Skip to main content
Log in

When Veblen meets Krugman: social network and city dynamics

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

The present paper explores the role of the social structure of suburban areas on city dynamics. We focus on relative concerns in the form of conspicuous consumption and introduce them into a standard economic geography model a la Krugman. We show that the level of social integration within the suburban areas of cities and the level of economic integration across cities are crucial in determining the city sizes. An interesting case arises with moderate trade costs when relatively small shares of income are devoted to the consumption of the differentiated good: if classes of workers are segregated (as in homogenous suburban areas), relative concerns tend to generate dispersed, medium-size cities; when workers of different classes socially interact, relative concerns contribute to foster socially integrated megalopolises. This result shows that keeping-up-with-the-Joneses motives may generate counterintuitive results when agents are able to choose their location.

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

Similar content being viewed by others

Notes

  1. See Borjas (1994) for a discussion on several aspects of the economics of immigration.

  2. For example, Quinn (2006) shows “relative deprivation” is a significative factor in domestic migration decisions among Mexicans.

  3. We choose the model by Forslid and Ottaviano (2003) because of its analytical tractability and because it turns out to be isomorphic to the core-periphery model by Krugman (1991).

  4. See Fujita et al. (1999), Fujita and Thisse (2002) and Baldwin et al. (2003) for a literature review on new economic geography models.

  5. There is also a vast theoretical literature on relative concerns that has been developed since the work of Veblen (1899). Among the many, see e.g., Abel (1990), Frey and Stutzer (2002), Hopkins and Kornienko (2004), Arrow and Dasgupta (2007) and Ghiglino and Goyal (2010).

  6. See, for example, Blanchflower and Oswald (2004), Layard (2005), Luttmer (2005) and Kuhn et al. (2011).

  7. See Saramaki et al. (2014), Eagle et al. (2009) and Wesolowski et al. (2013).

  8. We recall that in Forslid and Ottaviano (2003) there are two centrifugal forces, which are the so-called market size effect and the cost-of-living effect, and one centrifugal force, that is the market crowding effect. We define these forces as the “traditional” or “original” forces.

  9. As it will be specified later, a regional network is complete when in each given region, each worker compares his/her consumption with the average consumption of all the agents in the region weighted by the measure of the comparison group. Moreover, a regional network is regular when all agents of a given type occupy equivalent positions in the network and therefore can be treated as identical.

  10. For a review on the literature on social interactions and urban economics, see Ioannides (2013, chapter 5).

  11. The literature includes works without location choice, such as those by Johnson and Gilles (2000), Galeotti et al. (2006), Carayol et al. (2008).

  12. Similar segregation issues that may arize due to social network are analysed by de Marti and Zenou (2012) in a framework that does not consider an explicit spatial framework.

  13. Note that this formulation appears to be the interesting case to consider; see Frank (1985) for a discussion on differences in social sensitiveness across goods and Kuhn et al. (2011) for evidence that relative consumption effects are prominent with some goods but not others. It is also natural to assume that if consumption externalities are symmetric for both goods, then the social comparisons will simply wash out and equilibrium will be analogous to the equilibrium in an economy with no consumption externalities [see Arrow and Dasgupta (2007)].

  14. The first systematic analysis of the size effect in relative concerns was done in Ghiglino and Goyal (2010). More recently, Liu et al. (2011) also consider this issue.

  15. We focus on these two polar cases because they help us clarify the types of effects at work; however, we recognize that the intermediate specification with a small size effect may be more realistic. In this paper, we have mainly explored the role of relative concerns in location decision and city dynamics when the size of the comparison group matters. In the working paper Ghiglino and Nocco (2012), we explore the somewhat more standard pure average specification and precisely show where the change in specification affects the results. In fact, most of the results are not affected by the change and the broad picture remains unaffected.

  16. If \(\Lambda _{r}(i)=0\) then obviously \(\varPhi (X_{ir},X_{-ir})=X_{ir}\).

  17. Forslid and Ottaviano (2003) assume that \(L_{r}=L/2\) and \(H=1\).

  18. When performing the numerical analysis, we choose values of \(\alpha \) sufficiently small for the given values of the other parameters that ensure that \(W_{\alpha ir}\) is positive.

  19. Moreover, if we define \(E_{\Lambda _{r}(i)}\equiv p_{X_{r}}\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j\) as the total expenditure on the differentiated good by neighbours of individual i in region r, we can also point out what follows: for given wages \(w_{ir}\) and total expenditures of neighbours of individual i in region r on the differentiated good \(E_{\Lambda _{r}(i)}\), an increase in conspicuous consumption effects (that is an increase in \( \alpha \)) reduces the share of wage used by individuals to buy the traditional good, \(\frac{A_{ir}}{w_{ir}}=\left( 1-\mu \right) \left( 1-\frac{ \alpha S(\Omega (\Lambda _{r}(i)))}{1+\alpha S(\Omega (\Lambda _{r}(i)))} \frac{E_{\Lambda _{r}(i)}}{w_{ir}\int _{\Lambda _{r}(i)}\mathrm{d}j}\right) \), and, consequently, increases the share of wage used to buy the manufactured good.

  20. Alternatively, we could examine the location decision of skilled workers studying the current indirect utility differential, as in Forslid and Ottaviano (2003). However, we choose to work with the logarithm of the ratio of the current indirect utility levels because it gives an expression that is analytically more tractable, which, in any case, identifies the same critical values of \(\phi \) for the sustainability of the different types of equilibria.

  21. Of course the numerical examples would be affected.

  22. To have an insight of what happens, we consider the following thought experiment. For given values of \(\alpha \), of the wages and of the price indexes—which imply a given value of \(\left( n_{r}+n_{v}\phi \right) \), we evaluate the following expression

    $$\begin{aligned}&\tfrac{\partial X_{rr}(s)}{\partial H_{r}}-\left( \left. \tfrac{\partial X_{rr}(s)}{\partial H_{r}}\right| _{\alpha =0}\right) \\&\quad =\tfrac{\partial \left( \mu \frac{\left( \sigma -1\right) \left( L+H_{r}w_{H_{r}}\right) }{\sigma \beta \left( n_{r}+n_{v}\phi \right) }\frac{ 1+\alpha \left( H_{r}+L\right) }{1+\alpha \mu \left( H_{r}+L\right) }\right) }{\partial H_{r}}-\left( \left. \tfrac{\partial \left( \mu \frac{\left( \sigma -1\right) \left( L+H_{r}w_{H_{r}}\right) }{\sigma \beta \left( n_{r}+n_{v}\phi \right) }\right) }{\partial H_{r}}\right| _{\alpha =0}\right) \\&\quad =\alpha \mu \left( 1-\mu \right) \left( \sigma -1\right) \tfrac{L+2w_{H_{r}}H_{r}+Lw_{H_{r}}+w_{H_{r}}\alpha \mu \left( L+H_{r}\right) ^{2}}{\sigma \beta \left( L\alpha \mu +\alpha \mu H_{r}+1\right) ^{2}\left( n_{r}+\phi n_{v}\right) }>0 \end{aligned}$$
  23. Indeed, at the symmetric equilibrium (with \(H_{r}=H/2\) and \(\tfrac{\left( L+H_{r}w_{H_{r}}\right) }{H_{r}+\left( H-H_{r}\right) \phi }=\tfrac{\left( L+H_{v}w_{H_{v}}\right) }{\left( H-H_{r}\right) +H_{r}\phi }\)), for given total expenditures and price indexes (that imply given values of \(\left( L+H_{r}w_{H_{r}}\right) \) and \(H_{r}+\left( H-H_{r}\right) \phi \)), the sign of \(\frac{\partial w_{H_{r}}}{\partial H_{r}}\) with \(w_{H_{r}}= \frac{\mu }{\sigma }\tfrac{\left( L+H_{r}w_{H_{r}}\right) }{H_{r}+\left( H-H_{r}\right) \phi }\left[ \frac{1+\alpha \left( H_{r}+L\right) }{1+\alpha \mu \left( H_{r}+L\right) }+\phi \frac{1+\alpha \left( H-H_{r}+L\right) }{ 1+\alpha \mu \left( H-H_{r}+L\right) }\right] \) is equal to the sign of \( \frac{\partial \left( \frac{1+\alpha \left( H_{r}+L\right) }{1+\alpha \mu \left( H_{r}+L\right) }+\phi \frac{1+\alpha \left( H-H_{r}+L\right) }{ 1+\alpha \mu \left( H-H_{r}+L\right) }\right) }{\partial H_{r}}=\alpha \left( 1-\mu \right) \frac{\left( H\alpha \mu +2L\alpha \mu +2\right) ^{2}\left( 1-\phi \right) }{4\left( L\alpha \mu +\alpha \mu H_{r}+1\right) ^{2}\left( H\alpha \mu +L\alpha \mu -\alpha \mu H_{r}+1\right) ^{2}}>0\) which is positive when \(\alpha >0\). Let us notice that the extent of this increase is larger when \(\phi \) is low and that it becomes smaller when \( \phi \) increases.

  24. An analogous expression can be obtained for the wage of unskilled workers net of the conspicuous effect \(W_{\alpha L_{r}}\). Specifically, we can substitute the total amount of consumptions by the neighbours of unskilled workers \(\int _{\Omega _{r}(L)}X_{jr}\mathrm{d}j=X_{H_{r}}H_{r}+X_{L_{r}}L\) in the definition of \(W_{\alpha L_{r}}\) in (9) and obtain that

    $$\begin{aligned} W_{\alpha L_{r}}\equiv 1-\alpha \mu \tfrac{L+H_{r}w_{H_{r}}}{1+\alpha \mu \left( L+H_{r}\right) } \end{aligned}$$

    This expression is important to check the positivity of agricultural demand by unskilled workers in (7).

  25. This is true when the skilled wage is above a minimum level smaller than 1, that is for \(w_{H_{r}}{>}L\alpha \mu /(1+L\alpha \mu )\).

  26. We assume that the no black hole condition, that is \(\mu <\sigma -1\), holds. This condition rules out the case in which the symmetric equilibrium is never stable when relative concerns are absent.

  27. In “Complete networks” section of Appendix, we point out that \(H<2\frac{\left( 1+L\alpha \right) }{ \alpha (\sigma -1)}\) is required to have a positive value of the wage net of the conspicuous effect for skilled workers in r in the symmetric equilibrium, that is \(W_{\alpha H_{r}}(H/2)>0\).

  28. Note that the intervals of costs associated with full agglomeration equilibria and with unstable symmetric equilibria may be different, as seen in the discussion below.

  29. This numerical analysis is performed with \(L=4\), \(H=10\), \(\sigma =2\) and \( \mu =0.11\).

  30. Specifically, the symmetry breaking point \(\phi _{b}\ \)is such that the symmetric equilibrium is stable for \(\phi \in (0,\phi _{b})\), and the sustain point \(\phi _{s}\) is the critical value of \(\phi \) such that full agglomeration is an equilibrium for \(\phi >\phi _{s}\).

  31. Notice that the same type of graphic in Fig. 1c is obtained for other values of alpha, that is \(\alpha =0.06\), \(\alpha =0.08\), \(\alpha =0.10\), \( \alpha =0.12\), \(\alpha =0.14\) and \(\alpha =0.16\), with the corresponding critical values given in Table 1 in “Complete networks” section of Appendix that are such that \(\phi _{b}<\phi _{s}<\phi _{d}<\phi _{u}\).

  32. For other models in which catastrophic agglomeration is replaced with gradual and partial agglomeration processes, see, e.g., Helpman (1998), Tabuchi (1998), Ludema and Wooton (1999), Ottaviano et al. (2002), Tabuchi and Thisse (2002), Murata (2003), Pflüger (2004), Nocco (2009), Berliant and Kung (2009), Pflüger and Suedekum (2011) and Ottaviano (2012).

  33. Not surprisingly, the masses H and L are both relevant for the stability of equilibria in the case of the complete network, while only H is relevant for the segregated network. Note that they are not relevant in the model with no relative concerns of Forslid and Ottaviano (2003).

  34. We notice that for the two cases represented in Fig. 2a, b the condition \(H<2\frac{\left( 1+L\alpha \right) }{\alpha (\sigma -1)}\), which is assumed in Proposition 3, holds.

  35. See, for instance, Tabuchi (1998), Tabuchi and Thisse (2006) and Gaigné and Thisse (2014). Basically, the urban structure of these models disappears when commuting costs are equal to zero.

  36. An analogous expression to (43) holds for \(w_{H_{v}}\).

  37. Specifically, we observe that \(A\left( H_{r}=H_{v}=H/2\right) \)=\(B\left( H_{r}=H_{v}=H/2\right) \)=\(\frac{\alpha \Omega _{s}+1}{\alpha \mu \Omega _{s}+1}\).

  38. This requires that the relative mass of skilled workers with respect to the mass of unskilled is not relatively too high, that is, \(H<2\frac{\left( 1+L\alpha \right) }{\alpha (\sigma -1)}\).

  39. We assume that the no black hole condition, that is \(\mu <\sigma -1\), holds. This condition rules out the case in which the symmetric equilibrium is never stable.

  40. We recall that we assume \(H<2\frac{\left( 1+L\alpha \right) }{\alpha (\sigma -1)}\) to have \(W_{\alpha H_{r}}(H/2)>0\).

  41. See expression (25) at page 236 in Forslid and Ottaviano (2003).

  42. We recall that we assume the no black hole condition corresponding to the case of \(\alpha =0\) holds, that is \(\mu <\sigma -1\).

  43. Moreover, its minimum value is attained at \(\phi =-\frac{d_{1}}{2d_{2}}>0\) when \(\alpha >0\) (and at \(\phi =0\) when \(\alpha =0\)), so that the parabola has a negative slope in \(\phi =0\) only if \(\alpha \) is positive.

  44. This requires that \(H<2\frac{\left( 1+L\alpha \right) }{\alpha (\sigma -1)}\).

  45. When \(\alpha =0\): the numerator is still an increasing function in \(\phi \) taking the value 0 when \(\phi =0\), but it is smaller for any other value of \(\phi \); the parabola is always increasing for \(\phi \in [0,1]\,\), it intersects the vertical axis at a lower value of \(d_{0}\), that is \(\ \sigma -\mu \), and it is equal to the numerator (\(2\sigma \)) when \(\phi =1\).

  46. Indeed, we know that when \(\phi =1\) the denominator is

    $$\begin{aligned} d_{2}\phi ^{2}+d_{1}\phi +d_{0}=\sigma \left( H\alpha \mu +L\alpha \mu +1\right) \left( H\alpha +2L\alpha -H\alpha \sigma +2\right) \end{aligned}$$

    while the numerator is

    $$\begin{aligned} g_{0}\left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} \phi ^{1-\frac{ \mu }{\sigma -1}}=\sigma \left( 1+\alpha \mu L\right) \left[ \tfrac{\alpha (H+L)+1}{\alpha L+1}\right] ^{\mu }\left\{ 2+\alpha \left[ 2L-H(\sigma -1) \right] \right\} \end{aligned}$$

    Hence, when \(\phi =1\), the argument of the logarithm in (46) is given by \(\left[ \tfrac{\alpha (H+L)+1}{\alpha L+1}\right] ^{\mu }\frac{\left( 1+\alpha \mu L\right) }{\left( H\alpha \mu +L\alpha \mu +1\right) }\), which is equal to 1 when \(\alpha =0\). With a positive value of \(\alpha \), we can show that \(\left[ \tfrac{\alpha (H+L)+1}{\alpha L+1}\right] ^{\mu }\frac{ \left( 1+\alpha \mu L\right) }{\left( H\alpha \mu +L\alpha \mu +1\right) }<1\) and, thus, agglomeration is never an equilibrium when \(\phi =1\).

    Proof. \(\left[ \tfrac{\alpha (H+L)+1}{\alpha L+1}\right] ^{\mu }\frac{\left( 1+\alpha \mu L\right) }{\left( H\alpha \mu +L\alpha \mu +1\right) }<1\) requires that \(\left[ \tfrac{\alpha (H+L)+1}{\alpha L+1}\right] ^{\mu }< \frac{H\alpha \mu +L\alpha \mu +1}{1+\alpha \mu L}\). We know that these two expressions, defined respectively as \(LHS=\left[ \tfrac{\alpha (H+L)+1}{ \alpha L+1}\right] ^{\mu }\) and \(RHS=\frac{H\alpha \mu +L\alpha \mu +1}{ 1+\alpha \mu L}\), are both equal to 1 when \(\mu =0\), and that they both increase in the range \(\mu \in [0,1]\) and assume the same value \( \frac{\alpha (H+L)+1}{\left( 1+\alpha L\right) }\) when \(\mu =1\). However, since \(\forall \mu \in (0,1)\) the \(\frac{\partial ^{2}LHS}{\partial \mu ^{2}} >0\) and \(\frac{\partial ^{2}RHS}{\partial \mu ^{2}}<0\), the LHS is convex in \(\mu \) and the RHS is concave in \(\mu \), which implies that \(LHS<RHS\). Q.E.D.

  47. The two critical points \(\phi _{s}\) and \(\phi _{u}\) in Fig. 3 correspond to those obtained in Table 1 for \(\alpha =0.16\) when \(L=4\), \(H=10\), \( \sigma =2\) and \(\mu =0.11\).

  48. The values of the parameters for which curves in Fig. 4a, b are drawn are equal to those used to derive Fig. 5a, b.

  49. The choice of a smaller value of L and \(\mu =0.11\) allows for the asymmetric stable equilibria when \(\alpha >0.04\).

  50. This result can be compared with the case in which \(\alpha =0\) that implies that: the numerator is still an increasing function in \(\phi \) taking the value 0 when \(\phi =0\) but it is smaller for any other value of \(\phi \in [0,1]\); the parabola is always increasing for \(\phi \in [0,1]\, \), it intersects the vertical axis at a lower value, that is\(\ \sigma -\mu \), and it is equal to the numerator (\(2\sigma \)) when \(\phi =1\).

  51. Specifically,

    $$\begin{aligned} k_{2}= & {} \alpha ^{2}H^{2}\mu \left( \sigma +1\right) \left[ \sigma \left( \mu -2\right) +2\left( 1-\mu \right) \right] -2\alpha H\left[ \sigma \left( \sigma +\sigma \mu -1\right) +3\mu \left( \sigma +\mu -1\right) \right] \\&-4\left( \sigma +\mu -1\right) \left( \sigma +\mu \right) ; \\ k_{1}= & {} 2\alpha ^{2}H^{2}\mu ^{2}\left[ \sigma \left( \sigma -1\right) +1 \right] +4\alpha H\left[ \sigma \left( 1+\mu \right) \left( \sigma -1\right) +2\mu ^{2}\right] +8\left[ \sigma \left( \sigma -1\right) +\mu ^{2}\right] >0; \\ k_{0}= & {} \alpha ^{2}H^{2}\mu \left( \sigma -1\right) \left[ \sigma \mu -2\left( \sigma -1\right) \right] -2\alpha H\left[ \mu \left( \mu +3\right) +\sigma ^{2}\left( \mu +1\right) -\sigma \left( 5\mu +1\right) \right] \\&-4\left( \sigma -\mu -1\right) \left( \sigma -\mu \right) ; \\ h_{1}= & {} 2\left( \sigma +\mu \right) +\mu \alpha \left( \sigma +1\right) H>0;\\ h_{0}= & {} 2\left( \sigma -\mu \right) +\mu \alpha \left( \sigma -1\right) H>0. \end{aligned}$$
  52. Specifically, from (4), (9) and (10), we know that \(\mu =\frac{p_{X_{r}}\left( X_{ir}-\frac{\alpha S(\Omega (\Lambda _{r}(i)))}{1+\alpha S(\Omega (\Lambda _{r}(i)))}\frac{\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j}{\int _{\Lambda _{r}(i)}\mathrm{d}j}\right) }{W_{\alpha ir}}\).

  53. Where \(2+\alpha \left[ 2L-H(\sigma -1)\right] \) in the numerator has to be positive to have a positive value of \(W_{\alpha Hr}\) when \(H_{r}=H\). All the other factors in the numerator and denominator are positive.

  54. Proof. The argument of the logarithm in \(V^{mix}\left( H,1,\alpha \right) \) is smaller than 1 if \(\frac{\left[ H\alpha +2L\alpha +2+\alpha \mu \left( H+2L+2L^{2}\alpha +2HL\alpha \right) \right] }{\left( 1+\alpha \mu L\right) \left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} }<\left[ \alpha (H+L)+1\right] ^{\mu }\). We know that these two expressions, defined as \(LHS= \frac{\left[ H\alpha +2L\alpha +2+\alpha \mu \left( H+2L+2L^{2}\alpha +2HL\alpha \right) \right] }{\left( 1+\alpha \mu L\right) \left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} }\) and \(RHS=\left[ \alpha (H+L)+1 \right] ^{\mu }\), are respectively equal to \(\frac{H\alpha +2L\alpha +2}{ H\alpha +2L\alpha -H\alpha \sigma +2}>1\) and to 1 when \(\mu =0\), and that they both increase in the range \(\mu \in [0,1]\) and assume respectively the value \(2\frac{H\alpha +L\alpha +1}{H\alpha +2L\alpha -H\alpha \sigma +2}\) and \(H\alpha +L\alpha +1\) when \(\mu =1\). If \( H>2L/(\sigma -1)\) we find always that \(LHS>RHS\) for \(\mu \in [0,1]\) and therefore full agglomeration in r with \(\phi =1\) is never an equilibrium. However, if \(H<2L/(\sigma -1)\) full agglomeration in r with \( \phi =1\) is an equilibrium for relatively large value of \(\mu \) that ensure that \(LHS<RHS\). Q.E.D.

References

  • Abel, A.: Asset prices under habit formation and catching up with Jonesses. Am. Econ. Rev. 80(2), 38–42 (1990)

    Google Scholar 

  • Arrow, K., Dasgupta, P.: Conspicuous consumption; inconspicuous leisure, Mimeo. Stanford University and Cambridge University (2007)

  • Baldwin, R., Forslid, R., Martin, P., Ottaviano, G., Robert-Nicoud, F.: Economic Geography and Public Policy. Princeton University Press, Princeton (2003)

    Google Scholar 

  • Berliant, M., Kung, F.-C.: Bifurcations in regional migration dynamics. Reg. Sci. Urban Econ. 39, 714–720 (2009)

    Article  Google Scholar 

  • Bettencourt, L.M.A.: The origins of scaling in cities. Science 340(6139), 1438–1441 (2013). doi:10.1126/science.1235823

    Article  Google Scholar 

  • Blanchflower, D., Oswald, A.: Well-being over time in Britain and the USA. J. Public Econ. 88, 1359–1386 (2004)

    Article  Google Scholar 

  • Borjas, G.J.: The economics of immigration. J. Econ. Lit. 32, 1667–1717 (1994)

    Google Scholar 

  • Carayol, N., Roux, P., Yıldızo \(\breve{\text{ G }}\)lu, M.: Inefficiencies in a model of spatial networks formation with positive externalities. J. Econ. Behav. Organ. 67, 495-511 (2008)

  • De Marti, J., Zenou, Y.: Identity and social distance in friendship formation. Research Papers in Economics, Stockholm University, Department of Economics (2012)

  • Duranton, G.: Spatial economics. In: Durlauf, S. N., Blume, L. E. (eds.) The New Palgrave Dictionary of Economics, Second Edition (2008). The New Palgrave Dictionary of Economics Online. Palgrave Macmillan. 21 Oct 2010 doi:10.1057/9780230226203.1583. http://www.dictionaryofeconomics.com/article?id=pde2008_S000195

  • Eagle, N., Pentland, A., Lazer, D.: Inferring friendship structure by using mobile phone data. Proc. Natl. Acad. Sci. 106, 15274–15278 (2009)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Frank, R.: Choosing the Right Pond. Oxford University Press, Oxford (1985)

    Google Scholar 

  • Frey, B., Stutzer, A.: What can economists learn from happiness research? J. Econ. Lit. XL, 402–435 (2002)

    Article  Google Scholar 

  • Fujita, M., Krugman, P., Venables, A.J.: The spatial economy: cities, regions and international trade. MIT Press, Cambridge (1999)

    Google Scholar 

  • Fujita, M., Thisse, J.-F.: Economics of Agglomeration. Cities, Industrial Location, and Regional Growth. Cambridge University Press, Cambridge (2002)

    Book  Google Scholar 

  • Gaigné, C., Thisse, J.-F.: New economic geography: past and future. In: Fischer, M.M., Nijkamp, P. (eds.) Handbook of Regional Science, pp. 539–568. Springer, Berlin (2014)

    Chapter  Google Scholar 

  • Galeotti, A., Goyal, S., Kamphorst, J.: Network formation with heterogeneous players. Games Econ. Behav. 54, 353–372 (2006)

    Article  Google Scholar 

  • Ghiglino, C., Goyal, S.: Keeping up with the neighbors: social interaction in a market economy. J. Eur. Econ. Assoc. 8, 90–119 (2010)

    Article  Google Scholar 

  • Ghiglino, C., Nocco, A.: When Veblen meets Krugman. University of Essex Discussion Paper Series No 708 (2012)

  • Helpman, E.: The size of regions. In: Pines, D., Sadka, E., Zilcha, I. (eds.) Topics in Public Economics. Theoretical and Empirical Analysis, pp. 33–54. Cambridge University Press, Cambridge (1998)

    Google Scholar 

  • Henderson, J.V.: The sizes and types of cities. Am. Econ. Rev. 64, 640–656 (1974)

    Google Scholar 

  • Hopkins, E., Kornienko, T.: Running to keep in the same place: consumer choice as a game of status. Am. Econ. Rev. 94(4), 1085–1107 (2004)

    Article  Google Scholar 

  • Helsley, R.W., Strange, W.C.: Urban interactions and spatial structure. J. Econ. Geogr. 7, 119–138 (2007)

    Article  Google Scholar 

  • Helsley, R.W., Zenou, Y.: Social networks and interactions in cities. J. Econ. Theory 150, 426–466 (2014)

    Article  Google Scholar 

  • Ioannides, Y.: From Neighborhoods to Nations. The Economics of Social Interactions. Princeton University Press, Princeton (2013)

    Google Scholar 

  • Johnson, C., Gilles, R.P.: Spatial social networks. Rev. Econ. Design 5, 273–299 (2000)

    Article  Google Scholar 

  • Kuhn, P., Kooreman, P., Soetevent, A., Kapteyn, A.: The effects of lottery prizes on winners and their neighbors: evidence from the Dutch postcode lottery. Am. Econ. Rev. 101(5), 2226–2247 (2011)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Layard, R.: Happiness: Lessons from a New Science. Penguin Books, London (2005)

    Google Scholar 

  • Liu, X., Patacchini, E., Zenou, Y.: Peer effects in education, sport, and screen activities: local aggregate or local average?. CEPR Discussion Paper No. 8477 (2011)

  • Ludema, R., Wooton, I.: Regional Integration, Trade and Migration: Are Demand Linkages Relevant in Europe? In: Faini, R., De Melo, J., Zimmermann, K.F. (eds.) Migration. The Controversies and the Evidence, pp. 51–75. CEPR and Cambridge University Press, Cambridge (1999)

    Chapter  Google Scholar 

  • Luttmer, E.F.P.: Neighbors as negatives: relative earnings and well-being. Q. J. Econ. 120(3), 963–1002 (2005)

    Google Scholar 

  • Mossay, P., Picard, P.M.: On spatial equilibria in a social interaction model. J. Econ. Theory 146, 2455–2477 (2011)

    Article  Google Scholar 

  • Mossay, P., Picard, P.M.: Spatial segregation and urban structure. RIETI Discussion Paper Series 13-E-056 (2013)

  • Murata, Y.: Product diversity, taste heterogeneity, and geographic distribution of economic activities: market vs. non-market interactions. J. Urban Econ. 53, 126–144 (2003)

    Article  Google Scholar 

  • Nocco, A.: Preference heterogeneity and economic geography. J. Reg. Sci. 49, 33–56 (2009)

    Article  Google Scholar 

  • Ottaviano, G.I.P., Tabuchi, T., Thisse, J.-F.: Agglomeration and trade revisited. Int. Econ. Rev. 43, 409–435 (2002)

    Article  Google Scholar 

  • Ottaviano, G.I.P.: Agglomeration, trade and selection. Reg. Sci. Urban Econ. 42, 987–997 (2012)

    Article  Google Scholar 

  • Pflüger, M.: A simple, analytically solvable, chamberlinian agglomeration model. Reg. Sci. Urban Econ. 34, 565–573 (2004)

    Article  Google Scholar 

  • Pflüger, M., Suedekum, J.: On pitchforks and tomahawks. J. Reg. Sci. 51, 292–298 (2011)

    Article  Google Scholar 

  • Quinn, M.A.: Relative deprivation, wage differentials and Mexican migration. Rev. Dev. Econ. 10, 135–153 (2006)

    Article  Google Scholar 

  • Radu, D.: Social interactions in economic models of migration: a review and appraisal. J. Ethnic Migr. Stud. 34, 531–548 (2008)

    Article  Google Scholar 

  • Saramaki, J., Leicht, E.A., Lopez, E., Roberts, S.G.B., Reed-Tsochas, F., Dunbar, R.I.M.: Persistence of social signatures in human communication. Proc. Natl. Acad. Sci. 111, 942–947 (2014)

    Article  Google Scholar 

  • Schelling, T.C.: Dynamic models of segregation. J. Math. Sociol. 1, 143–186 (1971)

  • Schläpfer, M., Bettencourt, L.M.A., Grauwin, S., Raschke, M., Claxton, R., Smoreda, Z., West, G.B., Ratti, C.: The scaling of human interactions with city size. J. R. Soc. Interface (2014). doi:10.1098/rsif.2013.0789

    Google Scholar 

  • Stark, O., Taylor, J.E.: Migration incentives, migration types: the role of relative deprivation. Econ. J. 101, 1163–1178 (1991)

    Article  Google Scholar 

  • Tabuchi, T.: Urban agglomeration and dispersion: a synthesis of Alonso and Krugman. J. Urban Econ. 44, 333–351 (1998)

    Article  Google Scholar 

  • Tabuchi, T., Thisse, J.-F.: Taste heterogeneity, labor mobility and economic geography. J. Dev. Econ. 69, 155–177 (2002)

    Article  Google Scholar 

  • Tabuchi, T., Thisse, J.-F.: Regional specialization, urban hierarchy, and commuting costs. Int. Econ. Rev. 47, 1295–1317 (2006)

    Article  Google Scholar 

  • Veblen, T.: The Theory of the Leisure Class. Macmillan, New York (1899)

    Google Scholar 

  • Wesolowski, A., Eagle, N., Noor, A.M., Snow, R.W., Buckee, C.O.: The impact of biases in mobile phone ownership on estimates of human mobility. J. R. Soc. Interface (2013). doi:10.1098/rsif.2012.0986

    Google Scholar 

Download references

Acknowledgments

We are grateful to Gianmarco Ottaviano, Jacques Thisse and participants at the Conference “New challenges for macroeconomic regulation: financial crisis, stabilization policy and sustainable development” (GREQAM Marseille, 9–11 June 2011), 11th SAET Conference (Ancao (Faro), Portugal, 26 June–1 July 2011), SED Annual Meeting (Ghent, Belgium, 7–9 July 2011), ETSG Conference 2011 (Copenhagen Business School, 8–10 September 2011), 58th Annual North American Meetings of the Regional Science Association International (Miami, 9–12 November 2011), ITSG Meeting “International Trade, Finance and Migration” (Roma, University La Sapienza, 16–17 February 2012), XVII DEGIT (Dynamics, Economic Growth, International Trade) Conference, University of Milan, 13-14 September 2012, III International conference “Industrial Organization and Spatial Economics” (Center for Market Studies and Spatial Economics and National Research University Higher School of Economics, Saint-Petersburg, August 25–26, 2014) and Crenos seminar (Università di Cagliari) for helpful comments. The usual disclaimer applies.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Ghiglino.

Appendices

Appendix 1

1.1 Demand side and price index

In this Appendix, we compute the consumer’s demand for the differentiated good \(X_{ir}\) and for the agricultural good \(A_{ir}\) in region r, the indirect utility function \(U(A_{ir},\varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) )\), the individual demand in region r for variety v (\( X_{ir}(v)\)) and we define the price index \(p_{X_{r}}\).

Each individual i in region r solves the program

$$\begin{aligned}&\mathrm{Max}_{A_{ir},X_{ir}}U(A_{ir},\varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) )=A_{ir}^{1-\mu }\left( \varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) \right) ^{\mu }\\&\mathrm{s.t.}\qquad A_{ir}+\int \limits _{s\in N}p_{r}(s)X_{ir}(s)ds=w_{ir} \end{aligned}$$

with

$$\begin{aligned} \varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) =X_{ir}+\alpha S(\Omega (\Lambda _{r}(i)))\left[ X_{ir}-\frac{\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j}{ \int _{\Lambda _{r}(i)}\mathrm{d}j}\right] \end{aligned}$$

As \(\int _{\Lambda _{r}(i)}\mathrm{d}j=\Omega (\Lambda _{r}(i))\) and \(X_{ir}=\left( \displaystyle \int \limits _{s\in N}X_{ir}(s)^{\frac{\sigma -1}{\sigma }}ds\right) ^{\frac{ \sigma }{\sigma -1}}\), the Lagrangean function is

$$\begin{aligned} L= & {} A_{ir}^{1-\mu }\left[ \left[ 1+\alpha S(\Omega (\Lambda _{r}(i)))\right] \left( \int \limits _{s\in N}X_{ir}(s)^{\frac{\sigma -1}{\sigma }}ds\right) ^{ \frac{\sigma }{\sigma -1}}\right. \\&\qquad \qquad \left. -\alpha S(\Omega (\Lambda _{r}(i)))\frac{ \int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j}{\Omega (\Lambda _{r}(i))} \right] ^{\mu } \\&\quad +\,\lambda \left( w_{ir}-A_{ir}-\int \limits _{s\in N}p_{r}(s)X_{ir}(s)ds\right) \end{aligned}$$

with the first-order conditions with respect to \(A_{ir}\), the consumption of variety s, \(X_{ir}(s)\), and of variety v, \(X_{ir}(v)\), and of \(\lambda \) , respectively, given by

$$\begin{aligned}&\left( 1-\mu \right) A_{ir}^{-\mu }\varPhi ^{\mu }=\lambda \end{aligned}$$
(29)
$$\begin{aligned}&\mu A_{ir}^{1-\mu }\varPhi ^{\mu -1}\left[ 1+\alpha S(\Omega (\Lambda _{r}(i))) \right] \left( x\right) ^{\frac{\sigma }{\sigma -1}-1}X_{ir}(s)^{\frac{ \sigma -1}{\sigma }-1}=\lambda p_{r}(s) \end{aligned}$$
(30)
$$\begin{aligned}&\mu A_{ir}^{1-\mu }\varPhi ^{\mu -1}\left[ 1+\alpha S(\Omega (\Lambda _{r}(i))) \right] \left( x\right) ^{\frac{\sigma }{\sigma -1}-1}X_{ir}(v)^{\frac{ \sigma -1}{\sigma }-1}=\lambda p_{r}(v) \end{aligned}$$
(31)
$$\begin{aligned}&w_{ir}=A_{ir}+\int \limits _{s\in N}p_{r}(s)X_{ir}(s)ds \end{aligned}$$
(32)

with \(x=\int \limits _{s\in N}X_{ir}(s)^{\frac{\sigma -1}{\sigma }}ds\). Considering the ratio of (30) and (31), we obtain:

$$\begin{aligned} \frac{X_{ir}(s)^{-\frac{1}{\sigma }}}{X_{ir}(v)^{-\frac{1}{\sigma }}}=\frac{ p_{r}(s)}{p_{r}(v)} \end{aligned}$$
(33)

Let us notice that from (33), we obtain that

$$\begin{aligned} X_{ir}(s)=\left( \frac{p_{r}(v)}{p_{r}(s)}\right) ^{\sigma }X_{ir}(v) \end{aligned}$$

which can be substituted in the definition of \(X_{ir}\equiv \left( \displaystyle \int \limits _{s\in N}X_{ir}(s)^{\frac{\sigma -1}{\sigma }}ds\right) ^{\frac{ \sigma }{\sigma -1}}\) to get

$$\begin{aligned} X_{ir}(v)=\dfrac{p_{r}(v)^{-\sigma }}{\left( \displaystyle \int \limits _{s\in N}\left( p_{r}(s)\right) ^{1-\sigma }ds\right) ^{\frac{\sigma }{\sigma -1}}}X_{ir}\ \ \ \end{aligned}$$
(34)

From (29), (31) and (34), we get that

$$\begin{aligned} A_{ir}=\frac{\left[ \left[ 1+\alpha S(\Omega (\Lambda _{r}(i)))\right] X_{ir}-\alpha S(\Omega (\Lambda _{r}(i)))\frac{\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j }{\Omega (\Lambda _{r}(i))}\right] \left( 1-\mu \right) p_{X_{r}}}{\mu \left[ 1+\alpha S(\Omega (\Lambda _{r}(i)))\right] } \end{aligned}$$
(35)

where \(p_{X_{r}}\equiv \left( \int \limits _{s\in N}(p_{r}(s))^{1-\sigma }ds\right) ^{\frac{1}{1-\sigma }}\) is the price index of manufactured varieties. Then, we substitute (35) in (32) and we use also (34) to obtain the consumer’s demand for the differentiated good in region r

$$\begin{aligned} X_{ir}=\frac{\mu }{p_{X_{r}}}\left( w_{ir}+p_{X_{r}}\frac{1-\mu }{\mu }\frac{ \alpha S(\Omega (\Lambda _{r}(i)))}{1+\alpha S(\Omega (\Lambda _{r}(i)))} \frac{\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j}{\Omega (\Lambda _{r}(i))}\right) \end{aligned}$$
(36)

Previous expression can be substituted in (35) to obtain the consumer’s demand for the agricultural good in region r, that is

$$\begin{aligned} A_{ir}=\left( 1-\mu \right) \left( w_{ir}-p_{X_{r}}\frac{\alpha S(\Omega (\Lambda _{r}(i)))}{1+\alpha S(\Omega (\Lambda _{r}(i)))}\frac{\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j}{\Omega (\Lambda _{r}(i))}\right) \end{aligned}$$
(37)

Then, substituting (36) and (37) in

$$\begin{aligned} U(A_{ir},\varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) )=A_{ir}^{1-\mu }\left( \varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) \right) ^{\mu } \end{aligned}$$

we obtain the indirect utility function:

$$\begin{aligned}&U(A_{ir},\varPhi \left( X_{ir},X_{-ir},\Lambda _{r}(i)\right) )=\frac{\left( 1-\mu \right) ^{1-\mu }\mu ^{\mu }\left( 1+\alpha S(\Omega (\Lambda _{r}(i)))\right) ^{\mu }}{\left( p_{X_{r}}\right) ^{\mu }} \\&\left( w_{ir}-p_{X_{r}}\frac{\alpha S(\Omega (\Lambda _{r}(i)))}{1+\alpha S(\Omega (\Lambda _{r}(i)))}\frac{\int _{\Lambda _{r}(i)}X_{jr}\mathrm{d}j)}{\Omega (\Lambda _{r}(i))}\right) \end{aligned}$$

Expression (34) can be used to derive an expression for the expenditure in manufacturing, \(E_{ir}\), that is

$$\begin{aligned} E_{ir}=\int \limits _{s\in N}p_{r}(s)X_{ir}(s)ds=p_{X_{r}}X_{ir} \end{aligned}$$

This expression can be substituted into (34) to get the individual demand in region r for variety v

$$\begin{aligned} X_{ir}(v)=\frac{p_{r}(v)^{-\sigma }}{p_{X_{r}}^{1-\sigma }}E_{ir} \end{aligned}$$

1.2 Individual demand of the differentiated good by workers in region \(\mathbf {r}\)

In this Appendix, we show how we derive the individual demand by the skilled and by the unskilled consumers in region r. As \(\int _{\Lambda _{r}(i)}\mathrm{d}j=\Omega (\Lambda _{r}(i))\), the demands for each skilled individual i in region r as given by (7) and (8) can be rewritten as

$$\begin{aligned} A_{H_{r}}= & {} \left( 1-\mu \right) \left[ w_{H_{r}}-p_{X_{r}}\frac{\alpha S(\Omega _{r}(H))}{1+\alpha S(\Omega _{r}(H))}\frac{\left( h_{H_{r}}X_{H_{r}}+l_{H_{r}}X_{L_{r}}\right) }{\Omega _{r}(H)}\right] , \\ X_{H_{r}}= & {} \tfrac{\mu }{p_{X_{r}}}\left[ w_{H_{r}}+p_{X_{r}}\frac{1-\mu }{ \mu }\frac{\alpha S(\Omega _{r}(H))}{1+\alpha S(\Omega _{r}(H))}\frac{ \left( h_{H_{r}}X_{H_{r}}+l_{H_{r}}X_{L_{r}}\right) }{\Omega _{r}(H)}\right] ; \nonumber \end{aligned}$$
(38)

while for each unskilled individual i in region r the demands are

$$\begin{aligned} A_{L_{r}}= & {} \left( 1-\mu \right) \left[ w_{L_{r}}-p_{X_{r}}\frac{\alpha S(\Omega _{r}(L))}{1+\alpha S(\Omega _{r}(L))}\frac{\left( h_{L_{r}}X_{H_{r}}+l_{L_{r}}X_{L_{r}}\right) }{\Omega _{r}(L)}\right] , \\ X_{L_{r}}= & {} \frac{\mu }{p_{X_{r}}}\left[ w_{L_{r}}+p_{X_{r}}\frac{1-\mu }{ \mu }\frac{\alpha S(\Omega _{r}(L))}{1+\alpha S(\Omega _{r}(L))}\frac{ \left( h_{L_{r}}X_{H_{r}}+l_{L_{r}}X_{L_{r}}\right) }{\Omega _{r}(L)}\right] . \nonumber \end{aligned}$$
(39)

Considering the second Eqs. in (38) and in (39 ), we obtain a system of two equations in the two unknowns \(X_{H_{r}}\) and \( X_{L_{r}}\) given by

$$\begin{aligned} \left\{ \begin{array}{l} X_{H_{r}}=w_{H_{r}}\frac{\mu }{p_{X_{r}}}+\frac{\alpha \left( 1-\mu \right) S(\Omega _{r}(H))}{1+\alpha S(\Omega _{r}(H))}\frac{\left( h_{H_{r}}X_{H_{r}}+l_{H_{r}}X_{L_{r}}\right) }{\Omega _{r}(H)} \\ X_{L_{r}}=\frac{\mu }{p_{X_{r}}}+\frac{\alpha \left( 1-\mu \right) S(\Omega _{r}(L))}{1+\alpha S(\Omega _{r}(L))}\frac{\left( h_{L_{r}}X_{H_{r}}+l_{L_{r}}X_{L_{r}}\right) }{\Omega _{r}(L)} \end{array} \right. \end{aligned}$$
(40)

where we used the fact that \(w_{L_{r}}=1\) and where \(\Omega _{r}(H)=h_{H_{r}}+l_{H_{r}}\), \(\Omega _{r}(L)=h_{L_{r}}+l_{L_{r}}\).

Appendix 2

In the case of the additive specification, we know that \(S(\Omega (\Lambda _{r}(i)))=\Omega (\Lambda _{r}(i))\). Hence, with \(\Omega _{r}(H)=h_{H_{r}}+l_{H_{r}}\) and \(\Omega _{r}(L)=h_{L_{r}}+l_{L_{r}}\), the system in (40) becomes

$$\begin{aligned} \left\{ \begin{array}{l} \frac{1+\alpha (h_{H_{r}}+l_{H_{r}})-\alpha \left( 1-\mu \right) h_{H_{r}}}{ 1+\alpha (h_{H_{r}}+l_{H_{r}})}X_{H_{r}}=w_{H_{r}}\frac{\mu }{p_{X_{r}}}+ \frac{\alpha \left( 1-\mu \right) }{1+\alpha (h_{H_{r}}+l_{H_{r}})} l_{H_{r}}X_{L_{r}} \\ \frac{1+\alpha (h_{L_{r}}+l_{L_{r}})-\alpha \left( 1-\mu \right) l_{L_{r}}}{ 1+\alpha (h_{L_{r}}+l_{L_{r}})}X_{L_{r}}=\frac{\mu }{p_{X_{r}}}+\frac{\alpha \left( 1-\mu \right) }{1+\alpha (h_{L_{r}}+l_{L_{r}})}h_{L_{r}}X_{H_{r}} \end{array} \right. \end{aligned}$$
(41)

This system is solved to obtain the individual demands of the differentiated good by the skilled workers and by the unskilled workers in region r, i.e. \(X_{H_{r}}\) and \(X_{L_{r}}\).

1.1 Complete networks

Let us now consider the complete network and derive the expression for skilled wages as a function of \(H_{r}\). Substituting \(X_{rr}(s)\) and \( X_{rv}(s)\) from (26) and (27) in (19) and making use of (14), the wage paid to skilled workers in region r must satisfy the following equation

$$\begin{aligned} w_{H_{r}}=\frac{\mu }{\sigma }\left[ \frac{\left( L+H_{r}w_{H_{r}}\right) }{ H_{r}+\left( H-H_{r}\right) \phi }\frac{1+\alpha \Omega _{r}}{1+\alpha \mu \Omega _{r}}+\phi \frac{\left( L+H_{v}w_{H_{v}}\right) }{\left( H-H_{r}\right) +H_{r}\phi }\frac{1+\alpha \Omega _{v}}{1+\alpha \mu \Omega _{v}}\right] \end{aligned}$$
(42)

and an analogous expression holds for \(w_{H_{v}}\). Hence, we get a system of two linear equations in \(w_{H_{r}}\) and \(w_{H_{v}}\) that can be solved to obtain the two regional wages for skilled workers as an explicit function of a given distribution of workers, \(H_{r}\) and \(H_{v}\), between the two regions, and we find that the wage paid in region r to skilled workers is

$$\begin{aligned} w_{H_{r}}=\mu L\tfrac{\sigma A\left( H_{v}+\phi H_{r}\right) +\phi \sigma B\left( H_{r}+\phi H_{v}\right) +H_{v}\mu BA\left( \phi -1\right) \left( \phi +1\right) }{\phi \sigma ^{2}\left( H_{r}^{2}+H_{v}^{2}\right) -\phi \mu \sigma \left( AH_{r}^{2}+BH_{v}^{2}\right) -H_{v}H_{r}\left[ \mu \sigma \left( A+B\right) -\sigma ^{2}\left( \phi ^{2}+1\right) +\mu ^{2}BA\left( \phi -1\right) \left( \phi +1\right) \right] } \end{aligned}$$
(43)

where \(A=\frac{\alpha \Omega _{r}+1}{\alpha \mu \Omega _{r}+1}\) and \(B=\frac{ \alpha \Omega _{v}+1}{\alpha \mu \Omega _{v}+1}\).Footnote 36 This expression shows that the wage depends on the value of \(\alpha \), and A and B represent a measure of the effects produced by the proximity of neighbours, respectively, in r (where the number of neighbours is given by \( \Omega _{r}=H_{r}+L\)) and in v (where the number of neighbours is given by \(\Omega _{v}=H_{v}+L\)).

The wage of skilled workers in (43) evaluated in the case in which they are evenly distributed in the two regions is given by the following expression

$$\begin{aligned} w_{H/2}=\frac{2\mu L}{H}\frac{\alpha \Omega _{s}+1}{\sigma -\mu +\mu \alpha \Omega _{s}\left( \sigma -1\right) } \end{aligned}$$

where \(\Omega _{s}\equiv \left( H/2+L\right) \) is the number of neighbours in each region at the symmetric equilibrium.Footnote 37 It can be readily shown that the wage (\(w_{H_{r}}\)) increases in the symmetric equilibrium with \(\alpha \). Finally, evaluating the wage of skilled workers in (43) when they are all in region r we get that

$$\begin{aligned} w_{H_{r}}=\frac{\mu L}{H}\frac{2+\alpha \left( H+2L\right) \left( \mu +1\right) +2\alpha ^{2}\mu L(H+L)}{\left( \mu \alpha L+1\right) \left[ \sigma -\mu +\alpha \mu \left( H+L\right) \left( \sigma -1\right) \right] } \end{aligned}$$

In what follows, we show that Proposition 3 in Sect. 4.2 holds.

Proof of Proposition 3

First, consider the derivative of \(V\left( H_{r},\phi ,\alpha \right) \) evaluated at the symmetric equilibrium. It can be shown using the expressions for skilled wages derived for the symmetric equilibrium in this Appendix that

$$\begin{aligned}&V_{H_{r}}\left( H/2,\phi ,\alpha \right) \end{aligned}$$
(44)
$$\begin{aligned}&\quad =\frac{4\left( f_{2}\phi ^{2}+f_{1}\phi +f_{0}\right) }{H(\phi +1)\left( \sigma -1\right) \left[ \alpha \left( H+2L\right) +2\right] \left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} }*\\&\quad \frac{1}{\left\{ 2\left[ \sigma -\mu +\phi \left( \sigma +\mu \right) \right] +\mu \alpha \left( H+2L\right) \left[ \sigma -1+\phi \left( \sigma +1\right) \right] \right\} } \nonumber \end{aligned}$$
(45)

where the coefficients \(f_{0}\), \(f_{1}\) and \(f_{2}\) are functions of \(\mu \), \(\sigma \), L, H and \(\alpha \). We know that all factors in the denominator are positive, given that \(\sigma >1>\mu \) and that the term \( \left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} \) is positive when \( W_{\alpha H_{r}}\) evaluated at \(H_{r}=H/2\) is positive.Footnote 38 Thus, the sign of \( V_{H_{r}}\left( H/2,\phi ,\alpha \right) \) depends on the sign of the expression \(F\equiv f_{2}\phi ^{2}+f_{1}\phi +f_{0}\) in the numerator, and the symmetric equilibrium is stable when \(F<0\) and unstable when \(F>0\). However, we know that when \(\alpha =0\), \(F=a_{0}=8(1-\phi )(2\phi \sigma \mu -\mu \phi -\phi \sigma +\phi \sigma ^{2}+\mu ^{2}\phi -\mu ^{2}+\sigma -\sigma ^{2}+2\sigma \mu -\mu )\). In this case, \(a_{0}\) is the relevant term in determining the sign of \(V_{H_{r}}\left( H/2,\phi ,\alpha \right) ,\) and, as in Forslid and Ottaviano (2003), the symmetry breaking point \(\phi _{b}^{FO}=\frac{\left( \sigma -1-\mu \right) \left( \sigma -\mu \right) }{ \left( \mu +\sigma \right) \left( \mu +\sigma -1\right) }<1\) is such that the symmetric equilibrium is stable only for \(\phi \in (0,\phi _{b}^{FO})\).Footnote 39 Note that for \(\phi =1\), \(F=a_{0}=0.\) As \(\alpha \) becomes positive and rises, F becomes negative, that is \(F=4\mu H\alpha ^{2}\sigma (\sigma -1)(1-\mu )(2L+H)\left\{ \alpha \left[ H\left( \sigma -1\right) -2L \right] -2\right\} <0\),Footnote 40 and the symmetric equilibrium remains stable for \(\phi =1.\) By continuity the result holds for high levels of integration (i.e. large \(\phi \)) and the symmetric equilibrium is always stable for low trade costs. Moreover, for prohibitively high trade costs, when \(\phi =0\), \(F=f_{0}\) that is negative for \(\alpha =0\) as in this case \( f_{0}=8(\mu +1-\sigma ))(\sigma -\mu )=F<0\); by continuity the result holds for positive and sufficiently small values of \(\alpha \), with \(f_{0}=8(\mu +1-\sigma ))(\sigma -\mu )+a_{1}\alpha +a_{2}\alpha ^{2}+a_{3}\alpha ^{3}=F<0 \) where the coefficients \(a_{1}\), \(a_{2}\) and \(a_{3}\) are functions of \(\mu \), \(\sigma \), L and H. This proves the results in Proposition 3 on the symmetric equilibrium.\(\square \)

We now focus our attention on equilibria in which all skilled workers move to one region. Note that, provided these full agglomeration equilibria exist, they are stable, and in what follows we show that the final part of Proposition 3 holds.

The value of \(V\left( H_{r},\phi ,\alpha \right) \) when all skilled workers are located in region r is given by

$$\begin{aligned} V(H,\phi ,\alpha )=\ln \left\{ \frac{g_{0}\left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} \phi ^{1-\frac{\mu }{\sigma -1}}}{d_{2}\phi ^{2}+d_{1}\phi +d_{0}}\right\} \end{aligned}$$
(46)

where

$$\begin{aligned} g_{0}= & {} \sigma \left( 1+\alpha \mu L\right) \left[ \tfrac{\alpha (H+L)+1}{ \alpha L+1}\right] ^{\mu }>1 \\ d_{2}= & {} [1+\alpha \left( H+L\right) ][\mu \alpha L\left( \sigma +1\right) +\sigma +\mu ]>0; \\ d_{1}= & {} -H\alpha \sigma \left[ \mu \alpha \left( H+L\right) \left( \sigma -1\right) +\sigma -\mu \right] <0; \\ d_{0}= & {} (1+\alpha L)\left[ \mu \alpha \left( H+L\right) \left( \sigma -1\right) +\sigma -\mu \right] >0 \end{aligned}$$

When \(\alpha =0\), expression (46) becomes

$$\begin{aligned} V(H,\phi )=\ln \left( \frac{2\sigma \phi ^{1+\frac{\mu }{1-\sigma }}}{\sigma (1+\phi ^{2})-\mu (1-\phi ^{2})}\right) \end{aligned}$$
(47)

Equation (47) generates the sustain point \(\phi _{s}\), i.e. the value of \(\phi \) such that full agglomeration is an equilibrium for \(\phi >\phi _{s},\) as in Forslid and Ottaviano (2003).Footnote 41 When \( \alpha \) is positive, we know that \(g_{0}>1\) and that expression \(\phi ^{1+ \frac{\mu }{1-\sigma }}\) is increasing in \(\phi \in [0,1]\) from 0 (if \(\phi =0\)) to 1 (if \(\phi =1\)).Footnote 42 Given the sign of the parameters \(d_{2}\), \(d_{1}\) and \(d_{0}\), expression \(d_{2}\phi ^{2}+d_{1}\phi +d_{0}\) in the denominator of (46) is an upward opening parabola in \(\phi ~\)with positive value \(d_{0}\) when \(\phi =0\).Footnote 43 Moreover, we know that, when workers are all agglomerated in r, expression \(\left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} \) in the numerator must be positive to have \(W_{\alpha H_{r}}>0\).Footnote 44 Finally, the parabola in the denominator (\(d_{2}\phi ^{2}+d_{1}\phi +d_{0}\)) must be positive in order to have \(W_{\alpha H_{v}}>0\). Figure 3 presents the two possible scenarios for this parabola by means of the two continuous curves when \(\alpha \) is positive and sufficiently small. In Fig. 3, the numerator \(N=g_{0}\left\{ 2+\alpha \left[ 2L-H(\sigma -1) \right] \right\} \phi ^{1-\frac{\mu }{\sigma -1}}\) is represented by the dotted line, which characterizes an increasing function in \(\phi \) taking the value 0 when \(\phi =0\) and \(g_{0}\left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} >0\) when \(\phi =1\).Footnote 45 The lower parabola can be excluded with \(\alpha >0\) as it implies that with \( \phi =1\) the full agglomeration is an equilibrium, which contradicts direct computation.Footnote 46 The only relevant case is then the other parabola \(D=d_{2}\phi ^{2}+d_{1}\phi +d_{0}\) (that intersects the dotted curve in two critical points \(\phi _{s}\) and \(\phi _{u}\)).Footnote 47 In this case, full agglomeration in region r is an equilibrium only for intermediate values of \(\phi \) (when the higher parabola lies below the dotted curve) for \(\phi \in (\phi _{s},\phi _{u})\) as \(V(H,\phi ,\alpha )\) is positive. This proves the results in Proposition 3 on full agglomeration equilibria. \(\square \)

Fig. 3
figure 3

Full agglomeration equilibrium

A general overview: the bifurcation diagram.

To explore further the relationship between the strength of relative concerns (\(\alpha \)), the openness to trade (\(\phi \)) and the existence and stability of the equilibria we need to focus on numerical simulations of the model. First, let us perform the numerical analysis on a model in which \( L=20 \), \(H=10\), \(\sigma =2\) and \(\mu =0.4\) (Fig. 4a) or \(\mu =0.11\) (Fig. 4b).Footnote 48 We know that the sign of \( V_{H_{r}}\left( H/2,\phi ,\alpha \right) \) in (44) depends on the sign of the expression \(F\equiv f_{2}\phi ^{2}+f_{1}\phi +f_{0}\), and the symmetric equilibrium is stable when \(F<0\) and unstable when \(F>0\). The solid curves represent \(F=a_{0}\) as a function of \(\phi \) when \(\alpha =0\), as in Forslid and Ottaviano (2003). They show that in this case the symmetric equilibrium is stable only if \(F<0\), that is, only if \(\phi <\phi _{b}^{FO}\). Assume now the existence of relative concerns, \(\alpha >0\). Figure 4a, b represents the function F for \(\alpha =0.02\) , 0.04, 0.06, 0.08. They show that with relative concerns, as \(\phi \) rises the symmetric equilibrium becomes stable as soon as \(\phi >\phi _{d}\), where \(\phi _{d}\) is the “dispersion point” (so-called as the symmetric equilibrium is stable for \(\phi >\phi _{d}\)). Hence, the symmetric equilibrium is stable both for small values of the freeness of trade \(\phi ,\) such that \(\phi <\phi _{b}\), and for high levels of integration, such that \( \phi \) is above the new critical level \(\phi _{d}\).

Then, we can use expression (46) to find the other two critical points \(\phi _{s}\) and \(\phi _{u}\) for the equilibrium with full agglomeration. Full agglomeration in region r is an equilibrium only for \( \phi \in (\phi _{s},\phi _{u})\). As \(\phi \) rises from 0 and reaches the sustain point, \(\phi _{s}\), full agglomeration becomes a stable equilibrium for \(\phi >\phi _{s}\), as in the case \(\alpha =0\). However, this equilibrium again disappears with \(\alpha >0\) after trade costs have decreased beyond a new critical level, that we name “unsustain point”, \(\phi _{u}\).

Fig. 4
figure 4

Stability of the symmetric equilibrium for the complete network: the plot of F

Table 1 shows the values of \(\phi _{s}\), \(\phi _{b}\), \(\phi _{u}\) and \( \phi _{d}\) obtained numerically for the model with \(L=4\), \(H=10\), \(\sigma =2\) and \(\mu =0.11\).Footnote 49 The analysis in Table 1 is summarized by the bifurcation diagrams reported in Fig. 1 in the text of the paper.

Table 1 Critical points for different values of \(\alpha \)

1.2 Segregated networks

In the case of the segregated network, the system in (41) becomes

$$\begin{aligned} \left\{ \begin{array}{l} \frac{1+\alpha H_{r}-\alpha \left( 1-\mu \right) H_{r}}{1+\alpha H_{r}} X_{H_{r}}=w_{H_{r}}\frac{\mu }{p_{X_{r}}} \\ \frac{1+\alpha L-\alpha \left( 1-\mu \right) L}{1+\alpha L}X_{L_{r}}=\frac{ \mu }{p_{X_{r}}} \end{array} \right. \end{aligned}$$
(48)

which can be solved for \(X_{H_{r}}\) and \(X_{L_{r}}\) to find respectively that

$$\begin{aligned} X_{H_{r}}= & {} \mu \frac{w_{H_{r}}}{p_{X_{r}}}\frac{1+\alpha H_{r}}{1+\alpha \mu H_{r}}\,\,\text { and} \\ X_{L_{r}}= & {} \mu \frac{1}{p_{X_{r}}}\frac{1+\alpha L}{1+\alpha \mu L}. \nonumber \end{aligned}$$
(49)

Making use of these solutions, we can rewrite the total demand in region r for variety s produced in region k in (13) as follows

$$\begin{aligned} X_{kr}(s)=\mu \frac{p_{kr}(s)^{-\sigma }}{p_{X_{r}}^{1-\sigma }}\left( w_{H_{r}}\frac{1+\alpha H_{r}}{1+\alpha \mu H_{r}}H_{r}+\frac{1+\alpha L}{ 1+\alpha \mu L}L\right) \end{aligned}$$
(50)

where the price indices and the prices are respectively given by (17), (18) and (16). Making use of (50), the wage in (19) can be rewritten as follows

$$\begin{aligned} w_{H_{r}}=\frac{\mu }{\sigma }\left[ \frac{\left( A_{r}w_{H_{r}}+A_{L}\right) }{H_{r}+\left( H-H_{r}\right) \phi }+\frac{\phi \left( A_{v}w_{H_{v}}+A_{L}\right) }{\left( H-H_{r}\right) +H_{r}\phi } \right] \end{aligned}$$

where \(A_{r}\equiv \frac{1+\alpha H_{r}}{1+\alpha \mu H_{r}}H_{r}\), \( A_{v}\equiv \frac{1+\alpha \left( H-H_{r}\right) }{1+\alpha \mu \left( H-H_{r}\right) }\left( H-H_{r}\right) \) and \(A_{L}\equiv \frac{1+\alpha L}{ 1+\alpha \mu L}L\). Previous expression can be considered together with the analogous expression obtained for \(w_{H_{v}}\) to get a system of two equations in two unknowns \(w_{H_{r}}\) and \(w_{H_{v}}\), that can be solved to find the two skilled regional wages given by

$$\begin{aligned} w_{H_{r}}=\frac{\mu A_{L}\left[ 2\sigma \phi H_{r}+\sigma H_{v}\left( 1+\phi ^{2}\right) -A_{v}\mu \left( 1-\phi \right) \left( \phi +1\right) \right] }{ D_{s}} \end{aligned}$$
(51)

and

$$\begin{aligned} w_{H_{v}}=\frac{\mu A_{L}\left[ 2\sigma \phi H_{v}+\sigma H_{r}\left( \phi ^{2}+1\right) -A_{r}\mu \left( 1-\phi \right) \left( \phi +1\right) \right] }{D_{s}} \end{aligned}$$
(52)

with the denominator of both wages given by \(D_{s}\equiv \sigma ^{2}\left( H_{v}+\phi H_{r}\right) \left( H_{r}+\phi H_{v}\right) -A_{r}\sigma \mu \left( H_{v}+\phi H_{r}\right) -\sigma \mu A_{v}\left( H_{r}+\phi H_{v}\right) +\mu ^{2}\left( 1-\phi \right) \left( 1+\phi \right) A_{r}A_{v}\).

Then, \(W_{\alpha ir}\) in (9) can be written for skilled workers in region r as

$$\begin{aligned} W_{\alpha H_{r}}=\frac{w_{H_{r}}}{1+\alpha \mu H_{r}}, \end{aligned}$$

while for unskilled it is given by

$$\begin{aligned} W_{\alpha L_{r}}=\frac{1}{1+L\alpha \mu } \end{aligned}$$

Hence, we can rewrite (23) as follows

$$\begin{aligned} V\left( H_{r},\phi ,\alpha \right) =\ln \left\{ \left[ \frac{\left( H-H_{r}\right) +H_{r}\phi }{H_{r}+\left( H-H_{r}\right) \phi }\right] ^{ \frac{\mu }{1-\sigma }}\left[ \frac{1+\alpha H_{r}}{1+\alpha \left( H-H_{r}\right) }\right] ^{\mu }\frac{\frac{w_{H_{r}}}{1+\alpha \mu H_{r}}}{ \frac{w_{H_{v}}}{1+\alpha \mu \left( H-H_{r}\right) }}\right\} \end{aligned}$$
(53)

where the two regional wages for skilled workers \(w_{H_{r}}\) and \(w_{H_{v}}\) can be substituted respectively from (51) and (52).

We first focus on the equilibrium with full agglomeration, and we show that what stated in Proposition 4 in Sect. 5 for segregated networks holds.

Proof of Proposition 4

Expression (53) evaluated when all skilled workers are located in region r (i.e. \(H_{r}=H\) and \( H_{v}=0\)) is

$$\begin{aligned} V^{s}(H,\phi ,\alpha )=\ln \left\{ \frac{2\sigma \left( 1+\alpha H\right) ^{\mu }\phi ^{1+\frac{\mu }{1-\sigma }}}{\mu H\left[ \sigma -1+\phi ^{2}\left( \sigma +1\right) \right] \alpha +\left[ \sigma -\mu +\phi ^{2}\left( \sigma +\mu \right) \right] }\right\} \end{aligned}$$

The value of \(V^{s}(H,\phi ,\alpha )\) in this case does not depend on L. On the one hand, when \(\phi \rightarrow 0\) (that is, in autarky), \( V^{s}\left( H,0,\alpha \right) \rightarrow -\infty \) and full agglomeration is not an equilibrium. On the other hand, with complete integration, \(\phi =1 \), we find that \(V^{s}\left( H,\phi ,\alpha \right) =\ln \frac{\left( 1+\alpha H\right) ^{\mu }}{\left( 1+H\alpha \mu \right) }\), which is negative, provided \(\alpha >0,\) because \(\frac{\partial \left( \frac{\left( 1+\alpha H\right) ^{\mu }}{\left( 1+H\alpha \mu \right) }\right) }{\partial \alpha }=-\mu \alpha H^{2}\frac{\left( 1-\mu \right) }{H\alpha +1}\frac{ \left( H\alpha +1\right) ^{\mu }}{\left( H\alpha \mu +1\right) ^{2}}<0\) and full agglomeration is not an equilibrium either with complete integration. Moreover, when \(\alpha \) is positive and sufficiently small full agglomeration is an equilibrium only for intermediate values of openness to trade as the numerator, which is a concave function in \(\phi \) that increases from 0 when \(\phi =0\) to \(2\sigma \left( 1+\alpha H\right) ^{\mu }\) when \(\phi =1\), and the denominator, which is a convex parabola in \(\phi \) that increases from its minimum value \(\mu H\left( \sigma -1\right) \alpha +\sigma -\mu >0\) when \(\phi =0\) to \(2\sigma \left( 1+H\alpha \mu \right) \) when \(\phi =1\), intersect twice in \(\phi _{s}\) and \(\phi _{u}\), both in the range (0,1).Footnote 50 \(\square \)

We now consider the symmetric equilibrium, and we show that we can derive the results of the numerical analysis mentioned in the final part of Sect. 5 for the symmetric equilibrium. The derivative of \(V\left( H_{r},\phi ,\alpha \right) \) in (53) with respect to \(H_{r}\) evaluated at the symmetric equilibrium is given by the expression

$$\begin{aligned} V_{H_{r}}^{s}\left( H/2,\phi ,\alpha \right) =\frac{4\left( k_{2}\phi ^{2}+k_{1}\phi +k_{0}\right) }{H(1+\phi )(\sigma -1)(2+\alpha H)\left( h_{1}\phi +h_{0}\right) } \end{aligned}$$

where the coefficients \(k_{2}\), \(k_{1}\), \(k_{0}\), \(h_{1}\) and \(h_{0}\) are functions of \(\mu \), \(\sigma \), H and \(\alpha \) and they do not depend on L.Footnote 51 As the denominator of \(V_{H_{r}}^{s}\left( H/2,\phi ,\alpha \right) \) is positive, the sign of \(V_{H_{r}}^{s}\left( H/2,\phi ,\alpha \right) \) depends on that of the parabola \(G\equiv k_{2}\phi ^{2}+k_{1}\phi +k_{0}\), with the symmetric equilibrium stable (vs. unstable) only when G is negative (vs. positive). We focus our analysis on the simulated model with \( H=10\) and \(\sigma =2.\) Figure 5a (on the left) represents the value of G as a function of \(\phi \) in the case of \(\alpha =0\) (solid curve), \(\alpha =0.02\) (dash dot), \(\alpha =0.04\) (dash), \(\alpha =0.06\) (dot) and \(\alpha =0.08\) (long dash) when \(\mu =0.4\).

Fig. 5
figure 5

Stability of the symmetric equilibrium for the segregated network: the plot of G

Figure 5b (on the right) represents G for the same values except that now \(\mu =0.11\), where \(\mu \) is proportional to the share of income, net of the conspicuous effect, devoted to acquire the differentiated good.Footnote 52 The numerical analysis shows that the introduction of weak relative concerns tends to stabilize the symmetric equilibrium for high levels of integration \( \phi .\) Stability always holds for \(\phi =1\) as then \(G=4\alpha ^{2}H^{2}\mu \sigma (\sigma -1)(\mu -1)<0\). The size of the interval of \(\phi \) on which stability holds is increasing in \(\alpha \) and decreasing with \(\mu .\) Figure 5a also shows that for intermediate levels of integration \(\phi \) the symmetric equilibrium is destabilized when \(\mu \) is large. Indeed, in this case the agglomerative effect they produce is stronger than the dispersion effect. On the other hand, Fig. 5b shows that when \(\mu \) is relatively low the symmetric equilibrium can be stabilized for low and intermediate values of economic integration \(\phi \) provided \(\alpha \) is sufficiently large.

1.3 The mixed case

Let us consider the case of the additive specification. Then when region r has a complete (integrated) network, we know that the aggregate demand in r of variety s produced in r, \(X_{rr}(s)\), is given by (26), while the aggregate demand in r of variety s produced in v can be obtained from (27) and it is given by

$$\begin{aligned} X_{vr}(s)=\mu \frac{\left( \sigma -1\right) \tau ^{-\sigma }}{\sigma \beta \left( n_{r}+n_{v}\phi \right) }\frac{\left( L+H_{r}w_{H_{r}}\right) \left( L\alpha +\alpha H_{r}+1\right) }{1+\alpha \mu \left( L+H_{r}\right) } \end{aligned}$$

On the other hand, given that region v has two segregated networks, we know that the aggregate demands in v of variety s produced respectively in v and in r can be obtained from (50). Hence, the wage in (19) in region r in the mixed case can be rewritten as follows

$$\begin{aligned} w_{H_{r}}=\frac{\mu }{\sigma }\left[ \frac{\left( L+H_{r}w_{H_{r}}\right) }{ \left( H_{r}+H_{v}\phi \right) }A+\phi \frac{\left( w_{H_{v}}A_{v}+A_{L}\right) }{\left( H_{v}+H_{r}\phi \right) }\right] \end{aligned}$$

while that obtained in region v is given by

$$\begin{aligned} w_{H_{v}}=\frac{\mu }{\sigma }\left[ \frac{\left( w_{H_{v}}A_{v}+A_{L}\right) }{\left( H_{v}+H_{r}\phi \right) }+\frac{\phi \left( L+H_{r}w_{H_{r}}\right) }{\left( H_{r}+H_{v}\phi \right) }A\right] \end{aligned}$$

We use these last two equations to find the wages of skilled workers in the two regions, which are respectively

$$\begin{aligned}&w_{H_{r}}=\mu \\&\dfrac{\sigma \phi ^{2}A_{L}H_{v}-AL\mu A_{v}+AL\sigma H_{v}+\sigma \phi A_{L}H_{r}+AL\mu \phi ^{2}A_{v}+AL\sigma \phi H_{r}}{ \sigma ^{2}H_{r}H_{v}{+}\sigma ^{2}\phi H_{r}^{2}{+}\sigma ^{2}\phi H_{v}^{2}{+}A\mu ^{2}A_{v}H_{r}{+}\sigma ^{2}\phi ^{2}H_{r}H_{v}{-}\sigma \mu A_{v}H_{r}{-}\sigma \mu \phi A_{v}H_{v}{-}A\sigma \mu \phi H_{r}^{2}{-}A\mu ^{2}\phi ^{2}A_{v}H_{r}{-}A\sigma \mu H_{r}H_{v}} \end{aligned}$$

and

$$\begin{aligned}&w_{H_{v}}=\mu \\&\frac{\sigma A_{L}H_{r}-A\mu A_{L}H_{r}+\sigma \phi A_{L}H_{v}+AL\sigma \phi ^{2}H_{r}+A\mu \phi ^{2}A_{L}H_{r}+AL\sigma \phi H_{v}}{\sigma ^{2}H_{r}H_{v}{+}\sigma ^{2}\phi H_{r}^{2}{+}\sigma ^{2}\phi H_{v}^{2}{+}A\mu ^{2}A_{v}H_{r}{+}\sigma ^{2}\phi ^{2}H_{r}H_{v}{-}\sigma \mu A_{v}H_{r}{-}\sigma \mu \phi A_{v}H_{v}{-}A\sigma \mu \phi H_{r}^{2}{-}A\mu ^{2}\phi ^{2}A_{v}H_{r}{-}A\sigma \mu H_{r}H_{v}} \end{aligned}$$

where the two denominators in the two equations are equal.

Then, we find that with \(W_{\alpha Hr}\) for skilled workers in the integrated region r given by

$$\begin{aligned} W_{\alpha Hr}\equiv w_{H_{r}}-\alpha \mu \frac{L+H_{r}w_{H_{r}}}{L\alpha \mu +\alpha \mu H_{r}+1} \end{aligned}$$

and \(W_{\alpha Hv}\) for skilled workers in the segregated region v

$$\begin{aligned} W_{\alpha Hv}=\frac{w_{H_{v}}}{\alpha \mu H_{v}+1}, \end{aligned}$$

the log of the indirect utility levels \(V\left( H_{r},\phi ,\alpha \right) \) in the mixed case is given by

$$\begin{aligned} V^{mix}\left( H_{r},\phi ,\alpha \right) =\ln \left[ \left( \tfrac{p_{X_{v}} }{p_{X_{r}}}\right) ^{\mu }\left( \tfrac{1+\alpha \left( L+H_{r}\right) }{ 1+\alpha H_{v}}\right) ^{\mu }\tfrac{w_{H_{r}}-\alpha \mu \frac{ L+H_{r}w_{H_{r}}}{L\alpha \mu +\alpha \mu H_{r}+1}}{\frac{w_{H_{v}}}{\alpha \mu H_{v}+1}}\right] \end{aligned}$$
(54)

Evaluating \(V^{mix}\left( H_{r},\phi ,\alpha \right) \) in (54) when all skilled workers are located in the integrated region r we obtain the following expressionFootnote 53

$$\begin{aligned} V^{mix}\left( H,\phi ,\alpha \right) =\ln \tfrac{\left( 1+\alpha \mu L\right) \left\{ 2+\alpha \left[ 2L-H(\sigma -1)\right] \right\} \left[ \alpha (H+L)+1\right] ^{\mu }\sigma \phi ^{1-\frac{\mu }{\sigma -1}}}{ \left\{ [L\alpha \mu \left( \sigma +1\right) +\mu +\sigma ][1+\alpha (H+L)]\right\} \phi ^{2}+(L\alpha +1)[\alpha \mu \left( \sigma -1\right) \left( H+L\right) +\sigma -\mu ]} \end{aligned}$$

From the inspection of \(V^{mix}\left( H,\phi ,\alpha \right) \), we know that agglomeration in r can be an equilibrium only for high or intermediate \( \phi \). When \(\phi =1\), the argument of the logarithm in \(V^{mix}\left( H,1,\alpha \right) \) is equal to 1 when \(\alpha =0\). With a positive value of \(\alpha \), we can show that the argument of the logarithm in \( V^{mix}\left( H,1,\alpha \right) \) is smaller than 1 and, thus, agglomeration is an equilibrium when \(\phi =1\), only for relatively large value of \(\mu \) provided that the number of unskilled workers is relatively large with respect to that of skilled workers, that is \(H<2L/(\sigma -1)\).Footnote 54 More generally, we know that in this last case full agglomeration in r is an equilibrium also for all values of \(\phi \in \left( \phi _{s},1\right) \).

Instead, evaluating \(V^{mix}\left( H_{r},\phi ,\alpha \right) \) in (54) when all skilled workers are located in the segregated region v, we find that

$$\begin{aligned}&V^\mathrm{{mix}}\left( 0,\phi ,\alpha \right) \\&\quad =\ln \tfrac{\left\{ (L\alpha +1)[H\alpha \mu \left( \sigma +1\right) +\sigma +\mu ]\phi ^{2}-H\sigma \alpha [\left( \sigma -1\right) \alpha \mu H+\sigma -\mu ]\phi +(L\alpha +1)\left[ H\alpha \mu \left( \sigma -1\right) +\sigma -\mu \right] \right\} }{2\sigma (L\alpha +1)^{1-\mu }\left( \alpha H+1\right) ^{\mu }\phi ^{1-\frac{\mu }{\sigma -1}}} \end{aligned}$$

From the inspection of \(V^{mix}\left( 0,\phi ,\alpha \right) \), we know that agglomeration in v can be an equilibrium only for high or intermediate \( \phi \). When \(\phi =1\), the argument of the logarithm in \(V^{mix}\left( 0,1,\alpha \right) \) is equal to 1 when \(\alpha =0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ghiglino, C., Nocco, A. When Veblen meets Krugman: social network and city dynamics. Econ Theory 63, 431–470 (2017). https://doi.org/10.1007/s00199-015-0940-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-015-0940-5

Keywords

JEL Classification

Navigation