Skip to main content

Advertisement

Log in

Satellite City Formation for a Spatial Economic Model

Bifurcation Mechanism in a Hexagonal Domain

  • Published:
Networks and Spatial Economics Aims and scope Submit manuscript

Abstract

The economic agglomeration of one large city surrounded by satellite cities is observed worldwide and is a topic of keen economic interest. We theoretically investigate where such satellite cities emerge in a two-dimensional economic space in which discrete locations are evenly distributed in a regular-hexagonal domain. To elucidate this emergence, we introduce two viewpoints: (1) the bifurcation mechanism of the full agglomeration at the geographical center in this domain (mono-center), which produces satellite cities around this center, and (2) the existence of invariant patterns, which are equilibria for any value of the transport cost parameter. Theoretically-predicted agglomeration patterns are ensured to exist as stable equilibria for a spatial economic model proposed by Forslid and Ottaviano (2003). We theoretically find one large central city surrounded by hexagonal satellite cities that is a two-dimensional counterpart of the core-periphery pattern (Krugman 1991). Moreover, we demonstrate that spatial patterns of twin cities, three cities, and racetrack cities are absorbed into the mono-center as the transport cost decreases. These transitions are ubiquitously observed in the two-dimensional spatial platform with the geographical center.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

Data Availability

We do not analyze or generate any datasets because our work proceeds within a theoretical and mathematical approach.

Notes

  1. For example, with a ring and heterogeneous star network topologies, Barbero and Zofío (2016) analyzed the agglomeration and dispersion forces of the core-periphery model.

  2. Several studies for spatial agglomeration have been conducted on a square lattice (Clarke and Wilson 1983; Weidlich and Haag 1987; Weidlich and Munz 1990). Moreover, Stelder (2005) carried out a simulation of agglomeration for cities in Europe using a grid of points. However, a hexagonal lattice is employed in this paper since it has a finer spatial resolution than a square lattice.

  3. Akamatsu et al. (2019) elucidated the bifurcation mechanism of the FO model in racetrack economy. This mechanism, however, was restricted to the bifurcation from the uniform state.

  4. For example, two places (Krugman 1991), three places (Castro et al. 2012), and racetrack (Ikeda et al. 2019b) are spatial platforms without the geographical center.

  5. Only mobile workers are input for manufacturing production in the core-periphery model proposed by Krugman (1991), whereas both mobile and inmobile workers are the input in the FO model. This is the critical difference between the Krugman and FO models.

  6. See, e.g, Golubitsky et al. (Sec. 13.3.1, 1988) for the theoretical detail of the orbit decomposition.

  7. The full agglomeration was shown to be an invariant solution in a racetrack economy for any number of places in Castro et al. (2012).

  8. To name a few, twin cities were studied by Krugman (1991), the triangular cities by Castro et al. (2012), and racetrack cities by Akamatsu et al. (2019).

  9. Such pattern was also observed in Ikeda et al. (2017b) and was called a core-satellite pattern (Gaspar et al. 2020).

  10. For one-dimensional space (a long narrow economy), a similar result is observed (Gaspar et al. 2020).

  11. Figures 11(a) and 13(c) contain all stable bifurcating solutions from the mono-center.

  12. The emergence of central city from racetrack cities was observed in an ad-hoc manner in Ikeda et al. (2017b).

  13. Since the solution space of the governing equation is the \(\left( K-1\right)\)-dimensional simplex with a constant total population, the eigenvector \(\varvec{\eta }^*=(1,\ldots ,1)\) and the associated eigenvalue \(e^*\) must be excluded in the investigation of stability and sustainability of the solutions.

References

  • Ago T, Isono I, Tabuchi T (2006) Locational disadvantage of the hub. Ann Reg Sci 40(4):819–848

    Article  Google Scholar 

  • Aizawa H, Ikeda K, Osawa M, Gaspar J (2020) Breaking and sustaining bifurcations in \(S_N\)-invariant equidistant economy. Int J Bifurcation Chaos 30(16):2050240

  • Akamatsu T, Mori T, Osawa M, Takayama Y (2019) Endogenous agglomeration in a many-region world. arXiv:1912.05113

  • Akamatsu T, Takayama Y, Ikeda K (2012) Spatial discounting, fourier, and racetrack economy: A recipe for the analysis of spatial agglomeration models. J Econ Dyn Control 36(11):1729–1759

    Article  Google Scholar 

  • Baldwin R, Forslid R, Martin P, Ottaviano G, Robert-Nicoud F (2003) Economic geography and public policy. Princeton University Press

    Google Scholar 

  • Barbero J, Zofío JL (2016) The multiregional core-periphery model: The role of the spatial topology. Netw Spat Econ 16(2):469–496

    Article  Google Scholar 

  • Castro SB, Correia-da Silva J, Mossay P (2012) The core-periphery model with three regions and more. Pap Reg Sci 91(2):401–418

    Google Scholar 

  • Clarke M, Wilson AG (1983) The dynamics of urban spatial structure: progress and problems. J Reg Sci 23(1):1–18

    Article  Google Scholar 

  • Flam H, Helpman E (1987) Industrial policy under monopolistic competition. J Int Econ 22(1):79–102

    Article  Google Scholar 

  • Forslid R, Ottaviano GI (2003) An analytically solvable core-periphery model. J Econ Geogr 3(3):229–240

    Article  Google Scholar 

  • Fujita M, Krugman P, Mori T (1999a) On the evolution of hierarchical urban systems. Eur Econ Rev 43(2):209–251

  • Fujita M, Krugman P, Venables A (1999b) The spatial economy: Cities, regions, and international trade. MIT press

  • Fujita M, Mori T (1997) Structural stability and evolution of urban systems. Reg Sci Urban Econ 27(4–5):399–442

    Article  Google Scholar 

  • Gaspar JM, Castro SB, Correia-da Silva J (2018) Agglomeration patterns in a multi-regional economy without income effects. Econ Theor 66(4):863–899

    Article  Google Scholar 

  • Gaspar JM, Ikeda K, Onda M (2021) Global bifurcation mechanism and local stability of identical and equidistant regions: Application to three regions and more. Reg Sci Urban Econ 86:103597

    Article  Google Scholar 

  • Golubitsky M, Stewart I, Schaeffer DG (1988) Singularities and Groups in Bifurcation Theory, Vol. 2. Springer-Verlag

  • Ikeda K, Aizawa H, Gaspar JM (2020) How and where satellite cities form around a large city: Bifurcation mechanism of a long narrow economy. MPRA Paper No. 104748

  • Ikeda K, Kogure Y, Aizawa H, Takayama Y (2019a) Invariant patterns for replicator dynamics on a hexagonal lattice. Int J Bifurcation Chaos 29(06):1930014

  • Ikeda K, Murota K, Akamatsu T, Takayama Y (2017a) Agglomeration patterns in a long narrow economy of a new economic geography model: Analogy to a racetrack economy. Int J Econ Theory 13(1):113–145

  • Ikeda K, Murota K, Takayama Y (2017b) Stable economic agglomeration patterns in two dimensions: Beyond the scope of central place theory. J Reg Sci 57(1):132–172

  • Ikeda K, Onda M, Takayama Y (2018) Spatial period doubling, invariant pattern, and break point in economic agglomeration in two dimensions. J Econ Dyn Control 92:129–152

    Article  Google Scholar 

  • Ikeda K, Onda M, Takayama Y (2019b) Bifurcation theory of a racetrack economy in a spatial economy model. Netw Spat Econ 19(1):57–82

  • Ikeda K, Osawa M, Takayama Y (2022) Time evolution of city distributions in germany. Netw Spat Econ 22(1):125–151

    Article  Google Scholar 

  • Krugman P (1991) Increasing returns and economic geography. J Polit Econ 99(3):483–499

    Article  Google Scholar 

  • Mori T (1997) A modeling of megalopolis formation: the maturing of city systems. J Urban Econ 42(1):133–157

    Article  Google Scholar 

  • Sandholm WH (2010) Population games and evolutionary dynamics. MIT press

  • Stelder D (2005) Where do cities form? a geographical agglomeration model for europe. J Reg Sci 45(4):657–679

    Article  Google Scholar 

  • Tabuchi T, Thisse JF (2011) A new economic geography model of central places. J Urban Econ 69(2):240–252

  • Takayama Y, Ikeda K, Thisse JF (2020) Stability and sustainability of urban systems under commuting and transportation costs. Reg Sci Urban Econ 84:103553

    Article  Google Scholar 

  • Weidlich W, Haag G (1987) A dynamic phase transition model for spatial agglomeration processes. J Reg Sci 27(4):529–569

    Article  Google Scholar 

  • Weidlich W, Munz M (1990) Settlement formation. Ann Reg Sci 24(2):83–106

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to Prof. T. Mori, Prof. S. Mun, and Prof. T. Ago for very detailed and useful comments. We thank participants to the 34th Applied Regional Science Conference for insightful comments. We are also grateful to two anonymous referee for useful comments and suggestions. We acknowledge the financial support of the Japan Society for the Promotion of Science under the grants 18K04380 and 21K04299.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hiroki Aizawa.

Ethics declarations

Conflict of Interest

The authors of this manuscript declare that there is no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A. Detail of the FO Model

The fundamental logic and the governing equation of a multi-places version of the model by Forslid and Ottaviano (2003) are presented (see Akamatsu et al. 2012). The budget constraint is given by

$$\begin{aligned} p^\textrm{A}_i C_i^\textrm{A}+\sum _{j \in P} \int _0^{n_j}p_{ji}(\ell )q_{ji}(\ell )d\ell =Y_i, \end{aligned}$$
(24)

where \(p^\textrm{A}_i\) represents the price of the A-sector good in place i, \(C_i^\textrm{A}\) is the consumption of A-sector goods in place i, \(P= \left\{ 1,...,K\right\}\), \(n_j\) is the number of varieties produced in place j, \(p_{ji}(\ell )\) denotes the price of a variety \(\ell\) in place i produced in place j, \(q_{ji}(\ell )\) is the consumption of variety \(\ell \in [0, n_j]\) in place i produced in place j, and \(Y_i\) is the income of an individual in place i. The incomes (wages) of skilled workers and unskilled workers are represented respectively by \(w_i\) and \(w_i^\textrm{L}\).

An individual at place i maximizes the utility in (5) subject to the budget constraint in (24). This maximization yields the following demand functions

$$\begin{aligned} C_i^\textrm{A}=(1-\mu )\frac{Y_i}{p_i^\textrm{A}}, \qquad C_i^\textrm{M}=\mu \frac{Y_i}{\rho _i}, \qquad q_{ji}(\ell )= \mu \frac{\rho _i^{\sigma -1}Y_i}{p_{ji}(\ell )^{\sigma }}, \end{aligned}$$

where \(\rho _i\) denotes the price index of the differentiated products in place i, and is given by

$$\begin{aligned} \rho _i=\left( \sum _{j \in P} \int _0^{n_j}p_{ji}(\ell )^{1-\sigma }d\ell \right) ^{1/(1-\sigma )}. \end{aligned}$$
(25)

Because the total income in place i is \(w_i \lambda _i+w_i^\textrm{L}\), the total demand \(Q_{ji}(\ell )\) in place i for a variety \(\ell\) produced in place j is given as

$$\begin{aligned} Q_{ji}(\ell )=\mu \frac{\rho _i^{\sigma -1}}{p_{ji} (\ell )^{\sigma }}(w_i\lambda _i+w_i^\textrm{L}). \end{aligned}$$
(26)

The A-sector is perfectly competitive and produces homogeneous goods under constant-returns-to-scale, and requires one unit of unskilled labor per unit of output. The A-sector good is traded freely across locations and is chosen as the numéraire. In equilibrium, \(p_i^\textrm{A}=w_i^\textrm{L}=1\) for each i.

The M-sector output is produced under increasing-returns-to-scale and Dixit–Stiglitz monopolistic competition. A firm incurs a fixed input requirement of \(\alpha\) units of skilled labor and a marginal input requirement of \(\beta\) units of unskilled labor. An M-sector firm located in place i chooses \(( p_{ij}(\ell ) \mid j \in P)\) that maximizes its profit

$$\begin{aligned}&\Pi _i(\ell )=\sum _{j \in P} p_{ij}(\ell )Q_{ij}(\ell )-\left( \alpha w_i+\beta x_i(\ell )\right) , \end{aligned}$$
(27)

where \(x_i(\ell )\) denotes the total supply of variety \(\ell\) produced in place i and \(\alpha w_i+\beta x_i(\ell )\) signifies the cost function introduced by Flam and Helpman (1987).

With the use of the iceberg form of the transport cost, we have

$$\begin{aligned} x_i(\ell )=\sum _{j \in P} \tau _{ij}Q_{ij}(\ell ). \end{aligned}$$
(28)

Then the profit function of the M-sector firm in place i, given in (27) above, can be rewritten as

$$\begin{aligned}&\Pi _i(\ell )=\sum _{j \in P} p_{ij}(\ell )Q_{ij}(\ell )- \left( \alpha w_i+\beta \sum _{j \in P} \tau _{ij}Q_{ij}(\ell ) \right) , \end{aligned}$$

which is maximized by the firm. The first-order condition for this profit maximization yields the following optimal price

$$\begin{aligned} p_{ij}(\ell )=\frac{\sigma \beta }{\sigma -1} \tau _{ij}. \end{aligned}$$
(29)

This result implies that \(p_{ij}(\ell )\), \(Q_{ij}(\ell )\), and \(x_i(\ell )\) are independent of \(\ell\). Therefore, the argument \(\ell\) is suppressed subsequently.

In the short run, skilled workers are immobile between places, i.e., their spatial distribution \(\varvec{\lambda }=(\lambda _i \mid i \in P)\) is assumed to be given. The market equilibrium conditions consist of three conditions: the M-sector goods market clearing condition, the zero-profit condition attributable to the free entry and exit of firms, and the skilled labor market clearing condition. The first condition is written as (28) above. The second one requires that the operating profit of a firm, given in (27), be absorbed entirely by the wage bill of its skilled workers. This gives

$$\begin{aligned} w_i =\frac{1}{\alpha }\left\{ \sum _{j \in P} p_{ij}Q_{ij} -\beta x_i \right\} . \end{aligned}$$
(30)

The third condition is expressed as \(\alpha n_i=\lambda _i\) and the price index \(\rho _i\) in (25) can be rewritten using (29) as

$$\begin{aligned} \rho _i=\frac{\sigma \beta }{\sigma -1} \left( \frac{1}{\alpha }\sum _{j \in P} \lambda _j d_{ji}\right) ^{1/(1-\sigma )}. \end{aligned}$$
(31)

The market equilibrium wage \(w_i\) in (30) can be represented as

$$\begin{aligned} w_i =\frac{\mu }{\sigma } \sum _{j \in P} \frac{d_{ij}}{\Delta _j}(w_j \lambda _j+1) \end{aligned}$$
(32)

using \(d_{ji}= \tau _{ji}^{1-\sigma }= \phi \, ^{m(i,j)}\), (26), (28), (29), and (31). Here, \(\Delta _{j} = \sum _{k \in P} d_{kj}\lambda _k\). Equation (32) can be rewritten, using \(\varvec{w}=(w_1, \ldots , w_K)\), as \(\varvec{w}= \frac{\mu }{\sigma } \ D\Delta ^{-1} (\Lambda \varvec{w}+\varvec{1})\), which is solved for \(\varvec{w}\) as

$$\begin{aligned} \varvec{w}= \frac{\mu }{\sigma } \left( I- \frac{\mu }{\sigma } \ D\Delta ^{-1}\Lambda \right) ^{-1} D\Delta ^{-1} \varvec{1} \end{aligned}$$
(33)

with I being the identity matrix, \(\varvec{1}=(1,\ldots ,1)^\top\), and

$$\begin{aligned} D=(d_{ij}), \quad \Delta =\textrm{diag}(\Delta _1,\ldots ,\Delta _{K}), \quad \Lambda =\textrm{diag}(\lambda _1,\ldots ,\lambda _{K}). \end{aligned}$$
(34)

B. Classifications of Stationary Points

The mono-center is classified into a corner solution, for which some places have zero population. We can appropriately permute the components of \(\varvec{\lambda }\), without loss of generality, to arrive at

$$\begin{aligned} \hat{\varvec{\lambda }}= (\varvec{\lambda }_+, \varvec{0}_{K-m}), \end{aligned}$$

where all components of \(\varvec{\lambda }_+ = (\lambda _1, \ldots , \lambda _m)\) are positive and \(\textbf{0}_{K-m}\) is the \((K-m)\)-dimensional zero vector. \(\varvec{0}_{K-m}\) is present for corner solutions (\(K > m\)). For corner solutions, the governing equation (4) and associated Jacobian matrix can be rearranged, respectively, as (Ikeda et al. 2018)

$$\begin{aligned} \hat{\varvec{F}}= \begin{pmatrix} {\varvec{F}}_+ (\hat{\varvec{\lambda }},\phi ) \\ {\varvec{F}}_0 (\hat{\varvec{\lambda }},\phi ) \end{pmatrix}, \qquad \hat{J} = \frac{\partial \hat{\varvec{F}}}{\partial \hat{\varvec{\lambda }}}= \begin{pmatrix} J_+ &{} J_{+0} \\ O &{} J_{0} \end{pmatrix}, \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} J_+&= \textrm{diag}(\lambda _1, \lambda _2, \ldots , \lambda _{m}) \times \left( {\displaystyle \frac{\partial (v_i - \bar{v})}{\partial \lambda _j} } \mid i,j = 1, \ldots , m \right) , \\ J_{+0}&= \textrm{diag}(\lambda _1, \lambda _2, \ldots , \lambda _{m}) \times \left( \frac{\partial (v_i - \bar{v})}{\partial \lambda _j} \mid i = 1, \ldots , m; \ j = m + 1, \ldots , K \right) , \\ J_0&= \textrm{diag}(v_{m+1} - \bar{v}, \ldots , v_{K} - \bar{v}), \end{aligned} \end{aligned}$$
(35)

and diag\((\cdots )\) denotes a diagonal matrix with the entries in parentheses.

The stability condition of a stable corner solution is decomposed into two conditions:

$$\begin{aligned} \left\{ \begin{array}{ll} \text {Stability condition for } \varvec{\lambda}_+\!\!: &{} \text {all eigenvalues of }J_+\text { are negative.} \\ \text {Sustainability condition:}&{} \text {all diagonal entries of } J_0 \text { are negative.} \end{array}\right. \end{aligned}$$

Both of these two conditions are satisfied if and only if all eigenvalues of \(\hat{J}\) are negative.Footnote 13

Critical points are those which have one or more zero eigenvalue(s) of the Jacobian matrix \(\hat{J}\). Critical points are classified into a bifurcation point with singular \(J_+\) or \(J_0\), and a limit point of \(\phi\) with singular \(J_+\). We classify bifurcation points into a break bifurcation point with singular \(J_+\) and a corner bifurcation point with singular \(J_0\). The corner bifurcation from the mono-center is theoretically investigated in this paper.

C. Theoretical Details

1.1 C.1 Proof of Lemma 1

Since \(T(g)\varvec{\lambda }= \varvec{\lambda }\) for a subgroup \(G'\) of G, we have \(T(g)\varvec{v}(\varvec{\lambda }, \phi ) = \varvec{v}(\varvec{\lambda }, \phi )\) for \(g \in G'\) by (10). This means that \(v_i\) in the same orbit is permutable. This suffices for the proof.

1.2 C.2 Proof of Lemma 2

We can rearrange the components of \(\varvec{\lambda }_{P_l}\) to arrive at \(\hat{\varvec{\lambda }} = (\frac{1}{m}\varvec{1}_m, \varvec{0}_{K - m})\), where \({\varvec{1}}_m\) is the m-dimensional all-one vector, and \({\varvec{0}}_{K - m}\) is the \((K-m)\)-dimensional zero vector. The \(m~(=N_l)\) places belonging to \({\varvec{\lambda }}_+ = \frac{1}{m}\varvec{1}_m\) are permuted each other by the geometrical transformation by an element of a subgroup of G and \({\varvec{\lambda }}_+\) is invariant with respect to the permutation. By equivariance (10), we have \(v_1=\cdots =v_m\), as well as \(\lambda _1=\cdots =\lambda _m=1/m\). Thus we have \(v_i = \bar{v} \ (i \in P_l)\).

1.3 C.3 Proof of Proposition 1

By Lemma 2, we have \(v_{i}-\bar{v}=0\) \((i=1,\ldots ,m)\). Thus, \({\varvec{F}}_+(\frac{1}{m}{\varvec{1}}_m,\textbf{0}_{K-m},\phi )=\textbf{0}_{m}\) holds. For \(K-m\) places with no population, we have \(\lambda _{j}=0\). Hence, \({\varvec{F}}_0(\frac{1}{m}{\varvec{1}}_m,\textbf{0}_{K-m}, \phi )=\textbf{0}_{K-m}\) holds. This shows that \(({\varvec{\lambda }}_+,{\varvec{\lambda }}_0,\phi ) =(\frac{1}{m}\varvec{1}_m,\textbf{0}_{K-m}, \phi )\) serves as a solution to (4) for any \(\phi\).

1.4 C.4 Proof of Proposition 2

By choosing \(G=E\), which leaves each place unchanged, each place forms an orbit. Hence, the full agglomeration at any place i is an invariant pattern by Proposition 1.

1.5 C.5 Proof of Proposition 3

The bifurcating conditions of the Jacobian matrix (13) are given by as follows:

$$\begin{aligned} v_1&= 0, \end{aligned}$$
(36)
$$\begin{aligned} v_{\alpha i} - v_c&= 0, \quad i = 1,\ldots ,n_1, \end{aligned}$$
(37)
$$\begin{aligned} v_{\beta i} - v_c&= 0, \quad i = 1,\ldots ,n_2. \end{aligned}$$
(38)

However, no bifurcation solution with the direction \((1, 0, \ldots , 0)\) in the space \(\sum ^K_{i=1} \lambda _i = 1\) emerges. Therefore, only (37) and (38) are the bifurcating conditions from the mono-center.

1.6 C.6 Proof of Lemma 3

By the product form of the replicator dynamics in (3), \(\tilde{F}_i(\varvec{x},\psi )\) takes the product form: \(\tilde{F}_i(\varvec{x},\psi )=x_i G_i(\varvec{x},\psi )\) \((i\in P_l)\). Since the group \(\textrm{D}_6\) is generated by the elements r and s, it suffices to consider the symmetry condition \(T(g)\tilde{\varvec{F}}(\varvec{x})=\tilde{\varvec{F}}(T(g)\varvec{x})\) for \(g=r,s\). This condition for \(g=r\) gives the form (19) for some function \(R(\varvec{x},\psi )\) and that for \(g=s\) gives the symmetry condition (20).

1.7 C.7 Proof of Proposition 4

To begin with, we consider \(\varvec{x}=w(1,1,1,1,1,1)\). Then (19) reduces to a single condition \(wR(w,w,w,w,w,w,\psi )=0\). Since \(w=0\) corresponds to the pre-bifurcation solution, we focus on a relation \(R(w,w,w,w,w,w,\psi )=0\). Because this relation in general has a solution of the form \(\psi = a w +\mathrm{(higher \ order \ terms)}\) for some real constant a in the neighborhood of the bifurcation point, the bifurcating solution of the form \(\varvec{x}=w(1,1,1,1,1,1)\) exists. The other five cases can be treated similarly.

1.8 C.8 The Detail of the Bifurcation Equation for the Orbit of Type \(\beta i\)

Using the following Lemma, we can obtain the bifurcating solutions for Type \(\beta i\) (i.e., Proposition 5).

Lemma 4

Bifurcation equation for Type \(\beta i\) orbit is derived as follow: some i, which comprises twelve points, the bifurcation equation becomes:

$$\begin{aligned} \tilde{\varvec{F}}_1(\varvec{x},\psi )= & {} x_1 R(x_1,x_2,x_3,x_4,x_5,x_6,x_7,x_8,x_9,x_{10},x_{11},x_{12},\psi )=0, \\ \tilde{\varvec{F}}_2(\varvec{x},\psi )= & {} x_2 R(x_2,x_1,x_{12},x_{11},x_{10},x_9,x_8,x_7,x_6,x_5,x_4,x_3,\psi )=0, \\ \tilde{\varvec{F}}_3(\varvec{x},\psi )= & {} x_3 R(x_3,x_4,x_5,x_6,x_7,x_8,x_9,x_{10},x_{11},x_{12},x_1,x_2,\psi )=0, \\ \tilde{\varvec{F}}_4(\varvec{x},\psi )= & {} x_4 R(x_4,x_3,x_2,x_1,x_{12},x_{11},x_{10},x_9,x_8,x_7,x_6,x_5,\psi )=0, \\ \tilde{\varvec{F}}_5(\varvec{x},\psi )= & {} x_5 R(x_5,x_6,x_7,x_8,x_9,x_{10},x_{11},x_{12},x_1,x_2,x_3,x_4,\psi )=0, \\ \tilde{\varvec{F}}_6(\varvec{x},\psi )= & {} x_6 R(x_6,x_5,x_4,x_3,x_2,x_1,x_{12},x_{11},x_{10},x_9,x_8,x_7,\psi )=0, \\ \tilde{\varvec{F}}_7(\varvec{x},\psi )= & {} x_7 R(x_7,x_8,x_9,x_{10},x_{11},x_{12},x_1,x_2,x_3,x_4,x_5,x_6,\psi )=0, \\ \tilde{\varvec{F}}_8(\varvec{x},\psi )= & {} x_8 R(x_8,x_7,x_6,x_5,x_4,x_3,x_2,x_1,x_{12},x_{11},x_{10},x_9,\psi )=0, \\ \tilde{\varvec{F}}_9(\varvec{x},\psi )= & {} x_9 R(x_9,x_{10},x_{11},x_{12},x_1,x_2,x_3,x_4,x_5,x_6,x_7,x_8,\psi )=0, \\ \tilde{\varvec{F}}_{10}(\varvec{x},\psi )= & {} x_{10} R(x_{10},x_9,x_8,x_7,x_6,x_5,x_4,x_3,x_2,x_1,x_{12},x_{11},\psi )=0, \\ \tilde{\varvec{F}}_{11}(\varvec{x},\psi )= & {} x_{11} R(x_{11},x_{12},x_1,x_2,x_3,x_4,x_5,x_6,x_7,x_8,x_9,x_{10},\psi )=0, \\ \tilde{\varvec{F}}_{12}(\varvec{x},\psi )= & {} x_{12} R(x_{12},x_{11},x_{10},x_9,x_8,x_7,x_6,x_5,x_4,x_3,x_2,x_1,\psi )=0, \end{aligned}$$

where \(\varvec{x}=(x_1,\ldots ,x_{12})=\{\lambda _i \mid i \in \beta i\}\) (Fig. 8(b)) and \(\psi =\phi -\phi ^\textrm{c}_l\).

Proof

By the product form of the replicator dynamics in (3), \(\tilde{F}_i(\varvec{x},\psi )\) takes the product form: \(\tilde{F}_i(\varvec{x},\psi )=x_i R_i(\varvec{x},\psi )\) \((i\in P_{\beta i})\). Since the group \(\textrm{D}_6\) is generated by the elements r and s, it suffice to consider the symmetry condition \(T(g) \tilde{\varvec{F}}(\varvec{x})=\tilde{\varvec{F}}(T(g)\varvec{x})\) for \(g=r,s\). This condition for \(g=r\) gives the form for \(\tilde{\varvec{F}}_{i}\) \((i=1,3,\ldots ,11)\) for some function \(R(\varvec{x},\psi )\) and that for \(g=s\) gives the form for \(\tilde{\varvec{F}}_{i}\) \((i=2,4,\ldots ,12)\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aizawa, H., Ikeda, K. & Kogure, Y. Satellite City Formation for a Spatial Economic Model. Netw Spat Econ 23, 529–558 (2023). https://doi.org/10.1007/s11067-023-09586-z

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11067-023-09586-z

Keywords

Navigation