Skip to main content
Log in

The socially optimal and equilibrium locations of two stores or libraries with consumer search

  • Original Paper
  • Published:
The Annals of Regional Science Aims and scope Submit manuscript

Abstract

This paper examines the socially optimal (and also equilibrium) locations of two stores or libraries on a linear market of unit length. If each consumer has probability \(w\) of finding a desired product at each store, then the socially optimal locations are never completely centralized for full information, but are completely centralized for \(w \le 0.5\) when costly visit search is necessary. The Nash equilibrium locations of two stores, and various alternative models for the socially optimal locations of two stores, are also examined.

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.

Similar content being viewed by others

Notes

  1. A few of these consider more than two firms, but they are applicable to two firms.

  2. This simple search sequence contrasts with the much more complicated search problems in Narula et al. (1987) and Harwitz et al. (1998, 2000), in which a consumer determines the optimal sequence in which to visit several stores.

  3. Simply set the partial derivatives of (3) with respect to \(x_1\) and \( x_2\) equal to zero and then solve for \( x_1\) and \( x_2\). Calculating the second derivative matrix of (3) shows that AETD is a strictly convex function of \( x_1\) and \( x_2\) over the feasible solution set (see Definition 1), so we have a local and global minimum of AETD.

  4. If there were three stores (or libraries), the consumer would not necessarily visit the closest store first (see footnote 2), even if (as in this paper) the prices of the stores are assumed to be fixed and equal (see paragraph 3 of Sect. 10).

  5. If \(0 \le x \le x_1\), and store 1 does not have the product, then the consumer continues directly on to store 2, and then returns home. The right- hand side of (4) could also be written as \(2w( x_1 - x) + 2(1 - w)(x_2 - x)\). If \( x_1\le x \le (x_1 + x_2)/2\), and store 1 does not have the product, then the consumer passes by his home on the way to store 2, and then returns home for good.

  6. Note also that the level curve that is tangent to the third boundary locus is strictly concave (when expressed in terms of \(x_2\) as a function of \(x_1\)) in a neighborhood of (0.5, 0.5). The level curve that forms the boundary of the less-than-or-equal set might curl around and be a closed curve that is interior to the loci \(x_2 = x_1, x_2 = 0, \hbox {and}\; x_1 = 1\) (this is not the feasible solution set) except at (0.5, 0.5). It is not necessary to determine its exact shape, since all that matters is that its only intersection with the feasible set is at (0.5, 0.5).

  7. Proposition 4 is the same as Proposition 1 of Braid (2008), and Proposition 1 in Sect. 3 of this paper can be derived from it if we set \(b = {w^{2}}\) and \(c = w(1-w)\). Also, a variation of Proposition 4 with quadratic transportation costs is stated in a preliminary section of Braid (2011). However, the socially optimal location results for costly visit search (see Propositions 2 and 5), which are the main results of this paper, are completely new to the literature.

  8. Variable prices might very well lead to different results. If \(w = 1\), there is a central location tendency, but a price equilibrium does not exist if the firms are inside the quartiles (see d’Aspremont et al. 1979). This can be avoided with mixed-price strategies (see Osborne and Pitchik 1987), or with travel costs that are proportional to distance squared (see d’Aspremont et al. 1979), in which case the sellers move apart to the locations (0, 1). Also, if \(w < 1\), the equilibrium prices might be infinite for any set locations, since each seller has a monopoly over consumers for whom it has the desired good and the other seller does not. To avoid this, it might be possible to introduce a reservation price of each consumer (as in Economides 1984).

  9. If the length of the jurisdiction is \(L\), AETD would also be multiplied by \(L^{2}\).

  10. I was assuming that each consumer has the same completely inelastic demand. This assumption is often, though far from always, used in spatial models. Mulligan (2012) generalizes this in several ways. In this section, I generalize it in the simplest possible way, by assuming that each consumer has the same (finite) reservation price, as in Beckmann (1972), Economides (1984) and Salop (1979).

  11. As one of many possible alternatives, it might be possible to put one store at location 3/10, and consumers in the interval [1/10, 1/2] would visit it (since the equation \(wv = 2td\) implies that \(d =\) (1/2)(1/6)(12/5) \(=\) 1/5), and the other store at location 7/10, and consumers in the interval [1/2, 9/10] would visit it. However, even though a (slightly) greater number of consumers would visit a store, fewer would find what they want (each consumer who travels is visiting one store rather than two), and the overall consumer surplus would be less (and the overall producer surplus would also be lower, if the model is interpreted as stores, and if the price, assumed to be fixed, exceeds marginal costs, assumed to be constant).

  12. Since some consumers do not make a trip with any of these socially optimal locations, it is natural to imagine that it might be socially optimal to have more than two stores (or libraries). In fact, in terms of libraries, the significant cost of getting downtown for many consumers is a reason for a large city to have a central (main) library and a significant number of branch libraries (paragraphs 2 and 3 of Sect. 10 are somewhat related to this).

  13. See also Section 5, and see Proposition 1 of Braid (2008), but he considers only the full-information case (see fn. 7).

  14. Visit search is also assumed in much of Braid (2014), which has variable prices but fixed store locations.

References

  • Anderson SP, de Palma A (1988) Spatial price discrimination with heterogeneous products. Rev Econ Stud 55:573–592

  • Anderson SP, Neven D (1991) Cournot competition yields spatial agglomeration. Int Econ Rev 32:793–808

    Article  Google Scholar 

  • Beckmann MJ (1972) Spatial Cournot oligopoly. Pap Reg Sci Assoc 28:37–47

  • Berry BJL, Garrison WL (1958a) The functional bases of the central place hierarchy. Econ Geogr 34:145–154

    Article  Google Scholar 

  • Berry BJL, Garrison WL (1958b) A note on central place theory and the range of a good. Econ Geogr 34:304–311

    Article  Google Scholar 

  • Bigman D, ReVelle C (1978) The theory of welfare considerations in public facility location problems. Geogr Anal 10:229–240

    Article  Google Scholar 

  • Biscaia R, Mota I (2013) Models of spatial competition: a critical review. Pap Reg Sci 92:851–871

    Google Scholar 

  • Braid RM (1996) The optimal locations of branch facilities and main facilities with consumer search. J Reg Sci 36:217–234

    Article  Google Scholar 

  • Braid RM (1999) The socially optimal locations of three stores with stockouts or limited product selections. Econ Lett 64:363–368

    Article  Google Scholar 

  • Braid RM (2008) Spatial price discrimination and the locations of firms with different product selections or product varieties. Econ Lett 98:342–347

    Article  Google Scholar 

  • Braid RM (2011) Bertrand-Nash mill pricing and the locations of two firms with partially overlapping product selections. Pap Reg Sci 90:197–211

    Article  Google Scholar 

  • Braid RM (2014) Search costs decrease prices in a model of spatial competition. Pap Reg Sci 93:125–139

    Google Scholar 

  • d’Aspremont C, Gabszewicz JJ, Thisse J-F (1979) On Hotelling’s ‘Stability in competition’. Econometrica 47:1145–1150

    Article  Google Scholar 

  • de Fraja G, Norman G (1993) Product differentiation, pricing policy and equilibrium. J Reg Sci 33:343–363

  • de Palma A, Ginsburgh V, Papageorgiou Y, Thisse J-F (1985) The principle of minimum differentiation holds under sufficient heterogeneity. Econometrica 53:767–781

  • Eaton BC, Lipsey RG (1982) An economic theory of central places. Econ J 92:56–72

    Article  Google Scholar 

  • Economides N (1984) The principle of minimum differentiation revisited. Eur Econ Rev 24:345–368

    Article  Google Scholar 

  • Economides N (1986) Minimal and maximal product differentiation in Hotelling’s duopoly. Econ Lett 21:67–71

    Article  Google Scholar 

  • Fischer JH, Harrington JE (1996) Product variety and firm agglomeration. Rand J Econ 27:281–309

    Article  Google Scholar 

  • Fujita M, Thisse J-F (2002) Economics of agglomeration: cities, industrial location, and regional growth. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Gupta B (1992) Sequential entry and deterrence with competitive spatial price discrimination. Econ Lett 38:487–490

    Article  Google Scholar 

  • Hamilton JH, Thisse J-F, Weskamp A (1989) Spatial discrimination: Bertrand versus Cournot in a model of location choice. Reg Sci Urban Econ 19:87–102

    Article  Google Scholar 

  • Harwitz M, Lentnek B, Rogerson P, Smith TE (1998) Optimal search on spatial paths with recall: theoretical foundations. Pap Reg Sci 77:301–327

    Article  Google Scholar 

  • Harwitz M, Lentnek B, Rogerson P, Smith TE (2000) Optimal search on spatial paths with recall, part II: computational procedures and examples. Pap Reg Sci 79:293–305

    Article  Google Scholar 

  • Hsu SK (1983) Pricing in an urban spatial monopoly: a general analysis. J Reg Sci 23:165–175

    Article  Google Scholar 

  • Massam BH (1974) A note on expected distances. Geogr Anal 4:377–381

    Google Scholar 

  • Meza S, Tombak M (2009) Endogenous location leadership. Int J Ind Organ 27:687–707

    Article  Google Scholar 

  • Mulligan G (2012) Stochastic shopping and consumer heterogeneity. Stud Reg Sci 42:93–107

    Article  Google Scholar 

  • Mulligan G, Partridge MD, Carruthers JI (2012) Central place theory and its reemergence in regional science. Ann Reg Sci 48:405–431

    Article  Google Scholar 

  • Narula SC, Lentnek B, Harwitz M (1987) A contextual analysis of the journey-to-shop with price uncertainty. J Reg Sci 27:403–418

    Article  Google Scholar 

  • Osborne MJ, Pitchik C (1987) Equilibrium in Hotelling’s model of spatial competition. Econometrica 55:911–922

    Article  Google Scholar 

  • Quinzii M, Thisse J-F (1990) On the optimality of central places. Econometrica 58:1101–1119

    Article  Google Scholar 

  • ReVelle C (1991) Siting ambulances and fire companies. J Am Plan Assoc 57:471–484

    Article  Google Scholar 

  • Salop SC (1979) Monopolistic competition with outside goods. Bell J Econ 10:141–156

    Article  Google Scholar 

  • Stahl K (1982a) Differentiated products, consumer search, and locational oligopoly. J Ind Econ 31:97–113

    Article  Google Scholar 

  • Stahl K (1982b) Location and spatial pricing theory with nonconvex transportation cost schedules. Bell J Econ 13:575–582

    Article  Google Scholar 

  • Tabuchi T, Thisse J-F (1995) Asymmetric equilibria in spatial competition. Int J Ind Organ 13:213–227

    Article  Google Scholar 

  • Wolinsky A (1983) Retail trade concentration due to consumers’ imperfect information. Bell J Econ 14:275–282

    Article  Google Scholar 

Download references

Acknowledgments

I would like to thank two anonymous referees for helpful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ralph M. Braid.

Appendix

Appendix

This Appendix derives Eqs. (13) and (14), which are used in Propositions 7 and 8 of Sect. 7. In the full-information case of Sect. 3, if the two stores (libraries) are at the locations described in Proposition 1, it follows by substituting these locations into (3), multiplying by tD and adding 2\(f\), that aggregate social costs are

$$\begin{aligned} \hbox {ASC} = 2f + (\mathrm{wtD}/2)[2 - w - (1/2)w^2]. \end{aligned}$$
(15)

If, on the other hand, there is a single combined store (library) at location 0.5, the aggregate expected travel distance is the probability that a consumer makes a trip, which is \(2w - w^{2}\), multiplied by the average round-trip travel distance of consumers who do make the trip, which is 0.5. Thus, aggregate social costs are simply

$$\begin{aligned} \hbox {ASC}= F + (\mathrm{tD}/2)(2w - w^2) = F + (\mathrm{wtD}/2)(2 - w). \end{aligned}$$
(16)

Comparing (15) and (16), it follows that one single combined store in the middle is better than two separate stores if and only if

$$\begin{aligned} (2f - F)/(\mathrm{tD}) > (1/4)w^3, \end{aligned}$$
(17)

which is the same as (13) of the text.

In the costly visit-search case of Sect. 4, if \(0 < w \le 0.5\) and the two stores (libraries) are at the locations described by (12) in Proposition 2, it follows by substituting these locations into (8), multiplying by tD and adding 2\(f\), that aggregate social costs are

$$\begin{aligned} \hbox {ASC} = 2 f + \mathrm{tD}(1/2) \quad \hbox {for}\; 0 < w \le 0.5. \end{aligned}$$
(18)

If there is a single combined store (library) at location 0.5, the aggregate expected travel distance is the probability that the consumer makes a trip, which is 1 (since consumers do not know ahead of time whether they will find what they want), multiplied by the average round-trip travel distance of consumers who do make the trip, which is 0.5. Aggregate social costs are thus

$$\begin{aligned} \hbox {ASC }= F + \mathrm{tD}(1/2). \end{aligned}$$
(19)

Comparing (18) and (19), the single combined store at location 0.5 is always better than two separate stores at the same location 0.5 as long as \(F < 2f\), which is an obvious result. Consider now the other case in Proposition 2. If \(0.5 \le w \le 1\), and if the two separate stores are at the locations described by (11) in Proposition 2, then substituting these locations into (8), multiplying by tD, and then adding 2\(f\), shows that aggregate social costs are

$$\begin{aligned} \hbox {ASC} = 2 f + (\mathrm{tD}/2)[(1 + w)^{-2} (5w + w^2 - 4w^3)]\quad \hbox {for}\; 0.5 \le w \le 1. \end{aligned}$$
(20)

If there is a single combined store at location 0.5, the aggregate social costs are given by (19). Comparing (19) and (20), it is seen that one single store in the middle is better than two separate stores if and only if

$$\begin{aligned} (2f - F)/(\mathrm{tD}) > (1/2) (1 + w)^{-2} (1 - 3w + 4w^{3})\quad \hbox {for}\; 0.5 \le w \le 1. \end{aligned}$$
(21)

which is the same as (14) of the text.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Braid, R.M. The socially optimal and equilibrium locations of two stores or libraries with consumer search. Ann Reg Sci 53, 123–136 (2014). https://doi.org/10.1007/s00168-014-0620-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00168-014-0620-6

JEL Classification

Navigation