Skip to main content
Log in

Consumer store choice in Asian markets

  • Published:
Marketing Letters Aims and scope Submit manuscript

Abstract

Consumers in Asian markets often need to make choices between highuncertainty stores and low-uncertainty stores. For example, an iPhone may be purchased from an Apple authorized retailer (a low-uncertainty store) or from an unauthorized seller (a high-uncertainty one). We build a game-theoretical model to capture the crucial roles of store uncertainty and consumer risk aversion in store choice. A model-based explanation is provided for the coexistence of charging a higher retail price and having a higher product demand. Interestingly, our finding shows that store uncertainty and risk aversion have the potential to enhance the overall market profitability under the full market coverage, whereas such profitability will be reduced under the partial coverage.Moreover, we find that consumer learning might not be beneficial to the market as a whole. While consumer learning increases the expected overall market profitability under the partial coverage, such learning actually decreases the expected overall profitability if the market is fully covered. Not only is the precision of signals during learning relevant, but also the number of signals per se plays an indispensable role in the market. In addition, we find that the second-mover advantage possibly enjoyed by a high-uncertainty store is an important force that changes market outcomes.

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. More generally, the low-uncertainty stores are the more trusted and more reliable sellers perceived by consumers. There are some examples of the low-uncertainty stores in Asian markets: official stores and flagship stores of Tmall related to Taobao in China; Reliance Fresh, Shoppers Stop, and Pantaloons Retail in India; and Hypermart and Matahari Department Store in Indonesia. However, the less reliable and relatively low-end retailers where shoppers expect additional uncertainties in the values of products offered could be considered as high-uncertainty stores.

  2. The model of our study is developed based on the store uncertainty that is salient in Asia, where the huge population and market size, the largely immature nature of the retail markets, and risk-averse cultures make the research issue become relevant and substantively important. Also, more generally, the essential idea of this model is applicable to a larger set of circumstances such as service uncertainty and assortment uncertainty from trip to trip. We thank the guest editor and an anonymous reviewer for this insight.

  3. If the product uncertainty is higher at the HS, we can merge the incremental φ, which then indicates the overall store-induced uncertainty at the HS. Furthermore, if the σ 2 is zero, there is no uncertainty at the LS, and this is a special case of our more general setting. Assuming that σ 2 is zero does not change our main results. We thank an anonymous reviewer for suggesting this point.

  4. The reason is that, for the marginal consumer located at x 1 (as being analyzed later), \(\frac {\partial x_{1}}{\partial p_{2}}=\frac {1}{2t}\). That is, for the same price increase at store 2, a higher t means that a lower incremental change in the demand at store 1, holding other things being unchanged.

  5. Our goal is to explain what effect of uncertainties related to consumer store choice has on the retail market. Main qualitative insights, however, remain unchanged when a non-zero product cost is introduced.

  6. Notice that we assume that φ r < 6t, such that \(p_{2}^{\ast }>0\) and \( q_{2}^{\ast }>0\). Otherwise, the HS cannot charge a positive price and will get zero demand. Hence, t cannot be zero and must be greater than \(\frac { \varphi r}{6}\).

  7. This study focuses on the scenario of consumer learning that influences the expected product value and the variance concerning the store uncertainty at the HS. The related derivations as well as how the profit change depends the various model parameters are in the Appendix.

References

  • Bagwell, L. S., & Bernheim, B. D. (1996). Veblen effects in a theory of conspicuous consumption. American Economic Review, 86(3), 349–373.

    Google Scholar 

  • Baker, J., Parasuraman, A., Grewal, D., & Voss, G. B. (2002). The influence of multiple store environment cues on perceived merchandise value and patronage intentions. Journal of Marketing, 66(2), 120–141.

    Article  Google Scholar 

  • Balasubramanian, S. (1998). Mail versus mall: a strategic analysis of competition between direct marketers and conventional retailers. Marketing Science, 17(3), 181–195.

    Article  Google Scholar 

  • BBC (2014). China’s gigantic market and its growing firms. Business News, May 7.

  • Benjamin, D., Choi, J., & Strickland, A.J. (2010). Social identity and preferences. American Economic Review, 100(4), 1913–1928.

    Article  Google Scholar 

  • Bettman, J.R., Luce, M.F., & Payne, J.W. (1998). Constructive consumer choice processes. Journal of Consumer Research, 25(3), 187–217.

    Article  Google Scholar 

  • Bhardwaj, P., Chen, Y., & Godes, D. (2008). Buyer-initiated vs. seller-initiated information revelation. Marketing Science, 54(6), 1104–1114.

    Google Scholar 

  • Bloomberg Businessweek (2013). Why foreign retailers stumble in China. September 6. Accessed May 2, 2014. www.businessweek.com/articles/2013-09-06/why-foreign-retailers-stumble-in-china.

  • Briesch, R., Chintagunta, P., & Fox, E. (2009). How does assortment affect grocery store choice? Journal of Marketing Research, 46(2), 176–189.

    Article  Google Scholar 

  • Briesch, R., Dillon, W., & Fox, E. (2013). Category positioning and store choice: the role of destination categories. Marketing Science, 32(3), 488–509.

    Article  Google Scholar 

  • Chen, Y., & Xie, J. (2005). Third-party product review and firm marketing strategy. Marketing Science, 24(2), 218–240.

    Article  Google Scholar 

  • Chen, Y., & Xie, J. (2008). Online consumer review: word-of-mouth as a new element of marketing communications mix. Management Science, 54(3), 477–491.

    Article  Google Scholar 

  • Chen, T., Kalra, A., & Sun, B. (2009). Why do consumers buy extended service contracts? Journal of Consumer Research, 36(4), 611–623.

    Article  Google Scholar 

  • Chintagunta, P, Goettler, R., & Kim, M. (2012). New drug diffusion when forward-looking physicians learn from patient feedback and detailing. Journal of Marketing Research, 49(6), 807–821.

    Article  Google Scholar 

  • Erdem, T., & Keane, M. (1996). Decision-making under uncertainty: capturing dynamic brand choice processes in turbulent consumer goods markets. Marketing Science, 15(1), 1–20.

    Article  Google Scholar 

  • Forbes (2010). How eBay failed in China. Accessed May 1, 2014. www.forbes.com/sites/china/2010/09/12/how-ebay-failed-in-china.

  • Forehand, M.R., & Deshpande, R. (2001). What we see makes us who we are: priming ethnic self-awareness and advertising response. Journal of Marketing Research, 38(3), 336–348.

    Article  Google Scholar 

  • Fotheringham, A.S. (1988). Consumer store choice and choice set definition. Marketing Science, 7(3), 299–310.

    Article  Google Scholar 

  • India Retail Report (2011). Business Environment Outlook, 2(4), 31–33.

    Google Scholar 

  • Johnson, J.P., & Myatt, D.P. (2003). Multiproduct quality competition: fighting brands and product line pruning. American Economic Review, 93(3), 748–774.

    Article  Google Scholar 

  • Kalra, A., & Shi, M. (2010). Consumer value-maximizing sweepstakes and contests. Journal of Marketing Research, 47(2), 287–300.

    Article  Google Scholar 

  • Li, L. (1985). Cournot oligopoly with information sharing. RAND Journal of Economics, 16(4), 521–536.

    Article  Google Scholar 

  • Moe, W.W., & Schweidel, D.A. (2012). Online product opinions: incidence, evaluation, and evolution. Marketing Science, 31(3), 372–386.

    Article  Google Scholar 

  • Okada, E.M., & Hoch, S.J. (2004). Spending time versus spending money. Journal of Consumer Research, 31(2), 313–323.

    Article  Google Scholar 

  • PwC Report (2013). 2013 Outlook for the retail and consumer products sector in Asia. PricewaterhouseCoopers.

  • Roberts, J., & Urban, G. (1988). Modeling multiattribute utility, risk, and belief dynamics for new consumer durable brand choice. Management Science, 34(2), 167–185.

    Article  Google Scholar 

  • Shaffer, G., & Zettelmeyer, F. (2004). Advertising in a distribution channel. Marketing Science, 23(4), 619–628.

    Article  Google Scholar 

  • Sun, M. (2012). How does the variance of product ratings matter? Management Science, 58(4), 696–707.

    Article  Google Scholar 

  • Veblen, T.B. (1899). The theory of the leisure class: an economic study of institutions. London: Unwin Books; reprinted New York: Dover Publications, 1994.

  • Wall Street Journal (2014). The dos and don’ts of shopping on Alibaba’s Taobao. WSJ Digits, April 27.

  • Wang, W., Keh, H., & Bolton, L. (2010). Lay theories of medicine and a healthy lifestyle. Journal of Consumer Research, 37(1), 80–97.

    Article  Google Scholar 

  • Wang, Y., Bell, D.R., & Padmanabhan, V. (2009). Manufacturer-owned retail stores. Marketing Letters, 20(2), 107–124.

    Article  Google Scholar 

  • Wu, X., & Zhang, F. (2014). Home or overseas? An analysis of sourcing strategies under competition. Management Science, 60(5), 1223–1240.

    Article  Google Scholar 

  • Zhao, Y., Yang, S., Narayan, V., & Zhao, Y. (2013). Modeling consumer learning from online product reviews. Marketing Science, 32(1), 153–169.

    Article  Google Scholar 

Download references

Acknowledgments

The authors thank Yuxin Chen, Xiucheng Fan, Qingyun Jiang, Yi Liu, Seshan Ramaswami, Guohua Wan, Z. John Zhang, the guest editor and anonymous reviewers for their insights on model building and for their helpful comments and suggestions. This research is supported by Humanity and Social Science Research Foundation of Ministry of Education of China (10YJC630263).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yusong Wang.

Appendix

Appendix

1.1 Proof of Lemma 1

Adding the two first-order conditions (the second-order conditions are checked as well), we have p 1 + p 2 = 2t; and subtracting leads to \( p_{1}-p_{2}=\frac {\varphi r}{3} .\) Therefore, the equilibrium prices are \( p_{1}^{\ast }=t+\frac {\varphi r}{6}\) at the LS, and \(p_{2}^{\ast }=t-\frac { \varphi r}{6}\) at the other one. Hence, \(p_{1}^{\ast }\) increases with t and φ r, whereas \(p_{2}^{\ast }\) increases with t but decreases with φ r. Here, we have the price premium \(p_{1}^{\ast }>p_{2}^{\ast }\) and \(p_{1}^{\ast }-p_{2}^{\ast }=\frac {\varphi r}{3}\).

Based on the results of \(p_{1}^{\ast }\) and \(p_{2}^{\ast }\) , the equilibrium demands for store 1 and store 2 are \(q_{1}^{\ast }=\frac {1}{2t} (t+\frac {\varphi r}{6})\) and \(q_{2}^{\ast }=\frac {1}{2t}(t-\frac {\varphi r}{6 })\), respectively. Hence, \(q_{1}^{\ast }\) increases with φ r but decreases with t, whereas \(q_{2}^{\ast }\) increase with t but decreases with φ r. Here, the LS has a demand advantage \( q_{1}^{\ast }>q_{2}^{\ast }\), and \(q_{1}^{\ast }-q_{2}^{\ast }=\frac {\varphi r}{6t}\). Then, equilibrium profits can be obtained: \({\Pi }_{1}^{\ast }=\frac {1 }{2t}(t+\frac {\varphi r}{6})^{2}\) and \({\Pi }_{2}^{\ast }=\frac {1}{2t}(t-\frac { \varphi r}{6})^{2}\), where \({\Pi }_{1}^{\ast }>{\Pi }_{2}^{\ast }\) and \({\Pi }_{1}^{\ast }-{\Pi }_{2}^{\ast }=\frac {\varphi r}{3}\). \({\Pi }_{1}^{\ast }\) increases with t and φ r, whereas \({\Pi }_{2}^{\ast }\) increases with t but decreases with φ r. Notice that we focus on the more interesting case where φ r < 6t, such that the price and demand at store 2 are positive. However, if this condition is not satisfied (corresponding to the situation where the store uncertainty and the risk aversion are too high), there will be only one seller (i.e., store 1) in the market.

1.2 Proof of Proposition 1

  1. (a)

    Under the full market coverage. Assume that φ r < 6t, and denote the total profit with store uncertainty and risk aversion (i.e., φ r > 0) as \({\Pi }_{full}^{\ast }|_{\varphi r>0}={\Pi }_{1}^{\ast }+{\Pi } _{2}^{\ast }=\frac {1}{2t}(t+\frac {\varphi r}{6})^{2} +\frac {1}{2t}(t-\frac { \varphi r}{6})^{2}\). The total profit in the benchmark cases is \({\Pi }_{full}^{\ast }|_{\varphi r=0}=\frac {1}{2t}t^{2} +\frac {1}{2t}t^{2}=t\). Hence, \({\Delta }_{full}={\Pi }_{full}^{\ast }|_{\varphi r>0}-{\Pi }_{full}^{\ast }|_{\varphi r=0}=\frac {1}{t}(\frac {\varphi r}{6})^{2}>0\), where Δ f u l l increases with φ r and decreases with t.

  2. (b)

    Under the partial market coverage. Based on the locations of the marginal consumers, the demands are \(q_{1}=\widehat {x}_{1}=\frac {1}{t}(v- \frac {r}{2}\sigma ^{2}-p_{1})\) and \(q_{2}=\widehat {x}_{2}=\frac {1}{t}(v- \frac {r}{2}(\sigma ^{2}+\varphi )-p_{2})\). Solving the maximization problem of each store \(\max \limits _{p_{i}}q_{i}p_{i}\), we get the optimal retail prices as \(p_{1}=\frac {1}{2}(v-\frac {r}{2}\sigma ^{2})\) and \(p_{2}=\frac {1}{2 }(v-\frac {r}{2}(\sigma ^{2}+\varphi ))\). The demands and profits are \(q_{1}= \frac {1}{4t}\left (2v-\sigma ^{2}r\right ) \), \(q_{2}=\frac {1}{4t}(2v-\sigma ^{2}r-\varphi r)\), \({\Pi }_{1}=\frac {1}{16t}\left (2v-\sigma ^{2}r\right )^{2}\) and \({\Pi }_{2}=\frac {1}{16t}(2v-\sigma ^{2}r-\varphi r)^{2}\). Under the partial coverage, the condition q 1 + q 2 < 1 must be satisfied, and this means that \(v<\frac {r}{2}(\sigma ^{2}+\frac {1}{2}\varphi )+t\). Thus, we get the price premium p 1 > p 2 and demand advantage q 1 > q 2 under the partial coverage. The total profit for two stores under the partial coverage is \({\Pi }_{partial}^{\ast }=\frac {1}{16t}\left (2v-\sigma ^{2}r\right )^{2}+ \frac {1}{16t}(2v-\sigma ^{2}r-\varphi r)^{2}.\) Hence, the total profit under the partial coverage is lower that in the benchmark cases of φ r = 0: \( {\Delta }_{partial}={\Pi }_{partial}^{\ast }|_{\varphi r>0}-{\Pi }_{partial}^{\ast }|_{\varphi r=0}<0\), where |Δ p a r t i a l | increases with φ r and v, but decreases with t and σ 2. In the partially-covered retail market, notice that we will have two stores if \(v>\frac {r}{2}(\sigma ^{2}+\varphi )\), only store 1 if \(\frac {r}{2}\sigma ^{2}<v\leq \frac {r}{2} (\sigma ^{2}+\varphi )\), and no store if \(v\leq \frac {r}{2}\sigma ^{2}\).

1.3 Proof of Proposition 2

  1. (a)

    Under the full market coverage. Suppose that the marginal consumer who is indifferent between shopping at the two stores is located at \({x_{1}^{L}}\), where the superscript denotes the scenario with consumer learning. This means that

    $$v-\frac{r}{2}\sigma^{2}-t{x_{1}^{L}}-p_{1}=E[\widetilde{\widetilde{v}} |s_{1},{\ldots} ,s_{n}]-\frac{r}{2}V[\widetilde{\widetilde{v}}|s_{1},\ldots ,s_{n}]-t\left(1-{x_{1}^{L}}\right)-p_{2}. $$

    Hence, we have \({x_{1}^{L}}\)=\(\frac {1}{2t}(t+\frac {r}{2(1/\varphi +n/\sigma _{\varepsilon }^{2})}-\frac {n/\sigma _{\varepsilon }^{2}}{1/\varphi +n/\sigma _{\varepsilon }^{2}}\frac {{\Sigma }_{i=1}^{n}s_{i}}{n}-p_{1}+p_{2}).\) Since \(E[\frac {{\Sigma }_{i=1}^{n}s_{i}}{n}]=E[\widetilde {v}^{\prime }]=0\), the average value of signals \(\frac {{\Sigma }_{i=1}^{n}s_{i}}{n}\) is an unbiased estimator of the true value \(\widetilde {v}^{\prime }\), and then the expected location of \({x_{1}^{L}}\) is \(E[{x_{1}^{L}}]=\frac {1}{2t}(t+\frac { \varphi ^{L}r}{2}-p_{1}+p_{2})\), where we define \(\varphi ^{L}\equiv \frac {1 }{1/\varphi +n/\sigma _{\varepsilon }^{2}}<\varphi \).

    Following similar analyses as the basic model, we find that the expected price premium with learning \(E[p_{1}^{\ast L}-p_{2}^{\ast L}]=\frac {\varphi ^{L}r}{3}<p_{1}^{\ast }-p_{2}^{\ast }=\frac {\varphi r}{3}\) (the price premium without learning). The expected demand advantage with learning \( E[q_{1}^{\ast L}-q_{2}^{\ast L}]=\frac {\varphi ^{L}r}{6t}<\) \(q_{1}^{\ast }-q_{2}^{\ast }=\frac {\varphi r}{6t}\) (the demand advantage without learning). As \(\varphi ^{L}=\frac {1}{1/\varphi +n/\sigma _{\varepsilon }^{2}}\), the decreases in the price premium and demand advantage depend critically on the number of signals n and the corresponding signal precision \(1/\sigma _{\varepsilon }^{2}\).

    The expected total profit for the two stores with learning \(E[{\Pi }_{full}^{\ast L}]={\Pi }_{1}^{\ast L}+{\Pi }_{2}^{\ast L}=\frac {1}{2t}(t+\frac { \varphi ^{L}r}{6})^{2}\) \(+\frac {1}{2t}(t-\frac {\varphi ^{L}r}{6})^{2}<{\Pi }_{full}^{\ast }=\frac {1}{2t}(t+\frac {\varphi r}{6})^{2}\) \(+\frac {1}{2t}(t- \frac {\varphi r}{6})^{2}\) (the total profit without learning). Through partial differentiations, we get that the absolute value of the difference \( |E[{\Pi }_{full}^{\ast L}]-{\Pi }_{full}^{\ast }|\), which increases with φ, r, n, and \(1/\sigma _{\varepsilon }^{2},\) but decreases with t. Define \({\Delta }_{full}^{L}\equiv E[{\Pi }_{full}^{\ast L}|_{\varphi r>0}-{\Pi }_{full}^{\ast L}|_{\varphi r=0}]=\frac {1}{t}(\frac {\varphi ^{L}r}{6} )^{2}>0\), and we have \({\Delta }_{full}^{L}<\) Δ f u l l . Thus the qualitative result in Proposition 1 remains unchanged, while the improvement of the overall profitability gets smaller.

  2. (b)

    Under the partial market coverage. The expected price premium with learning is \(E[{p_{1}^{L}}-{p_{2}^{L}}]=\frac {1}{2}(v-\frac {r}{2}\sigma ^{2})-\frac {1}{2}(v-\frac {r}{2}(\sigma ^{2}+\varphi ^{L}))=\frac {\varphi ^{L}r}{4}\) which is less than the price premium without learning (i.e., \( \frac {\varphi r}{4}\)); and the expected demand advantage is \( E[{q_{1}^{L}}-{q_{2}^{L}}]=\frac {1}{2t}\left (v-\frac {r}{2}r\sigma ^{2}\right ) - \frac {1}{2t}(v-\frac {r}{2}(\sigma ^{2}+\varphi ^{L}))=\frac {\varphi ^{L}r}{4t }\) which is less than the demand advantage without learning (i.e., \(\frac { \varphi r}{4t}\)). Furthermore, the expected joint profit under the partial coverage with learning is \(E[{\Pi }_{partial}^{\ast L}]=\frac {\left (2v-r\sigma ^{2}\right )^{2}+(2v-r\sigma ^{2}-r\varphi ^{L})^{2}}{16t}\), which is greater than the joint profit without learning \({\Pi }_{partial}^{\ast }=\frac {\left (2v-r\sigma ^{2}\right )^{2}+(2v-r\sigma ^{2}-r\varphi )^{2}}{16t}\), as φ L < φ. Hence, we have \( E[{\Pi }_{partial}^{\ast L}]-{\Pi }_{partial}^{\ast }>0\). Through partial differentiations, we get that the difference \(E[{\Pi }_{partial}^{\ast L}]-{\Pi }_{partial}^{\ast }\) increases with n and \(1/\sigma _{\varepsilon }^{2},\) but decreases with σ 2 and t. In addition, \(\frac {\partial E[{\Pi } _{partial}^{\ast L}]-{\Pi }_{partial}^{\ast }}{\partial r}=\frac {\left (\varphi -\varphi ^{L}\right ) \left (2v-\varphi r-\varphi ^{L}r-2\sigma ^{2}r\right ) }{8t}\geq 0\) if r is sufficiently low, i.e., \(r\leq \frac {2v}{ 2\sigma ^{2}+\varphi +\varphi ^{L}}\). This implies that the learning tends to be most impactful when the consumers’ risk aversion is in an intermediate range. Define \({\Delta }_{partial}^{L}\equiv E[{\Pi }_{partial}^{\ast L}|_{\varphi r>0}-{\Pi }_{partial}^{\ast L}|_{\varphi r=0}]<0\) and we have \( |{\Delta }_{partial}^{L}|<|{\Delta }_{partial}|\). Thus the qualitative result in Proposition 1 remains unchanged, while the decrease of the overall profitability under the partial market coverage becomes less pronounced.

1.4 Proof of Proposition 3

We solve the Stackelberg game through backward induction. In the second stage of the game, store 2’s profit maximization problem is \( \max \limits _{p_{2}}q_{2}p_{2}=\frac {1}{2t}(t-\frac {r}{2}\varphi +p_{1}-p_{2})p_{2}\), where the first-order condition is \(t-\frac {r}{2} \varphi +p_{1}-2p_{2}=0\). We focus on the full market coverage situation, as the sequence of stores’ moves does not impact the market outcomes under the partial coverage. Hence, we have the best response function for store 2: \( p_{2}(p_{1})=\frac {1}{2}(t-\frac {r}{2}\varphi +p_{1})\). In the first stage of the game, store 1 sets its price p 1 to maximize its profit, taking into account store 2’s best response function:

$$\max\limits_{p_{1}}q_{1}p_{1}=\frac{1}{2t}\left(t+\frac{r}{2}\varphi -p_{1}+p_{2}(p_{1})\right)p_{1}, $$

where the first-order condition is \(3t+\frac {r}{2}\varphi -2p_{1}=0\). This leads to the equilibrium price \(p_{1}^{\ast }=\frac {3t}{2}+\frac {\varphi r}{4 }\). Hence, we get \(p_{2}^{\ast }=\frac {5t}{4}-\frac {\varphi r}{8}\), and then the price premium is \(p_{1}^{\ast }-p_{2}^{\ast }=\frac {t}{4}+\frac {3\varphi r}{8}\), which is greater than \(\frac {\varphi r}{3}\) (i.e., the price premium under Nash). Consequently, the equilibrium demands are \(q_{1}^{\ast }=\frac { 6t+\varphi r}{16t}\) and \(q_{2}^{\ast }=\frac {10t-\varphi r}{16t}\).

Notice that the price and demand at store 2 are positive when φ r < 10t under the Stackelberg structure (cf. φ r < 6t under Nash). The demand and price at store 2 are zero under Nash, if φ r≥6t. This implies that the HS is more likely to be financially viable as a follower under the Stackelberg structure than as a competitor under Nash, when φ r is quite high (i.e., 6tφ r < 10t). Then, we have: (a) the LS has demand advantage \(q_{1}^{\ast }-q_{2}^{\ast }=\frac {\varphi r-2t}{ 8t}>0\) if 2t < φ r < 6t, and (b) the HS has demand advantage \( q_{1}^{\ast }-q_{2}^{\ast }<0\) if φ r < 2t. For the special case when φ r = 2t, neither store has the advantage. In addition, we have \({\Pi }_{Stackelberg}^{\ast }=\) \(\frac {(6t+\varphi r)^{2}}{64t}+\frac {(10t-\varphi r)^{2}}{128t}\), thus we define

$${\Delta}_{S}\equiv {\Pi}_{Stackelberg}^{\ast }|_{\varphi r>0}-{\Pi}_{Stackelberg}^{\ast }|_{\varphi r=0}=\frac{\varphi r(4t+3\varphi r)}{128t}. $$

As we have proved previously that \({\Delta }_{N}\equiv {\Pi }_{Nash}^{\ast }|_{\varphi r>0}-{\Pi }_{Nash}^{\ast }|_{\varphi r=0}=\frac {1}{t}(\frac { \varphi r}{6})^{2}\), \({\Delta }_{S}-{\Delta }_{N}=\frac {\varphi r(36t-5\varphi r) }{1152t}\). Therefore, Δ S > Δ N if \(\varphi r<\frac {36t}{5}\) . Recall that φ r < 6t for \(p_{2}^{\ast }>\) 0 and \(q_{2}^{\ast }>0\) under the Nash game. Hence Δ S > Δ N for φ r < 6t.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, Y., Bell, D.R. Consumer store choice in Asian markets. Mark Lett 26, 293–308 (2015). https://doi.org/10.1007/s11002-015-9352-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11002-015-9352-3

Keywords

Navigation