ be a G-fibration, and let $E^n_B$ be the corresponding fibre product. Define a G-action on $E^n_B$ as $g\cdot(e_1,\dots,e_n)=(ge_1,\dots,ge_n)$. Since p is G-map, we have $p(ge_i)=gp(g_i)=gp(e_j)=p(ge_j)$. Thus, this action is well-defined. The G-action on $E^I_B$ is defined as follows. Let $\gamma\in E^I_B$. Then define $(g\cdot \gamma)(t)=g\gamma(t)$. Note that $p(g\cdot \gamma)=gp(\gamma)$ and since $p(\gamma)$ is constant, $p(g\cdot \gamma)$ is also constant. Therefore, $E^I_B$ also admits a G-action. Observe that for $g\in G$ we have:
Thus, $\Pi_n$ is G-equivariant.
Definition 4.1. The sequential equivariant parametrized topological complexity of a G-fibration $p:E\to B$, denoted by $\mathrm{TC}_{G,n}[p:E\to B]$, is defined as
We recall that Colman and Grant observed in [Reference Colman and Grant12] that the fibration πn is a G-fibration (also it follows from [Reference Grant30, proposition 2.5]). This can be proved using an equivariant analogue of [Reference Spanier40, theorem 2.8.2]. Cohen, Farber and Weinberger [Reference Cohen, Farber and Weinberger8, Appendix] have shown that the map $\Pi_n$ is a fibration. We want to prove an equivariant analogue of their result.
In [Reference İlhan31], an equivariant analogue of lifting function was defined and equivariant analogue of [Reference Spanier40, theorem 2.7.8] was proved (see [Reference İlhan31, § 2.1 and proposition 2.1.3]). We also refer the reader to [Reference Grant30, Reference İlhan32, Reference Waner43 § 2].
Let $p:E\to B$ be a G-fibration. Let X be a topological space with a trivial G-action. Consider the space $E^X_B=\{f:X\to E \mid p\circ f \text{is constant}\}$ with the compact open topology, on which the G-action is defined as $(gf)(x)=g(f(x))$. Let (K, L) be a pair of CW-complexes with trivial G-action, and $\Pi:E^K_B\to E^L_B$ be the restriction map defined by $\Pi(f)=f|_{L}$. Note that Π is a G map.
Proposition 4.2. The map Π is a G-fibration.
Proof. Since $p:E\to B$ is a G-fibration, it follow that (see [Reference İlhan31, proposition 2.1.3] or [Reference Grant30, lemma 2.2]), there exist a G-lifting function $\lambda: \bar{B}\to E^I$, where $\bar{B}=\{(e,\gamma)\in E\times B^I \mid p(e)=\gamma(0)\}$, such that $p\circ \lambda(e,\gamma)=\gamma$ and $\lambda(e,\gamma)(0)=e$. Consider the trivial action of G on L × I. An element h of $(E^L_B)^I$ can be considered as a map $h:L\times I\to E$ such that for any $t\in I$, the image $h(L\times \{t\})\subseteq E$ lies in a single fibre of p. Note that, we need to construct a G-lifting function $\Lambda: \bar{E}^L_B\to (E^K_B)^I$ for Π, where
That is, Λ should satisfy the conditions: $\Lambda(f,h)(x,t)=h(x,t)$ for all $x\in L$ and $t\in I$, and $\Lambda(f,h)(x,0)=f(x)$ for $x\in K$. Note that $\bar{E}^L_B$ has the diagonal G-action.
Let $h\in (E^L_B)^I$. Define a path ωh in B by $\omega_h(t)=p\circ h(x,t)$ for $x\in L$. Now define $\tilde{\Lambda}:\bar{E}^L_B\to (E^K_B)^I$ by $\tilde{\Lambda}(f,h)(x,t)=\lambda(f(x), \omega_h)(t)$. Note that for $g\in G$, we have $\tilde{\Lambda}(gf, gh)(x, t) =\lambda (gf(x), \omega_{gh})(t) = \lambda(gf(x), g\omega_h)(t) = g\lambda(f(x), \omega_h)(t)$. Thus, $\tilde{\Lambda}$ is a G-map. Observe that, $\tilde{\Lambda}(f,h)(x,0)=f(x)$ for $x\in K$. Thus $\tilde{\Lambda}$ satisfies the second condition of a G-lifting function. However, the condition $\tilde{\Lambda}(f,h)(x,t)=h(x,t)$ for all $x\in L$ is not guaranteed. Thus, we need a modification to $\tilde{\Lambda}$.
Define $\alpha_h:L\times I\times I\to E$ by
where $\omega_h^{[\tau,1]}$ is a path defined by $\omega_h^{[\tau,1]}(s)=\omega_h(\tau+(1-\tau)s)$ for $s\in I$. Note that for $x\in L$, we have $p\circ \alpha_h(x,\tau,t)=\omega_h(t)$, $\alpha_h(x,0,t)=\lambda(h(x,0),\omega_h)(t)=\tilde{\Lambda}(f,h)(x,t)$, and $\alpha_h(x,1,t)=h(x,t)$.
We will now define
where $\rho:K\times I\to K\times \{0\}\cup L\times I$ is a retraction. Here we have treated $\tilde{\Lambda}(f,g)$ as a function from $K\times \{0\}\times I\to E$. Since for $g\in G$, we have $g\alpha_h=\alpha_{gh}$ and $\tilde{\Lambda}$ is a G-map, we conclude that Λ is a G-map. One can check that $\Lambda(f,h)(x,t)=h(x,t)$ for all $x\in L$ and $\Lambda(f,h)(x,0)=f(x)$ for $x\in K$. Thus, Λ is a G-lifting function for Π.
Corollary 4.3. The G-map $\Pi_n:E^I_B\to E^n_B$ is a G-fibration.
Proof. The proof follows from proposition 4.2, by considering K = I as a CW-complex with its subcomplex $L=\{0,\frac{1}{n-1},\dots,\frac{n-2}{n-1},1\}$.
The following proposition is a straightforward application of definition 4.1.
(1) Let $p:E\to B$ be a fibration, and let G act trivially on E and B. Then $\mathrm{TC}_{G,n}[p:E\to B]=\mathrm{TC}_n[p:E\to B]$.
(2) If $B= \{\ast\}$, then $\mathrm{TC}_{G,n}[p:E\to B]=\mathrm{TC}_{G,n}(E)$.
(3) Suppose a G-fibration $p:E\to B$ is trivial. Then $\mathrm{TC}_{G,n}[p:E\to B]=\mathrm{TC}_{G,n}(F)$.
Proof. Parts (1) and (2) are obvious. We prove part (3) here. Let F and B be G-spaces and let $p: F\times B (=E)\to B$ be a trivial G-fibration. Then we have $E^n_B=F^n\times B$, $E^I_B=F^I\times B$ and $\Pi_n: E^I_B\to E^n_B$ becomes $\pi_n\times id: F^I\times B\to F^n\times B$. It is easy to see that $\mathrm{secat}_G(\pi_n\times id)=\mathrm{secat}_G(\pi_n)$. Thus we have
Remark 4.5. Note that if G acts trivially on F and B, then the conclusion of (8) follows from [Reference Farber and Kumar Paul23, example 3.2] and the n = 2 case from [Reference Cohen, Farber and Weinberger8, example 4.2].
In the following result, we establish an equivariant analogue of [Reference Daundkar14, lemma 2.4]. We also show that if the fixed point set of base space of a G-fibration is non-empty, then the sequential equivariant topological complexity of fibre is dominated by the sequential equivariant parametrized topological complexity.
Proposition 4.6. Let $p:E\to B$ be a G-fibration and let $\tilde{p}: \tilde{E}\to \tilde{B}$ be a pullback fibration corresponding to a G-map $\tilde{B}\to B$. Then,
Moreover, if $E^G\neq \emptyset$, then F admits G-action and
Proof. We consider the diagram of a pullback of G-fibrations
Note that f ʹ induces a G-map $\tilde{f'}: \tilde{E}^n_{\tilde{B}}\to E^n_B$ defined by $\tilde{f'}(e_1,\dots,e_n)=(f'(e_1),\dots,f'(e_n))$. The map $\tilde{f'}$ is well defined because of the commutativity of the above diagram. One can observe that the pullback of the G-fibration $\Pi_n: E^I_B\to E^n_B$ along $\tilde{f'}$ is isomorphic as a G-fibration to $\tilde{\Pi}_n:\tilde{E}^I_{\tilde{B}}\to \tilde{E}^n_{\tilde{B}}$. Thus,
Since $E^G\neq \emptyset$, we have $B^G\neq \emptyset$. Then it is easy to see that, for $e\in E^G$, the fibre $F:=p^{-1}(p(e))$ admits a G-action. Then (10) follows from (9) by taking the pullback along the inclusion $\{b\}\hookrightarrow B$.
In the following result, we show that, under certain conditions, having a continuous G-equivariant parametrized motion planning algorithm for a G-fibration is equivalent to the fibre being G-contractible.
Proposition 4.7. Let $p:E\to B$ be a G-fibration with $\mathrm{TC}_{G,n}[p:E\to B]=1$. Suppose there exists $e\in E^G\neq \emptyset$ such that fibre $F:=p^{-1}(p(e))$ is G-connected. Then F is G-contractible. Conversely, if F is G-contractible and $E^n_B$ is a G-CW-complex, then $\mathrm{TC}_{G,n}[p:E\to B]=1$.
Proof. We have $\mathrm{TC}_{G,n}(F)=1$ from (10). Then, from proposition 3.5, we get that $\mathrm{cat}_G(F^{n-1})=1$. This forces $\mathrm{cat}_G(F)=1$. That is, F is G-contractible.
Note that the fibre of $\Pi_n$ is $(\Omega F)^{n-1}$. Since $e\in E^G\neq \emptyset$, the space $(\Omega F)^{n-1}$ is admits a G-action. Additionally, $(\Omega F)^{n-1}$ is G-contractible because F is G-contractible. Then, from the equivariant obstruction theory, there exist a continuous section of $\Pi_n:E^I_B\to E^n_B$. This proves the converse.
In [Reference tom Dieck41], tom Dieck introduced the notion of principal $(G,\alpha, A)$-bundles, where G and A are topological groups and G acts on A via a continuous homomorphism $\alpha:G\to \mathrm{Aut}(A)$. For more details, see [Reference Murayama and Shimakawa37]. We now define these bundles.
Definition 4.8 ([Reference Murayama and Shimakawa37, definition 1.1])
Let E and B be left G-spaces. A locally trivial principal A-bundle $p:E\to B$ is called a principal $(G,\alpha,A)$-bundle if p is G-equivariant and $g( x\cdot a)= gx\cdot ga$ for every $g\in G$, $a\in A$ and $x\in E$.
For such bundles we establish a relation between sequential equivariant parametrized topological complexity and the G-equivariant category of $A^{n-1}$. We observe that the latter coincides with the sequential G-equivariant topological complexity of A. The following result generalizes [Reference Farber and Kumar Paul23, proposition 3.3].
Theorem 4.9 Let $p:E\to B$ be a principal $(G,\alpha, A)$-bundle such that A is G-connected. Then,
Proof. Note that G acts on A via a continuous homomorphism $\alpha:G\to \mathrm{Aut}(A)$. Therefore, for the identity $e\in A$, we have $g\cdot e=\alpha(g)(e)=e$. Thus, e is a fixed point of the G-action on A. This allows the path space $P_0(A)=\{\gamma\in A^I \mid \gamma(0)=e\}$ to admit a G-action. Moreover, we have assumed that A is G-connected. Thus, it follows from [Reference Colman and Grant12, corollary 4.7] that $\mathrm{secat}_G(q:P_0(A)\to A^{n-1})=\mathrm{cat}_G(A^{n-1})$, where q is a G-map defined by $q(\gamma):=(\gamma(\frac{1}{n-1}),\dots,\gamma(\frac{n-2}{n-1}),\gamma(1))$.
We now use a similar idea which was used in [Reference Farber and Kumar Paul23, proposition 3.3]. Consider the following commutative diagram:
where F and F ʹ are defined such that
Note that F and F ʹ are homeomorphisms. We will verify that they are indeed G-maps. Let $g\in G$. Then we have $F(g\gamma,gx)(t)= gx \cdot g\gamma(t)= g (x \cdot \gamma(t))=gF(\gamma,x)(t)$ and
Thus, F and F ʹ are G-equivariant homeomorphisms. Consequently, we get
The equality $\mathrm{cat}_G(A^{n-1})=\mathrm{TC}_{G,n}(A)$ follows from proposition 3.6.
4.1. Fibrewise G-homotopy invariance
This subsection aims to achieve a fibrewise G-homotopy invariance of G-equivariant parametrized topological complexity. We begin by proving a lemma which plays a crucial role in showing that the sequential G-equivariant parametrized topological complexity of fibrewise G-homotopy equivalent fibrations coincide.
(1) Suppose $p:E\to B$ and $p':E'\to B$ are G-fibrations. If there is a continuous G-map $f:E\to E'$ which fits into the following commutative diagram
then $\mathrm{secat}_G(p')\leq \mathrm{secat}_G(p)$.
(2) Suppose we have the following commutative diagram of G-maps:
where $p:E\to B$ and $p':E'\to B'$ are G-fibrations. Suppose $f'\circ f\simeq_G id_B$. Then
\begin{align*}\mathrm{secat}_G(p)\leq \mathrm{secat}_G(p').\end{align*}
Proof. For (1), consider a G-invariant open set U of B with a G-section sU of p. Note that $f\circ s_U$ is a G-map, and due to the commutativity of the above diagram, it defines a G-section of p ʹ. Thus, we get the desired inequality in (1).
For (2), consider the pullback fibration $q:B\times_{B'} E'\to B$ of a G-fibration $p':E'\to B'$ along $f:B\to B'$. Then, from [Reference Colman and Grant12, proposition 4.3], we have:
Let $\tilde{F}=F'\circ q'$, where $q':B\times_{B'} E'\to E'$ be the projection onto the second factor. Then, we have $p\circ\tilde{F}((b,x))=p\circ F'\circ q'((b,x))=f'\circ p'(x)$. Since $p'(x)=f(b)$, it follows that $p\circ\tilde{F}((b,x))=f'\circ f\circ q(b,x)$. In other words, $p\circ\tilde{F}=f'\circ f\circ q$. Since $f'\circ f\simeq_G id_B$, we have $p\circ\tilde{F}\simeq_G q$. Note that we have the following G-homotopy commutative diagram:
Now consider a G-homotopy $H_t:B\to B$ such that $H_0=f'\circ f$ and $H_1=id_B$. Using the G-homotopy lifting property, we obtain a G-homotopy $J_t:B\times_{B'} E'\to E$ such $J_0=\tilde{F}$, $p\circ J_t=H_t\circ q$ and $p\circ J_1=q$. Thus, from part (1), we get that $\mathrm{secat}_G(p)\leq \mathrm{secat}_G(q)$. We can now conclude the proof using (11).
We need the following definitions before introducing the equivariant analogue of fibrewise homotopy equivalent fibrations.
Definition 4.11. Let $p:E\to B$ and $q: E'\to B$ be G-fibrations. A fibrewise G-map from $p:E\to B$ to $q: E'\to B$ is a G-map $f:E\to E'$ such that $q\circ f=p$.
Definition 4.12. A fibrewise G-homotopy $F:E\times I\to E'$ is a G-map such that $q(F(-,t))=p$ for all $t\in I$. Thus, F is a G-homotopy between fibrewise G-maps $F(-,0)$ and $F(-,1)$.
Definition 4.13. Let $p:E\to B$ and $q: E'\to B$ be G-fibrations. Then p and q are said to be fibrewise G-homotopy equivalent if there exist fibrewise G-maps $f: E\to E'$ and $g:E'\to E$ such that there are fibrewise G-homotopies from $f\circ g$ to $Id_{E'}$ and from $g\circ f$ to IdE.
Farber and Paul [Reference Farber and Kumar Paul23] show that the sequential parametrized topological complexity of fibrewise homotopy equivalent fibrations coincides. We now prove the equivariant analogue of their result.
Proposition 4.14. Suppose the G-fibrations $p:E\to B$ and $p:E'\to B$ are fibrewise G-homotopy equivalent. Then
Proof. The commutative diagram of G-maps given in the hypothesis induces the following two commutative diagrams of G-maps:
where $F^I(\gamma)(t)=F(\gamma(t))$ and $F'^I$ defined similarly. Since $f'\circ f$ and $f\circ f'$ are fibrewise G-homotopic to the identity maps idE and $id_{E'}$, respectively, we have $f'^n\circ f^n\simeq_G id_{E^n_B}$ and $f^n\circ f'^n\simeq_G id_{E'^n_B}$. Thus, we obtain the desired equality using part (2) of lemma 4.10.
4.2. Bounds
In this subsection, we obtain various bounds on the sequential equivariant parametrized topological complexity. We note that the results in this subsection generalize the classical results of Farber [Reference Farber22, Reference Rudyak38] when we consider the trivial action of the group G, and the results of Colman-Grant [Reference Colman and Grant12] and Bayeh-Sarkar [Reference Bayeh and Sarkar3] when we consider the base space B to be a singleton.
We now establish the upper bound on the sequential equivariant parametrized topological complexity in terms of the equivariant category, and consequently, in terms of the dimension of the fibre product.
Proposition 4.15. Let $p:E\to B$ be a G-fibration with $E^G\neq \emptyset$ and $E^n_B$ is G-connected . Then
Moreover, if G acts locally smoothly on E and F is the fibre of a smooth fibre bundle $p:E\to B$, then
where P is the principal orbit (an orbit of maximal dimension) of the G-action on $E^n_B$.
Proof. One can observe that, since $E^G\neq \emptyset$, we have $(E^n_B)^G\neq \emptyset$. Consequently, $(E^I_B)^G\neq \emptyset$. Moreover, $E^n_B$ is G-connected, thus the first inequality of (12) follows from [Reference Colman and Grant12, proposition 4.4]. The right-most inequality of (12) follows from [Reference Marzantowicz35, corollary 1.12]. It follows from [Reference Bredon5, theorem IV.3.8] that $\mathrm{dim}(E^n_B/G)=\mathrm{dim}(E^n_B)-\mathrm{dim}(P)$. Then, the inequality in (13) follows from the fact that $\mathrm{dim}(E^n_B)\leq n\mathrm{dim}(F)+\mathrm{dim}(B)$.
Corollary 4.16. Suppose E admits a locally smooth, non-trivial and semi-free action of G, with $p:E\to B$ being a G-fibration such that $E^G\neq \emptyset$ and $E^n_B$ is G-connected. Then
Proof. It follows that if the action is locally smooth, non-trivial, and semi-free, then the principal orbit’s dimension coincides with the group’s dimension. Thus, the conclusion follows from proposition 4.15.
Remark 4.17. Consider $p:E\to B$ a G-fibration. When $B= \{\ast\}$, the inequality (14) also appears in [Reference Daundkar and Singh16, proposition 3.1]. Since $\mathrm{TC}_n(X)\leq \mathrm{TC}_{G,n}(X)$, we can conclude that the inequality (14) generalizes [Reference Daundkar13, corollary 4.7].
Proposition 4.18. Let $p:E\to B$ be a G-fibration such that the fibre of $p|_{E^H}:E^H\to B^H$ is path connected for all closed subgroups H of G and let $E^n_B$ be G-connected. Then
Proof. Since $p|_{E^H}:E^H\to B^H$ is path connected for all closed subgroups H of G, the map $\Pi_n|_{(E^I_B)^H}:(E^I_B)^H\to (E^n_B)^H$ is surjective. Therefore, the conclusion follows from [Reference Colman and Grant12, proposition 4.5].
Proposition 4.19. Let $p:E\to B$ be a G-fibration. If E is G-contractible space, then
Proof. Let $x\in E$ and γx be a constant path at x. Then define $h:E\to E^I_B$ by $h(x)=\gamma_x$. Observe that h is a G-map since $h(gx)=\gamma_{gx}=g\cdot \gamma_x=g\cdot h(x)$. Notice that h has a G-homotopy inverse h ʹ, which sends a path to its initial point. This implies, $E\simeq_{G}E^I_B$. This implies $E^I_B$ is a G-contractible space. Therefore, we can apply [Reference Colman and Grant12, proposition 4.6] to the fibration $\Pi_n:E^I_B\to E^n_B$ to get the desired result.
Now, the following result is clear.
Corollary 4.20. Let $p:E\to B$ be a G-fibration and let E be a G-contractible space. If $E^n_B$ is G-connected and $E^G\neq \emptyset$, then
Theorem 4.21 Let $p:E\to B$ be a G fibration. Let H and K be any closed subgroups of G such that EH is K-invariant. Then
Proof. Let U be a G-invariant open set of $E^n_B$ on which we have a local G-section s of $\Pi_n:E^I_B\to E^n_B$. Let $V=U\cap (E^H)^n_B $. Then, observe that V is K-invariant. Note that for $x\in (E^H)^n_B$, we have $hs(x)=s(hx)=s(x)$. Therefore, $s(x)\in (E^I_B)^H$. But $(E^I_B)^H=(E^H)^I_{B^H}$. Thus, s restricts to V. Therefore, we have a K-section $s|_{V}:V\to (E^H)^I_B$. This concludes the result.
The following is an obvious consequence of theorem 4.21.
Corollary 4.22. Let $p:E\to B$ be a G-fibration. Then for any closed subgroups H and K of G we have:
(1) $\mathrm{TC}_n[p|_{E^H}:E^H\to B^H]\leq \mathrm{TC}_{G,n}[p:E\to B]$,
(2) $\mathrm{TC}_{K,n}[p:E\to B]\leq \mathrm{TC}_{G,n}[p:E\to B]$.
Remark 4.23. If B is a singleton, then theorem 4.21 generalizes [Reference Colman and Grant12, proposition 5.3] and [Reference Bayeh and Sarkar3, proposition 3.14].
In the following result, we establish the relation between sequential equivariant parametrized topological complexity of a G-fibration and the equivariant category of its fibre and the equivariant category of the corresponding fibre product. The proof is inspired by [Reference Colman and Grant12, proposition 5.7].
Theorem 4.24 Let $p:E\to B$ be a G-fibration such that $E^n_B$ is G-connected, and let H be a stabilizer of some $e\in E$. Then,
In particular, if $E^G\neq \emptyset$, then
Proof. Let $e\in E$ and $b=p(e)$. We consider $F=p^{-1}(b)$. Note that $b\in B^H$. Therefore, F admits an H-action. We now define a map $f:F^{n-1}\to E^n_B$ by $f(x_1,\dots,x_{n-1})=(e,x_1,\dots,x_{n-1})$. Note that f is a well-defined H-equivariant map. Now consider the following pullback diagram:
where $Y=\{\gamma\in E^I_B \mid \gamma(0)=e\}$. Note that Y is H-invariant and H-contractible. Therefore, $\mathrm{cat}_H(F^{n-1})\leq \mathrm{secat}_{H}(q)$. From [Reference Colman and Grant12, proposition 4.3], we have:
Now from corollary 4.22, we have $\mathrm{TC}_{H,n}[p:E\to B] \leq \mathrm{TC}_{G,n}[p:E\to B]$. This proves the desired inequality.
We will now obtain the cohomological lower bound on the equivariant parametrized topological complexity. Let X be a G-space, and $X^h_G=EG\times_{G} X$ be the homotopy orbit space of X. The Borel G-equivariant cohomology of X is defined as $H^{\ast}_G(X;R):=H^{\ast}(X^h_G;R)$, where R is any commutative ring.
Let $p:E\to B$ be a G-fibration and $E^n_B$ be the corresponding fibre product. Note that the diagonal map $\triangle: E\to E^n_B$ is G-equivariant. Therefore, it induces a map $\triangle_G: E^h_G\to (E^n_B)^h_G$.
Theorem 4.25 Let $\triangle_G^{\ast}:H^{\ast}_G(E^n_B)\to H^{\ast}_G(E)$ be the induced map from $\triangle_G: E^h_G\to (E^n_B)^h_G$. Suppose there exists cohomology classes $u_i\in \mathrm{ker}(\triangle_G^{\ast})$ for $1\leq i\leq r$ such that $u_1\cup \dots \cup u_r\neq 0$. Then
Proof. Consider the following commutative diagram:
where h is a G-homotopy equivalence. Suppose $\mathrm{TC}_{G,n}[p:E\to B]\leq r$. Consider a parametrized motion planning cover $\{U_1,\dots, U_r\}$. Since h is G-homotopy equivalence, the inclusion $i_{U_i}:U_i\to E^n_B$ factors through the diagonal $\triangle: E\to E^n_B$ up to G-homotopy. Therefore, $i_{U_i}^{\ast}(u_i)=0$ for $1\leq i\leq r$. Hence ui’s are in the image of $H^{\ast}_G(E^n_B, U_i)\to H^{\ast}_G(E^n_B)$ for $1\leq i\leq r$. Since Ui covers $E^n_B$, we get that $u_1\cup\dots\cup u_r=0$ by the naturality of cup products. This completes the proof.
The upper bound on the equivariant sectional category of Serre G-fibrations has obtained in [Reference Grant30, theorem 3.5]. We now state this result with a slight modification.
Theorem 4.26 ([Reference Grant30, theorem 3.5])
Let $p:E\to B$ be a Serre G-fibration and let B be a G-CW complex with $\dim(B)\geq 2$. Suppose for all closed subgroups H of G, the fibre of $p^H:E^H\to B^H$ is $(m-1)$-connected. Then
We will now use theorem 4.26 to establish the upper bound on the equivariant parametrized topological complexity.
Theorem 4.27 Let $p:E\to B$ be a (Hurewicz) G-fibration and $E^n_B$ is a G-CW complex with $\mathrm{dim}(E^n_B)\geq 2$. Suppose for all subgroups H of G, the fibre of $p^H:E^H\to B^H$ is m-connected. Then
Proof. Since $\Pi_n:E^I_B\to E^n_B$ is a (Hurewicz) G-fibration, it is also a Serre G-fibration. We can observe that $(E^I_B)^H=(E^H)^I_{B^H}$ and $(E^n_B)^H=(E^H)^n_{B^H}$ for any closed subgroup H of G. Thus, the fibration $(\Pi_n)^H:(E^I_B)^H\to (E^n_B)^H$ is, in fact, a sequential parametrized path fibration associated with the fibration $p^H:E^H\to B^H$. In other words, we have $(\Pi_n)^H:(E^H)^I_{B^H}\to (E^H)^n_{B^H}$. Let XH denote the fibre of $p^H: E^H\to B^H$. Then, note that the fibre of $(\Pi_n)^H$ is $(\Omega X_H)^{n-1}$. Since XH is m-connected and $\pi_i((\Omega X_H)^{n-1})=(\pi_i(\Omega X_H))^{n-1}=(\pi_{i+1}(X_H))^{n-1}$, it follows that $(\Omega X_H)^{n-1}$ is $(m-1)$-connected. Then from theorem 4.26, we obtain the following inequality:
It is well known that there is a locally trivial fibration $E^n_B\to B$ with fibre F n. Therefore, we have $\mathrm{dim}(E^n_B)\leq n\mathrm{dim}(F)+\mathrm{dim}(B)$. This completes the proof.
5. Examples
5.1. Equivariant Fadell–Neuwirth fibrations
In this subsection, we compute sequential equivariant parametrized topological complexity of Fadell–Neuwirth fibrations.
We start with recalling the definition of the ordered configuration space. The ordered configuration space of s points on $\mathbb{R}^d$ is denoted by $F(\mathbb{R}^d,s)$ and defined as
Definition 5.1 ([Reference Fadell and Neuwirth20])
The maps
are called Fadell–Neuwirth fibrations.
The sequential parametrized topological complexity of these fibrations have been computed in [Reference Farber and Kumar Paul23] for odd $d\geq 3$ and in [Reference Farber and Kumar Paul24] for even $d\geq 2$. We now state their theorem.
Theorem 5.2 ([Reference Farber and Kumar Paul24, theorem 1.3])
Let $d\geq 2$ be an even integer, and $t\geq 1$, $s\geq 2$. Then
Theorem 5.3 ([Reference Farber and Kumar Paul23, theorem 8.1])
Let $d\geq 3$ be an odd integer, and $s \geq 2$, $t \geq 1$. Then
We now define an action of a permutation group $\Sigma_s$ (on s letters) on $F(\mathbb{R}^d,s+t)$. Let $\sigma\in \Sigma_s$. Define
Also $\Sigma_s$ acts on $F(\mathbb{R}^d,s)$ by permuting coordinates. Observe that the map p in definition 5.1 satisfies
Thus, p is a $\Sigma_s$-equivariant map. We now show that p is indeed a $\Sigma_s$-fibration.
The local triviality of p was shown in [Reference Fadell and Husseini21, theorem 1.1], and we also refer the reader to [Reference Cohen10, theorem 3.3]. We will use the ideas from these results to show the $\Sigma_s$-equivariant local triviality of p. Consider a point $q=(q_1,\dots,q_s)\in F(\mathbb{R}^d,s)$ and mutually disjoint open discs Di with centres qi. Then consider the open set $V=\cup_{\sigma\in \Sigma_s}D_{\sigma(1)}\times \dots\times D_{\sigma(s)}$ containing q. Note that V is $\Sigma_s$-invariant and a disjoint union. Note that the fibre Fq of p at q is given by $F(\mathbb{R}^d\setminus \{q_1,\dots,q_s\},t)$. We identify Fq with the last t-coordinates of $F(\mathbb{R}^d,s+t)$ via $(y_1,\dots,y_t)\mapsto (q_1,\dots,q_s,y_1,\dots,y_t)$ and assume that $\Sigma_s$ acts trivially on Fq. We define a homeomorphism $\Phi_q:V\times F_q\to p^{-1}(V)$ in a similar way as it was defined in [Reference Fadell and Husseini21, theorem 1.1]. We fix the notation $\bar{y}=(y_1,\dots,y_t)$. Let $\bar{x}=(x_1,\dots,x_s)\in V$. That means $\bar{x}\in D_{\sigma(1)}\times \dots\times D_{\sigma(s)}$ for some $\sigma\in \Sigma_s$. Define
where functions $\gamma_i(x_i):\mathbb{R}^d\to \mathbb{R}^d$ are homeomorphisms defined as in the [Reference Fadell and Husseini21, corollary 1.1] for $1\leq i\leq s$ and we have denoted $(\gamma_{i}(x))^{-1}$ by $\gamma_i^{-1}(x)$. Recall that these functions have the following properties: for $x\in D_i$, we have $\gamma_i(x)(D_i)=D_i$, $\gamma_i(x)(x)=q_i$, and $\gamma_i(x)(y)=y$ for $y\in \mathbb{R}^d\setminus D_i$. Note that
For $(x_1,\dots,x_s)\in D_1\times\dots\times D_s$, we note the important observation:
We will now prove that $\Phi_q$ is $\Sigma_s$-equivariant. Let $\bar{x}\in D_{\sigma(1)}\times\dots\times D_{\sigma(s)}$ and $\rho\in\Sigma_s$. Note that $\rho\cdot \bar{x}\in D_{\sigma(\rho(1))}\times\dots\times D_{\sigma(\rho(s))}$. Then we have
Observe that the sets $\{\gamma_{\sigma(r)}^{-1}(x_r) \mid 1\leq r\leq s\}$ and $\{\gamma_{\sigma(\rho(r))}^{-1}(x_{\rho(r)}) \mid 1\leq r\leq s\}$ coincides. Then using (16) we can see that
More precisely, if $y_i\in D_j$ for some $1\leq j \leq s$ and $\sigma(k)=j$, then we have
Moreover, if $\sigma(\rho(r))=j$, then
Since $\rho(r)=k$, we get $\gamma^{-1}_{\sigma(k)}(x_k)(y_i)=\gamma^{-1}_{\sigma(\rho(r))}(x_{\rho(r)})(y_i)$. This gives us,
This proves that $\Phi_q$ is $\Sigma_s$-equivariant.
Let $F=F(\mathbb{R}^d\setminus \{a_1,\dots,a_s\},t)$ be the fibre of p with $(a_1,\dots,a_s)\in F(\mathbb{R}^d,s)$ as chosen base point. We will now choose a stable homeomorphism (as in step-2 of [Reference Fadell and Husseini21, theorem 1.1]) $\alpha_q:F\to F_q$ and define $\Phi=\Phi_q\circ (id_V\times \alpha_q)$. Note that αq is equivariant as $\Sigma_s$ acts trivially on both F and Fq. We conclude that $\Phi:V\times F\to p^{-1}(V)$ is $\Sigma_s$-equivariant local trivialization. Finally, using [Reference İlhan32, theorem 2.3], we conclude that the fibration p is a G-fibration.
We are going to use theorem 4.27 to obtain an upper bound on $\mathrm{TC}_{\Sigma_s,n}[p:F(\mathbb{R}^d,s+t)\to F(\mathbb{R}^d, s)]$. We denote the fibration $p:F(\mathbb{R}^d,s+t)\to F(\mathbb{R}^d, s)$ by $p:E\to B$. Note that F is $(d-2)$-connected. Since any closed subgroup H of $\Sigma_s$ acts freely on E and B, we only need to consider $p^H:E^H\to B^H$ when H is the trivial subgroup. In this case, $p^H=p$. Then from theorem 4.27, we get the following expression.
This gives us
Since we have
we establish the following result.
(1) Let $d\geq 3$ be an odd integer, and $s\geq 2$, $t\geq 1$. Then
\begin{align*}\mathrm{TC}_{\Sigma_s,n}[p:F(\mathbb{R}^d,s+t)\to F(\mathbb{R}^d, s)]=nt+s.\end{align*}(2) Let $d\geq 2$ be an even integer, and $t\geq 1$, $s\geq 2$. Then
\begin{align*}\mathrm{TC}_{\Sigma_s,n}[p:F(\mathbb{R}^d,s+t)\to F(\mathbb{R}^d, s)]\in \{nt+s-1,nt+s\}.\end{align*}
5.2. Some generalized projective product spaces
Let τ be any involution on M. Consider the following generalized projective product space
We have a fibre bundle $p:X(M, S^{2m+1})\to \mathbb{R} P^{2m+1}$ with fibre M. We define a $\mathbb{Z}_2$-action on $X(M, S^{2m+1})$ by defining an involution
where $y=(y_0,y_1,\dots,y_{2m},y_{2m+1})$ and $\tilde{y}=(-y_1,y_0,\dots, -y_{2m+1},y_{2m})$. It can be observed that σ defines a free involution on $X(M, S^{2m+1})$. A $\mathbb{Z}_2$-action on $\mathbb{R} P^{2m+1}$ is defined by
We now observe the following:
This shows that $p:X(M,S^{2m+1})\to \mathbb{R} P^{2m+1}$ is a $\mathbb{Z}_2$-equivariant map. One can also show that p is in fact a G-fibration. Observe that
Since $\mathbb{Z}_2$ acts freely on $X(M,S^{2m+1})$, we have free diagonal $\mathbb{Z}_2$-action on $X(M,S^{2m+1})^n_{\mathbb{R} P^{2m+1}}$. Thus $X(M,S^{2m+1})^n_{\mathbb{R} P^{2m+1}}$ is $\mathbb{Z}_2$-connected. Therefore using proposition 4.18 we get the following inequality:
From [Reference Colman and Grant12, proposition 3.5] we get $\mathrm{cat}_{\mathbb{Z}_2}(X(M,S^{2m+1})^n_{\mathbb{R} P^{2m+1}})=\mathrm{cat}(X(M,S^{2m+1})^n_{\mathbb{R} P^{2m+1}}/\mathbb{Z}_2)$. Using the dimensional inequality for category, we obtain
In particular, if $M=\mathbb{R} P^k$ with any involution, then we have
Suppose $k=2^r$, then using [Reference Farber, Tabachnikov and Yuzvinsky26, corollary 8.2], we have
Acknowledgements
The author expresses sincere gratitude to the reviewer for numerous insightful suggestions, comments, and feedback that significantly improved the exposition and presentation of this article. In particular, the author thanks the reviewer for prompting the proof that $\Pi_n$ is a G-fibration, for suggesting the correct version of theorem 4.26 and the modifications in the proof of theorem 4.27, and for many other valuable suggestions. The author also thanks Soumen Sarkar for valuable feedback on an earlier version of the article, and Mark Grant for his valuable suggestions and for providing resources related to equivariant principal bundles, which were instrumental in proving theorem 4.9. Additionally, the author appreciates Tejas Kalelkar and Ramandeep Singh Arora for helpful discussions. The author acknowledges the support of NBHM through grant 0204/10/(16)/2023/R&D-II/2789.