Skip to main content
Log in

Competition, innovation, and the effect of R&D knowledge

  • Published:
Journal of Economics Aims and scope Submit manuscript

Abstract

This paper analyzes the effect of learning-by-doing in R&D on firms’ incentives to innovate. As a benchmark without learning it is shown that relaxing the usual assumption of imposed imitation yields additional strategic effects. Therefore, the leader’s R&D effort increases with the gap as she is trying to avoid competition in the future. When firms gain experience by performing R&D technological leaders rest on their laurels allowing followers to catch up. Contrary to the benchmark case, the leader’s innovation effort declines with the lead. This causes an equilibrium where the incentives to innovate are highest when competition is most intense.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. For an interesting depiction of early fundamentally statements on determinants of innovation see Caballero and Jaffe (1993).

  2. A brief review of literature is provided at the end of this section.

  3. This way of modeling is based upon the work of Doraszelski (2003).

  4. Note that learning-by-doing and organizational forgetting in R&D is not the same as learning-by-doing and organizational forgetting in production which is a well established approach in literature (see e.g. Cabral and Riordan 1994, 1997). In contrast to the latter, in our framework firms learn by doing research and development not by producing.

  5. See Doraszelski (2003) for empirical findings on knowledge accumulation and e.g. Gulledge and Womer (1986) for a comprehensive early work on learning-by-doing.

  6. See for example Argote et al. (1990) and Benkard (2000).

  7. This escape competition motive has been pointed out in previous theoretical papers on innovation, see e.g. Mookherjee and Ray (1991) for an early work.

  8. Vickers (2010) presents a good overview of how dynamic considerations affect innovation incentives when there is R&D competition.

  9. For example, Balasubramanian and Lieberman (2011) empirically estimate the relation of learning-by-doing on market structure. Learning-by-doing is also a common effect in the general-equilibrium models of endogenous growth. Early accordant works include Romer (1990) and Jones (1995).

  10. See for example Fudenberg et al. (1983), Harris and Vickers (1985), Grossman and Shapiro (1987), Harris and Vickers (1987), and Lippman and McCardle (1988).

  11. Acemoglu and Akcigit (2012) also consider the case where the follower might even be able to improve over the frontier technology.

  12. Extending the derived results to the more general case of differentiated goods would be possible at the cost of additional notation and a considerably higher complexity in derivation. In a differentiated goods model the innovation incentive is also determined by the difference in profits between the leader and the laggard. This difference in profits shows the same main characteristics with homogeneous and heterogeneous goods (see Fig. 1). Hence, only minor additional insights can be gained by such an extension as long as the degree of substitution is exogenous such that we restrict attention to the case of perfect substitutes.

  13. The stepsize is arbitrarily set equal to one. As long as the size is exogenous and constant all results remain unchanged with a different increment. However, allowing for different possible stepsizes of innovations may alter the outcome considerably.

  14. Note that due to this way of modeling we cannot interpret knowledge as capital in the usual way since knowledge is not an input factor in production and knowledge as such does not directly influence the production technology.

  15. For detailed information on Poisson processes see Ross (2007).

  16. For the sake of readability throughout the rest of the paper we will denote firm \(i\)’s competitor by \(j\), i.e., \(j\ne i\) will always hold. Besides, we suppress the indication of time where not necessary.

  17. A discussion on how realistic these assumptions are is given at the end of this section.

  18. This result is derived in Appendix A.

  19. However, this assumption is not crucial for the following results and just assures tractability.

  20. For the sake of readability we will suppress the identity of the firm where not necessary.

  21. As shown by Dutta (1991), in a model like ours the assumption of zero discounting is not crucial for the results. model discounted optimal policies and values converge to undiscounted optima. Besides, we were able to show numerically that the results basically hold with \(\rho >0\) in quality (with the additional feature that R&D investment eventually falls to zero). However, the analytical derivation for the discounted problem is excessively more complex.

  22. Note, however, that this result is not necessarily the only possible pattern, i.e., it is not possible to show uniqueness (see Appendix).

  23. It is exactly the strengthening of this avoid competition effect that drives the results of Acemoglu and Akcigit (2012).

  24. This result of the model is in line with recent economic analysis showing that stronger intellectual property rights not necessarily spur innovation (see e.g. Vickers 2010). The effects at place here are conflictive: On the one hand, imitation diminishes the strategic incentives of R&D, i.e. the Schumpeterian effect for the laggard and the avoid competition effect for the leader. So the incentives to invest would decrease. On the other hand, imitation makes catching up easier and thus the industry shows more often a neck-and-neck state with strong R&D competition and innovation incentives due to the escape-competition-effect. Thus, the overall outcome is not clear.

  25. See Acemoglu and Akcigit (2012) for a formal proof on the existence of a steady state.

  26. Although we do not model an entire closed economy and cannot provide a general equilibrium analysis, our model can easily be transferred into such a framework. Thus, we can draw conclusions on the economy’s growth rate from the growth rate of the industry or sector. In a general equilibrium framework with an economy consisting of a mass of 1 identical industries, the defined industry growth rate \(g\) equals the growth rate of the economy \(\frac{d \ln Y}{dt}\) with \(Y\) as the economy’s aggregate output.

  27. Note that different to technology levels there is no direct effect of the competitor’s knowledge on optimal investment.

  28. This relatively new, mainly macroeconomic approach jointly analyzes short-run instability and long-run growth due to innovations. Important papers in this strand of literature include for example Bental and Peled (1996), Matsuyama (1999), Francois and Lloyd-Ellis (2003), Maliar and Maliar (2004), Gabaix (2011), and Haruyama (2009). In these macroeconomic models, however, the motivation for fluctuations in aggregate growth and the link to long run growth are important issues that are irrelevant to our model.

  29. Again, we suppress the identification of the identity of the firm.

  30. The hazard rate being a linear (\(\alpha =1\)) function of knowledge is a special case where the function described by (13) is a horizontal line. The results are similar to the described cases and therefore not given in detail.

  31. For the very special case of \(\alpha =\eta \) and \(\left( \frac{\delta +\rho }{\alpha \gamma \Phi (\Delta )}\right) ^{\frac{\eta }{1-\eta }}=\delta ^{\eta }\) the two functions are identical and we have an infinite number of steady states.

  32. See for example Culbertson and Mueller (1985) on food-manufacturing industries and Lunn (1986), Lunn and Martin (1986), MacDonald (1994), Nickell (1996) and Tang (2006) on a variety of aggregated manufacturing industries. Additionally, in his survey, Gilbert (2006) concludes that “there is some evidence that competition promotes innovation when the measure of competition is an index of proximity of firms to a technological frontier.” This is the case in our theoretical framework.

  33. For the special case \(\eta =2\) effort is even strictly proportional to the incremental value.

  34. For the sake of readability we will suppress the identity of the firm where not necessary.

References

  • Acemoglu D, Akcigit U (2012) Intellectual property rights policy, competition and innovation. J Eur Econ Assoc 10(1):1–42

    Article  Google Scholar 

  • Aghion P, Harris C, Howitt P, Vickers J (2001) Competition, imitation and growth with step-by-step innovation. Rev Econ Stud 68(3):467–492

    Article  Google Scholar 

  • Aghion P, Bloom N, Blundell R, Griffith R, Howitt P (2005) Competition and innovation: an inverted-U relationship. Q J Econ 120(2):701–728

  • Argote L, Beckman SL, Epple D (1990) The persistence and transfer of learning in industrial settings. Manag Sci 36(2):140–154

    Article  Google Scholar 

  • Balasubramanian N, Lieberman MB (2011) Learning-by-doing and market structure. J Ind Econ 59(2):177–198

    Article  Google Scholar 

  • Belleflamme P, Vergari C (2011) Incentives to innovate in oligopolies. Manch School 79(1):6–28

    Article  Google Scholar 

  • Benkard CL (2000) Learning and forgetting: the dynamics of aircraft production. Am Econ Rev 90(4):1034–1054

    Article  Google Scholar 

  • Bental B, Peled D (1996) The accumulation of wealth and the cyclical generation of new technologies: a search theoretic approach. Int Econ Rev 37(3):687–718

    Article  Google Scholar 

  • De Bondt R, Vandekerckhove J (2012) Reflections on the relation between competition and innovation. J Ind Compet Trade 12:7–19

    Article  Google Scholar 

  • Budd C, Harris C, Vickers J (1993) A model of the evolution of duopoly: does the asymmetry between firms tend to increase or decrease? Rev Econ Stud 60(3):543–573

    Article  Google Scholar 

  • Caballero RJ, Jaffe AB (1993) How high are the giants’ shoulders: an empirical assessment of knowledge spillovers and creative destruction in a model of economic growth. In: Nber macroeconomics annual 1993, vol 8. MIT press, Cambridge, pp 15–86

  • Cabral LMB, Riordan MH (1994) The learning curve, market dominance, and predatory pricing. Econometrica 62(5):1115–1140

    Article  Google Scholar 

  • Cabral LMB, Riordan MH (1997) The learning curve, predation, antitrust, and welfare. J Ind Econ 45(2):155–169

    Article  Google Scholar 

  • Cohen WM, Levin RC (1989) Empirical studies of innovation and market structure. In: Schmalensee R, Willig R (eds) Handbook of Industrial Organization (vol 2). Elsevier, New York, pp 1059–1107

  • Culbertson J, Mueller W (1985) The Influence of market structure on technological performance in the food-manufacturing industries. Rev Ind Organ 2(1):40–54

    Article  Google Scholar 

  • Czarnitzki D, Kraft K (2004) An empirical test of the asymmetric models on innovative activity: who invests more into R&D, the incumbent or the challenger? J Econ Behav Organ 54(2):153–173

    Article  Google Scholar 

  • Dasgupta P, Stiglitz JE (1988) Learning-by-doing, market structure and industrial and trade policies. Oxford Econ Papers 40(2):246–268

    Google Scholar 

  • Denicolò V, Zanchettin P (2012) Leadership cycles in a quality-ladder model of endogenous growth. Econ J 122(561):618–650

    Article  Google Scholar 

  • Doraszelski U (2003) An R&D race with knowledge accumulation. RAND J Econ 34(1):20–42

    Article  Google Scholar 

  • Dutta PK (1991) What do discounted optima converge to? A theory of discount rate asymptotics in economic models. J Econ Theory 55(1):64–94

    Article  Google Scholar 

  • Francois P, Lloyd-Ellis H (2003) Animal spirits through creative destruction. Am Econ Rev 93(3):530–550

    Article  Google Scholar 

  • Fudenberg D, Gilbert R, Stiglitz J, Tirole J (1983) Preemption, leapfrogging and competition in patent races. Eur Econ Rev 22(1):3–31

    Article  Google Scholar 

  • Gabaix X (2011) The granular origins of aggregate fluctuations. Econometrica 79(3):733–772

    Article  Google Scholar 

  • Gilbert R (2006) Looking for Mr. Schumpeter: where are we in the competition-innovation debate? NBER Innov Policy Econ 6(1):159–215

  • Grossman GM, Shapiro C (1987) Dynamic R&D competition. Econ J 97:372–387

    Article  Google Scholar 

  • Gulledge TR, Womer NK (1986) The economics of made-to-order production. Springer, Berlin

  • Harris C, Vickers J (1985) Perfect equilibrium in a model of a race. Rev Econ Stud 52:193–209

    Article  Google Scholar 

  • Harris C, Vickers J (1987) Racing with uncertainty. Rev Econ Stud 54:1–22

    Article  Google Scholar 

  • Haruyama T (2009) R&D policy in a volatile economy. J Econ Dyn Control 33(10):1761–1778

    Article  Google Scholar 

  • Hörner J (2004) A perpetual race to stay ahead. Rev Econ Stud 71:1064–1088

    Article  Google Scholar 

  • Jones C (1995) R&D-based models of economic growth. J Polit Econ 103(4):759–784

    Article  Google Scholar 

  • Khanna T (1995) Racing behavior technological evolution in the high-end computer industry. Res Policy 24(6):933–958

    Article  Google Scholar 

  • Lerner J (1997) An empirical exploration of a technology race. RAND J Econ 28(2):228–247

    Article  Google Scholar 

  • Lippman SA, McCardle KF (1988) Preemption in R&D races. Eur Econ Rev 32(8):1661–1669

    Article  Google Scholar 

  • Lunn J (1986) An empirical analysis of process and product patenting: a simultaneous equation framework. J Ind Econ 34(3):319–330

    Article  Google Scholar 

  • Lunn J, Martin S (1986) Market structure, firm structure, and research and development. Q Rev Econ Bus 26(1):31–44

    Google Scholar 

  • MacDonald JM (1994) Does import competition force efficient production? Rev Econ Stat 76(4):721–727

    Article  Google Scholar 

  • Maliar L, Maliar S (2004) Endogenous growth and endogenous business cycles. Macroecon Dyn 8(05):559–581

    Google Scholar 

  • Matsuyama K (1999) Growing through cycles. Econometrica 67(2):335–347

    Article  Google Scholar 

  • Mookherjee D, Ray D (1991) On the competitive pressure created by the diffusion of innovations. J Econ Theory 54(1):124–147

    Article  Google Scholar 

  • Nickell SJ (1996) Competition and corporate performance. J Polit Econ 104(4):724–746

    Article  Google Scholar 

  • Reinganum JF (1989) The timing of innovation: research, development, and diffusion. In: Schmalensee R, Willig R (eds) Handbook of industrial organization (vol. 1). Elsevier, New York, pp 849–908

  • Romer PM (1990) Endogenous technological change. J Polit Econ 98(5):71–102

    Article  Google Scholar 

  • Ross SM (2007) Introduction to probability models, 9th edn. Academic Press, Amsterdam

    Google Scholar 

  • Segal I, Whinston MD (2007) Antitrust in innovative industries. Am Econ Rev 97(5):1703–1730

    Article  Google Scholar 

  • Sutton J (2007) Market structure: theory and evidence. In: Armstrong M, Porter R (eds) Handbook of industrial organization (vol 3). Elsevier, New York, pp 2301–2368

  • Sutton J (1991) Sunk costs and market structure. MIT Press, Cambridge

    Google Scholar 

  • Sutton J (1998) Technology and market structure: theory and history. MIT Press, Cambridge

    Google Scholar 

  • Tang J (2006) Competition and innovation behaviour. Res Policy 35(1):68–82

    Article  Google Scholar 

  • Vickers J (2010) Competition policy and property rights. Econ J 120(544):375–392

    Article  Google Scholar 

  • Vives X (2008) Innovation and competitive pressure. J Ind Econ 56(3):419–469

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander Steinmetz.

Additional information

I am grateful to Norbert Schulz for continuously insightful guidance and encouragement. I also thank Monika Schnitzer and Víctor Aguirregabiria for valuable advice and useful suggestions. The paper benefited from discussions at numerous conferences and seminars at the Universities of Toronto, Munich and Würzburg.

Appendix

Appendix

A The general model

To solve for the Markov-stationary equilibrium we use dynamic programming methods and therefore derive the Bellman equations. Defining the optimal programs for the firms \(i=1,2\) as \(V_i(s)\equiv \max _{\{z_i(\tau )\}}\Pi _i(s(t))\) s.t. the evolutions of the state variables \(s\equiv (k_1,k_2,x_1,x_2)\), the Bellman equations are given by

$$\begin{aligned} \rho V_i (s(t))&= \max _{z_i(t)}\left\{ \pi _i(s(t))-z_{i}(t)+\frac{1}{dt} E_t d V_i(s(t))\right\} , \end{aligned}$$
(15)

where the R&D effort of the competitor is taken as given. Given this general form we compute the differential \(d V_i(s(t))\) given the evolutions of the state variables and form expectations. This yields

$$\begin{aligned} E_t d V_i(s(t))&= \left[ \frac{\partial V_i(s)}{\partial k_i}\left( u(z_i)-\delta k_i\right) +\frac{\partial V_i(s)}{\partial k_{j}}\left( u(z_{j})-\delta k_{j}\right) \right] dt\\&+\big [V_i(x_i+1,\cdot )\!-\!V_i(s)\big ]h_i(z_i)dt\!+\!\left[ V_i(x_{j}\!+\!1,\cdot )\!-\!V_i(s)\right] h_{j}(z_{j})dt. \end{aligned}$$

The Bellman equations therefore read

$$\begin{aligned} \rho V_i (s(t))\!&= \!\max _{z_i(t)} \biggl \{\pi _i(s(t))\!-\!z_{i}(t)\!+\!\left[ \frac{\partial V_i(s)}{\partial k_i}\left( u(z_i)\!-\!\delta k_i\right) \!+\!\frac{\partial V_i(s)}{\partial k_{j}}\left( u(z_{j})\!-\!\delta k_{j}\right) \right] \nonumber \\&+\big [V_i(x_i+1,\cdot )-V_i(s)\big ]h_i(z_i)+\left[ V_i(x_{j}+1,\cdot )-V_i(s)\right] h_{j}(z_{j})\biggl \}.\nonumber \\ \end{aligned}$$
(16)

Then, the first-order conditions are

$$\begin{aligned}&-1+\frac{\partial V_i(s)}{\partial k_i}u'(z_i)+\big [V_i(x_i+1,\cdot )-V_i(s)\big ]h_i'(z_i)\mathop {=}\limits ^{!}0\nonumber \\&\quad \Leftrightarrow \frac{\partial V_i(s)}{\partial k_i}= \frac{1-\lambda u'(z_i)\big [V_i(x_i+1,\cdot )-V_i(s)\big ]}{u'(z_i)}. \end{aligned}$$
(17)

Current gain from not investing an additional unit, i.e., \(-1\) , must equal future gain from an additional unit of investment which is influenced by the change of knowledge stock (through the increase in effort) and the probability of a successful innovation.

This yields

$$\begin{aligned} d \frac{\partial V_i(s)}{\partial k_i}&= d \frac{1-\lambda u'(z_i)\big [V_i(x_i+1,\cdot )-V_i(s)\big ]}{u'(z_i)}\nonumber \\&= d \frac{1}{u'(z_i)}- \lambda d \big [V_i(x_i+1,\cdot )-V_i(s)\big ]. \end{aligned}$$
(18)

In the next step, we state the maximized Bellman equations from (16) as the Bellman equations where controls are replaced by their optimal values:

$$\begin{aligned} \rho V_i (s)&= \pi _i(s)-z_i(s)+\frac{\partial V_i(s)}{\partial k_i}\big (u(z_i(s))-\delta k_i\big ) +\frac{\partial V_i(s)}{\partial k_{j}}\left( u(z_{j}(s))-\delta k_{j}\right) \nonumber \\&+\big [V_i(x_i+1,\cdot )-V_i(s)\big ]h_i(z_i(s))+\left[ V_i(x_{j}+1,\cdot )-V_i(s)\right] h_{j}(z_{j}(s)).\nonumber \\ \end{aligned}$$
(19)

We now compute the derivatives with respect to the state variables \(k_i\) using the envelope theorem on the Bellman equation (16). This gives expressions for the shadow prices \(\frac{\partial V_i(s)}{\partial k_{i}}\):

$$\begin{aligned} \rho \frac{\partial V_i(s)}{\partial k_{i}}&=\frac{\partial ^2 V_i(s)}{\partial k_i^2}\big (u(z_i(s))-\delta k_i\big )-\delta \frac{\partial V_i(s)}{\partial k_{i}} +\frac{\partial ^2 V_i(s)}{\partial k_{i}\partial k_j }\left( u(z_{j}(s))-\delta k_{j}\right) \nonumber \\&\quad +\left[ \frac{\partial V_i(x_i\!+\!1,\cdot )}{\partial k_i}\!-\!\frac{\partial V_i(s)}{\partial k_{i}}\right] h_i(z_i(s))\!+\!\big [V_i(x_i\!+\!1,\cdot )\!-\!V_i(s)\big ]\alpha \gamma k_i^{\alpha -1}\nonumber \\&\quad + \left[ \frac{\partial V_i(x_{j}+1,\cdot )}{\partial k_i}-\frac{\partial V_i(s)}{\partial k_{i}}\right] h_{j}(z_{j}(s)). \end{aligned}$$
(20)

Furthermore,

$$\begin{aligned} \rho \frac{\partial V_i(s)}{\partial k_{j}}&= \frac{\partial ^2 V_i(s)}{\partial k_i \partial k_j}\big (u(z_i(s))-\delta k_i\big )+\frac{\partial ^2 V_i(s)}{\partial k_{j}^2}\left( u(z_{j}(s))-\delta k_{j}\right) -\delta \frac{\partial V_i(s)}{\partial k_{j}} \nonumber \\&+\left[ \frac{\partial V_i(x_i+1,\cdot )}{\partial k_j}-\frac{\partial V_i(s)}{\partial k_{j}}\right] h_i(z_i(s)) \nonumber \\&+\left[ \frac{\partial V_i(x_{j}+1,\cdot )}{\partial k_j}-\frac{\partial V_i(s)}{\partial k_{j}}\right] h_{j}(z_{j}(s))\nonumber \\&+\left[ V_i(x_j+1,\cdot )-V_i(s)\right] \alpha \gamma k_j^{\alpha -1}. \end{aligned}$$
(21)

Given the evolutions of the state variables we can compute the differentials of the shadow prices:

$$\begin{aligned} d \frac{\partial V_i(s)}{\partial k_{i}}&= \left[ \frac{\partial ^2 V_i(s)}{\partial k_i^2}\big (u(z_i)-\delta k_i\big ) +\frac{\partial ^2 V_i(s)}{\partial k_i \partial k_j}\left( u(z_{j})-\delta k_{j}\right) \right] dt \nonumber \\&+\left[ \frac{\partial V_i(x_i\!+\!1,\cdot )}{\partial k_i}\!-\!\frac{\partial V_i(s)}{\partial k_i}\right] dq_i(s) \!+\!\left[ \frac{\partial V_i(x_j\!+\!1,\cdot )}{\partial k_i}\!-\!\frac{\partial V_i(s)}{\partial k_i}\right] dq_j(s),\nonumber \\ \end{aligned}$$
(22)

and

$$\begin{aligned} d \frac{\partial V_i(s)}{\partial k_{j}}&= \left[ \frac{\partial ^2 V_i(s)}{\partial k_i \partial k_j}\big (u(z_i)-\delta k_i\big ) +\frac{\partial ^2 V_i(s)}{\partial k_j^2}\left( u(z_{j})-\delta k_{j}\right) \right] dt \nonumber \\&+\left[ \frac{\partial V_i(x_i\!+\!1,\cdot )}{\partial k_j}\!-\!\frac{\partial V_i(s)}{\partial k_j}\right] dq_i(s) \!+\!\left[ \frac{\partial V_i(x_j\!+\!1,\cdot )}{\partial k_j}\!-\!\frac{\partial V_i(s)}{\partial k_j}\right] dq_j(s),\nonumber \\ \end{aligned}$$
(23)

Replacing \(\frac{\partial ^2 V_i(s)}{\partial k_i^2}\big (u(z_i)-\delta k_i\big ) +\frac{\partial ^2 V_i(s)}{\partial k_i \partial k_i}\big (u(z_{j})-\delta k_{j}\big )\) in (22) by the same expressions from (20) gives

$$\begin{aligned} d \frac{\partial V_i(s)}{\partial k_{i}}&= \biggl \{(\rho +\delta ) \frac{\partial V_i(s)}{\partial k_{i}}-\left[ \frac{\partial V_1(x_i+1,\cdot )}{\partial k_i}-\frac{\partial V_i(s)}{\partial k_{i}}\right] h_i(z_i(s))\nonumber \\&-\left[ \frac{\partial V_i(x_{j}+1,\cdot )}{\partial k_i}-\frac{\partial V_i(s)}{\partial k_{1}}\right] h_{j}(z_{j}(s))\nonumber \\&-\left[ V_i(x_i+1,\cdot )-V_i(s)\right] \alpha \gamma k_i^{\alpha -1}\biggl \}dt \nonumber \\&+\left[ \frac{\partial V_i(x_i\!+\!1,\cdot )}{\partial k_i}\!-\!\frac{\partial V_i(s)}{\partial k_i}\right] dq_i(s) \!+\!\left[ \frac{\partial V_i(x_j\!+\!1,\cdot )}{\partial k_i}\!-\!\frac{\partial V_i(s)}{\partial k_i}\right] dq_j(s),\nonumber \\ \end{aligned}$$
(24)

and

$$\begin{aligned} d \frac{\partial V_i(s)}{\partial k_{j}}&= \biggl \{(\rho +\delta ) \frac{\partial V_i(s)}{\partial k_{j}} +\left[ \frac{\partial V_i(x_i+1,\cdot )}{\partial k_j}-\frac{\partial V_i(s)}{\partial k_{j}}\right] h_i(z_i(s))\nonumber \\&+\left[ \frac{\partial V_i(x_{j}+1,\cdot )}{\partial k_j}-\frac{\partial V_i(s)}{\partial k_{j}}\right] h_{j}(z_{j}(s))\nonumber \\&+\left[ V_i(x_j+1,\cdot )-V_i(s)\right] \alpha \gamma k_j^{\alpha -1}\biggr \} dt \nonumber \\&+\left[ \frac{\partial V_i(x_i\!+\!1,\cdot )}{\partial k_j}\!-\!\frac{\partial V_i(s)}{\partial k_j}\right] dq_i(s) \!+\!\left[ \frac{\partial V_i(x_j\!+\!1,\cdot )}{\partial k_j}\!-\!\frac{\partial V_i(s)}{\partial k_j}\right] dq_j(s),\nonumber \\ \end{aligned}$$
(25)

Finally, we replace the marginal values by marginal profits from the first order conditions (17). This yields

$$\begin{aligned}&d \frac{1}{u'(z_i)}- \lambda d \big [V_i(x_i+1,\cdot )-V_i(s)\big ]\nonumber \\&\quad = \biggl \{(\rho +\delta ) \frac{1-\lambda u'(z_i)\left[ V_i(x_i+1,\cdot )-V_i(s)\right] }{u'(z_i)}\nonumber \\&\quad \quad -\biggl [\frac{1-\lambda u'(z_i(x_i+1))\big [V_i(x_i+2,\cdot )-V_i(x_i+1)\big ]}{u'(z_i(x_i+1))}\nonumber \\&\quad \quad -\frac{1-\lambda u'(z_i) \big [V_i(x_i+1,\cdot )-V_i(s)\big ]}{u'(z_i)}\biggr ]h_i(z_i(s))\nonumber \\&\quad \quad -\biggl [\frac{1-\lambda u'(z_i(x_{-i}+1))\left[ V_i(s)-V_i(x_{-i}+1)\right] }{u'(z_i(x_{-i}+1))}\nonumber \\&\quad \quad -\frac{1-\lambda u'(z_i)\left[ V_i(x_i+1,\cdot )-V_i(s)\right] }{u'(z_i)}\biggr ]h_{j}(z_{j}(s))\nonumber \\&\quad \quad -\big [V_i(x_i+1,\cdot )-V_i(s)\big ]\alpha \gamma k_i^{\alpha -1}\biggl \}dt \nonumber \\&\quad \quad +\biggl [\frac{1-\lambda u'(z_i(x_i+1)) \big [V_i(x_i+2,\cdot )-V_i(x_i+1)\big ]}{u'(z_i(x_i+1))}\nonumber \\&\quad \quad -\frac{1-\lambda u'(z_i) \big [V_i(x_i+1,\cdot )-V_i(s)\big ]}{u'(z_i)}\biggr ]dq_i(s)\nonumber \\&\quad \quad +\biggl [\frac{1-\lambda u'(z_i(x_{-i}+1))\left[ V_i(s)-V_i(x_{-i}+1)\right] }{u'(z_i(x_{-i}+1))} \nonumber \\&\quad \quad -\frac{1-\lambda u'(z_i)\big [V_i(x_i+1,\cdot )-V_i(s)\big ]}{u'(z_i)}\biggr ]dq_j(s). \end{aligned}$$
(26)

This simplifies to

$$\begin{aligned}&d \frac{1}{u'(z_i)}- \lambda d \big [V_i(x_i+1,\cdot )-V_i(s)\big ]\nonumber \\&\quad = \biggl \{(\rho +\delta )\left( \frac{1}{u'(z_i)}-\lambda \big [V_i(x_i+1,\cdot )-V_i(s)\big ]\right) \nonumber \\&\quad \quad -\left[ \frac{1}{u'(z_i(x_i+1))} -\frac{1}{u'(z_i)}-\lambda \big [V_i(x_i+2,\cdot )+V_i(s)\big ]\right] h_i(z_i(s))\nonumber \\&\quad \quad -\left[ \frac{1}{u'(z_i(x_{-i}+1))}-\frac{1}{u'(z_i)}+\lambda \left[ V_i(x_{-i}+1)+V_i(x_i+1,\cdot )\right] \right] h_{j}(z_{j}(s))\nonumber \\&\quad \quad -\big [V_i(x_i+1,\cdot )-V_i(s)\big ]\alpha \gamma k_i^{\alpha -1}\biggl \}dt \nonumber \\&\quad \quad +\left[ \frac{1}{u'(z_i(x_i+1))}-\frac{1}{u'(z_i)} -\lambda \big [V_i(x_i+2,\cdot )+V_i(s)\big ]\right] dq_i(s)\nonumber \\&\quad \quad +\left[ \frac{1}{u'(z_i(x_{-i}+1))}-\frac{1}{u'(z_i)} +\lambda \big [V_i(x_{-i}+1)+V_i(x_i+1,\cdot )\big ] \right] dq_j(s).\qquad \quad \end{aligned}$$
(27)

On the other hand we can use the maximized Bellman equations together with the first order condition and get

$$\begin{aligned}&\rho V_i (s)=\pi _i(s)-z_i(s)+\frac{1-\lambda \big [V_i(x_i+1,\cdot )-V_i(s)\big ]}{u'(z_i)}\big (u(z_i(s))-\delta k_i\big )\nonumber \\&\quad +\frac{\partial V_i(s)}{\partial k_{j}}\left( u(z_{j}(s))-\delta k_{j}\right) \nonumber \\&\quad +\big [V_i(x_i+1,\cdot )-V_i(s)\big ]h_i(z_i(s))+\left[ V_i(x_{j}+1,\cdot )-V_i(s)\right] h_{j}(z_{j}(s)).\qquad \end{aligned}$$
(28)

We make use of the derivation above in the next subsections analyzing special cases of the general model.

B No learning-by-doing effects

Without knowledge acquisition, i.e., with \(\gamma =0\) and \(\Delta _i = x_i-x_j\) as the only state variable, the maximized Bellman equations (28), (19) respectively, simplify to:

$$\begin{aligned} \rho V_i(\Delta _i)&= \pi _i(\Delta _i)-z(u_i(\Delta _i))\nonumber \\&+\big [V_i(\Delta _i\!+\!1)\!-\!V_i(\Delta _i)\big ]h_i(u_i(\Delta _i))\!+\!\big [V_i(\Delta _i-1)\!-\!V_i(s)\big ]h_{j}(u_{j}(\Delta _i)).\nonumber \\ \end{aligned}$$
(29)

B.1 Proof of Lemma 1

Proof of Lemma 1 Using the envelope theorem and \(\lambda =1\) the first order condition for firm \(i\) yields

$$\begin{aligned} V_i(\Delta _i+1)-V_i(\Delta _i)=u_i(\Delta _i)^{\eta -1}. \end{aligned}$$
(30)

Note that each R&D effort is proportional to the incremental value that would result from innovating.Footnote 33 Inserting this in (29) gives

$$\begin{aligned} V_i(\Delta _i)=\tfrac{1}{\rho }\Bigl (\pi _i(\Delta _i)-z(u_i(\Delta _i))-z'(u_i(\Delta _i)) u_i(\Delta )-z'(u_i(\Delta _i-1)) u_{j}(\Delta _i)\Bigr )\nonumber \\ \end{aligned}$$
(31)

and

$$\begin{aligned} V_i(\Delta _i+1)&= \tfrac{1}{\rho }\Bigl (\pi _i(\Delta _i+1)-z(u_i(\Delta _i+1))-z'(u_i(\Delta _i+1)) u_i(\Delta _i+1)\nonumber \\&-z'(u_1(\Delta _i)) u_{j}(\Delta _i+1)\Bigr ). \end{aligned}$$
(32)

Using this in the first order condition yields

$$\begin{aligned}&u_i(\Delta _i)^{\eta -1} \left( \rho +u_i(\Delta _i) -u_{j}(\Delta _i+1)\right) \nonumber \\&\quad =\pi _i(\Delta _i+1)-z(u_i(\Delta _i+1))-\pi _i(\Delta _i)+z(u_i(\Delta _i))\nonumber \\&\qquad +u_i(\Delta _i+1)^{\eta } -(u_i(\Delta _i-1))^{\eta -1} u_{j}(\Delta _i). \end{aligned}$$
(33)

As the firms are ex ante symmetric, \(u_i(\Delta _i)=u_j(-\Delta _i)\) holds. This yields the reduced form R&D equations

$$\begin{aligned}&u_i(\Delta _i)^{\eta -1} \left( \rho +u_i(\Delta _i) -u_{i}(-\Delta _i-1)\right) \nonumber \\&= \pi _i(\Delta _i+1)-z(u_i(\Delta _i+1))-\pi _i(\Delta _i)+z(u_i(\Delta _i))\nonumber \\&\qquad +u_i(\Delta _i+1)^{\eta } -(u_i(\Delta _i-1))^{\eta -1} u_{i}(-\Delta _i). \end{aligned}$$
(34)

For the special case of \(\Delta _i=0\), this simplifies to

$$\begin{aligned}&u_i(0)^{\eta -1} \left( \rho +u_i(0) -u_{1}(-1)\right) \nonumber \\&\quad = 1-e^{-1}-\frac{1}{\eta }\left( u_i(1)^\eta -u_i(0)^{\eta }\right) +u_i(1)^{\eta } -(u_i(-1))^{\eta -1} u_{i}(0). \end{aligned}$$
(35)

Solving this for \(u(0)\) Footnote 34 and assuming \(\eta =2\) yields

$$\begin{aligned} u(0)=\sqrt{2-\frac{2}{e}+\rho ^2+u(1)^2}-\rho . \end{aligned}$$
(36)

For \(\Delta >0\) we have \(\pi (\Delta +1)-z(u(\Delta +1))-\pi (\Delta )+z(u(\Delta ))=e^{-\Delta -1}(1-e)+\frac{1}{\eta }\left( u(\Delta )^\eta -u(\Delta +1)^{\eta }\right) \) while when \(\Delta <0\) the increase in profit when moving one step ahead is zero, i.e., \(\pi (u(\Delta +1))-\pi (u(\Delta ))=0\). Using this we get the relations of optimal R&D stated in Lemma 1. \(\square \)

B.2 Proof of Proposition 1

Proof of Proposition 1. We restrict attention to non-fluctuating patterns. That means for all \(\Delta >0:\) if \(u(\Delta )>u(\Delta +1)\) then \(u(\Delta +1)>u(\Delta +2)\) and vice versa. The same has to be true for the follower, i.e., for all \(-\Delta <0:\) if \(u(-\Delta )>u(-\Delta -1)\) then \(u(-\Delta -1)>u(-\Delta -2)\) and vice versa. Hence, we are not able to rigorously prove the uniqueness of the equilibrium we derive. However, economic intuitions for fluctuating effort do not exist when time preference rate is set to zero.

From Eq. (5) in the first part of Lemma 1, obviously \(u(0)>u(1)\) results.

To see how the effort of the industry’s follower reacts on technological jumps we firstly analyze Eq. (7) given in the third part of Lemma 1.

This equation gives for \(-\Delta <0\)

$$\begin{aligned} \left( u(-\Delta )-\tfrac{1}{2}u(\Delta -1)\right) ^2=&\left( u(-\Delta +1)-\tfrac{1}{2}u(\Delta -1)\right) ^2\\&+u(-\Delta +1)u(\Delta -1)-u(-\Delta -1)u(\Delta ) \end{aligned}$$

and hence

$$\begin{aligned}&\mathrm {sign}\left\{ (u(-\Delta )-\tfrac{1}{2}u(\Delta -1))^2 - (u(-\Delta +1)-\tfrac{1}{2}u(\Delta -1))^2\right\} \\&\quad =\mathrm {sign}\big \{u(-\Delta +1)u(\Delta -1)- u(-\Delta -1)u(\Delta )\big \}. \end{aligned}$$

Here, we have to distinct different cases. Let us first assume that \(u(-\Delta )>\tfrac{1}{2}u(\Delta )\) \(\forall \Delta >0\) (implying \(u(-\Delta +1)>\tfrac{1}{2}u(\Delta -1)\) \(\forall \Delta >0\)). Then, we get

$$\begin{aligned}&\mathrm {sign}\big \{u(-\Delta ) - u(-\Delta +1)\big \}\nonumber \\&\quad =\mathrm {sign}\big \{u(-\Delta +1)u(\Delta -1)- u(-\Delta -1)u(\Delta )\big \}. \end{aligned}$$
(37)

This means efforts for the follower are decreasing with the gap when \(u(-\Delta +1)u(\Delta -1)< u(-\Delta -1)u(\Delta )\) holds. The opposite is true for \(u(-\Delta )<\tfrac{1}{2}u(\Delta )\) \(\forall \Delta >0\), i.e., follower’s effort increases if \(u(-\Delta +1)u(\Delta -1)> u(-\Delta -1)u(\Delta )\).

For \(\Delta =1\), these relations also compare the follower’s effort with the effort of a neck-and-neck firm. Analyzing the situation for \(-\Delta =-1\) yields

$$\begin{aligned} \mathrm {sign}\big \{u(-1) - u(0)\} =\mathrm {sign}\{u(0)^2- u(-2)u(1)\big \}. \end{aligned}$$

We already know that investment in neck-and-neck state is higher than effort of a firm being one step ahead, i.e., \(u(0)>u(1)\). This indicates that \(u(0)^2>u(-2)u(1)\) might hold, implicating \(u(-1) > u(0)\). This is true as long as \(u(-2)\) is not too large, i.e., \(u(-1)>u(0)>u(1)\) iff \(u(-2)<\tfrac{u(0)^2}{u(1)}>u(0)\). If \(u(-2)\) is large enough to outweigh the difference between \(u(0)\) and \(u(1)\) we have \(u(-1) > u(0)\). In that case the relation \(u(-2)>u(0)>u(-1)\) holds. That means the optimal pattern shows some kind of fluctuation.

We can illustrate the characteristics of the general Eq. (37) by means of the example of \(\Delta =2\). Equation (37) yields

$$\begin{aligned} \mathrm {sign}\big \{u(-2) - u(-1)\big \} =\mathrm {sign}\big \{u(-1)u(1)- u(-3)u(2)\big \}. \end{aligned}$$

As we are not looking for fluctuating patterns, we either have \(u(-3)>u(-2)> u(-1)\) or \(u(-3)<u(-2)< u(-1)\). In the first case, \(u(2)\) would have to be sufficiently small to ensure \(u(-1)u(1)>u(-3)u(2)\). In the case of \(u(-3)<u(-2)< u(-1)\), \(u(2)\) needs to be sufficiently large. Obviously, in either case do R&D efforts for leader and follower go into opposite directions when the gap increases. This clearly also holds true for the general case of \(\Delta >2\) and Eq. (37). More precisely, \(u(-\Delta ) \gtreqless u(-\Delta +1)\) if \(\frac{u(-\Delta +1)}{u(-\Delta -1)}\gtreqless \frac{u(\Delta )}{u(\Delta -1)}\). If now \(u(\Delta )>u(\Delta -1)\), \(u(-\Delta +1)> u(-\Delta )>u(-\Delta -1)\) has to hold and vice versa.

Keeping in mind that leader’s and follower’s effort move into opposite directions when the gap increases, let’s now analyze the leader’s optimal effort given by Eq. (6) in the second part of Lemma 1. We can directly see that for \(\Delta >0\)

$$\begin{aligned}&(u (\Delta )-u(-\Delta -1))^2\\&\quad =\left( u(\Delta +1)-u(-\Delta -1)\right) ^2\\&\qquad +2\left( u(-\Delta -1)u(\Delta +1)-u(\Delta -1)u(-\Delta )+e^{-\Delta }\left( 1-\tfrac{1}{e}\right) \right) , \end{aligned}$$

and hence

$$\begin{aligned}&\mathrm {sign}\big \{(u(\Delta )-u(-\Delta -1))^2 - (u(\Delta +1)-u(-\Delta -1))^2\big \}\\&\quad =\mathrm {sign}\big \{u(-\Delta -1)u(\Delta +1)- u(\Delta -1)u(-\Delta )+ e^{-\Delta }\left( 1-\tfrac{1}{e}\right) \big \} \end{aligned}$$

holds. Again, we have to distinct different cases. Let us first assume \(u(\Delta )<u(-\Delta -1)\) and \(u(\Delta )<u(-\Delta )\) \(\forall \Delta >0\). In that case the assumption \(u(-\Delta )>\tfrac{1}{2}u(\Delta )\) \(\forall \Delta >0\) holds as well. Besides, as we know leader’s and follower’s effort move into opposite directions, the assumptions implicate \(u(\Delta ')<u(-\Delta )\) \(\forall \Delta , \Delta '>0\).

Then, we get

$$\begin{aligned}&\mathrm {sign} \big \{u(\Delta +1)-u(\Delta )\big \} \nonumber \\&\quad =\mathrm {sign}\big \{u(-\Delta -1)u(\Delta +1)- u(\Delta -1)u(-\Delta )+ e^{-\Delta }\left( 1-\tfrac{1}{e}\right) \big \}. \end{aligned}$$
(38)

That means, we see increasing efforts for leaders when \(u(-\Delta -1)u(\Delta +1)> u(-\Delta )u(\Delta -1)- e^{-\Delta }\left( 1-\frac{1}{e}\right) \).

In the case of \(u(\Delta ')<u(-\Delta )\) \(\forall \Delta , \Delta '>0\) the leader’s effort can only be increasing if the follower’s effort is decreasing and furthermore if \(u(-\Delta -1)u(\Delta +1)> u(-\Delta )u(\Delta -1)- e^{-\Delta }\left( 1-\frac{1}{e}\right) \) holds. As in this case \(u(\Delta +1)> u(\Delta -1)\) holds, we have found an equilibrium where the leader’s increase in effort with an increase in gap is not too large. Besides, it is clear that beyond this \(u(-1)>u(0)>u(1)\) holds, since effort is decreasing for the follower and therefore \(u(-2)>u(0)>u(-1)\) cannot hold. The resulting pattern is that summarized in terms of investment by Proposition 1. Hence, we have shown that the described optimal behavior is indeed an equilibrium. \(\square \)

To show that no other equilibria exist is a very comprehensive task and needs quantifying analysis. Unfortunately, we are not able to analytically show the uniqueness of the equilibrium.

C The effect of learning-by-doing

C.1 Proof of Proposition 3

Proof of Proposition 3 With the simplifying assumption that investment in R&D does not influence a firm’s probability of success immediately, i.e., with \(\lambda =0\), the first order conditions read

$$\begin{aligned} -1+\frac{\partial V_i(s)}{\partial k_i}u'(z_i(s))=0, \end{aligned}$$
(39)

yielding

$$\begin{aligned} d\frac{\partial V_i(s)}{\partial k_i}=d\frac{ 1}{ u'(z_i(s))} . \end{aligned}$$
(40)

Therefore, we get the optimal rule describing the evolution of marginal profits:

$$\begin{aligned} -d\frac{ -1}{ u'(z_i(s))}&= \biggl \{-(\rho +\delta ) \frac{-1}{u'(z_i(s))} +\left[ \frac{-1}{u'(z_i(x_i+1,\cdot ))} -\frac{-1}{u'(z_i(s))}\right] \gamma k_i^{\alpha }\nonumber \\&\quad +\left[ \frac{-1)}{u'(z_i(x_j+1,\cdot ))}-\frac{-1}{u'(z_i(s))}\right] \gamma k_{j}^\alpha \nonumber \\&\quad -\bigg [V_i(x_i+1,\cdot )-V_i(s)\bigg ]\alpha \gamma k_i^{\alpha -1}\biggl \}dt \nonumber \\&\quad +\left[ \frac{-1}{u'(z_i(s))}-\frac{-1}{u'(z_i(x_i+1,\cdot ))}\right] dq_i(s)\nonumber \\&\quad +\left[ \frac{-1}{u'(z_i(s))}-\frac{-1}{u'(z_i(x_j+1,\cdot ))}\right] dq_j(s). \end{aligned}$$
(41)

The rule shows how marginal profit changes in a deterministic and stochastic way. While there is a one-to-one mapping from marginal profit to investment which allows some inferences about investment from (41), it would be more useful to have a rule for optimal investment itself. With firms’ instantaneous profits (3) and rate of knowledge acquisition \(u(z_i)=(\eta z_i)^{\frac{1}{\eta }}\) we get

$$\begin{aligned} d\frac{-1}{ u'(z_i(s))}\!=\!-\frac{\partial \frac{1}{u'(z_i(s))}}{\partial z_i(s)}dz_i(s)\!=\!\frac{u''(z_i(s))}{u'(z_i(s))^2}dz_i(s)\!=\!(1\!-\!\eta )(\eta z_i(s))^{-\frac{1}{\eta }}dz_i(s).\nonumber \\ \end{aligned}$$
(42)

Due to the modeling approach only the technological gap and not the technological levels as such matters for firms’ values, i.e., the effect of the competitor moving one step forward is the same as moving one step backwards. Thus, using (42) we can write

$$\begin{aligned} dz_i(s)&= \frac{(\eta z_i(s))^{\frac{1}{\eta }}}{\eta -1}\Biggl [ \biggl \{\frac{(\rho +\delta )}{u'(z_i(s))} +\left[ \frac{1}{u'(z_i(s))}-\frac{1}{u'(z_i(x_i+1,\cdot ))}\right] \gamma k_i^{\alpha }\nonumber \\&+\left[ \frac{1}{u'(z_i(s))}\!-\!\frac{1}{u'(z_i(x_i\!-\!1,\cdot ))}\right] \gamma k_{j}^\alpha \!-\!\big [V_i(x_i\!+\!1,\cdot )\!-\!V_i(s)\big ]\alpha \gamma k_i^{\alpha -1}\biggl \}dt \nonumber \\&+\left[ \frac{1}{u'(z_i(x_i+1,\cdot ))}-\frac{1}{u'(z_i(s))}\right] dq_i(s)\nonumber \\&+\left[ \frac{1}{u'(z_i(x_i-1,\cdot ))}-\frac{1}{u'(z_i(s))}\right] dq_j(s)\Biggl ]. \end{aligned}$$
(43)

These rules describe the evolution of investment under optimal behavior for the firms. Growth of investment depends on the right-hand side in a deterministic way on the typical sum of the depreciation and time preference rate per marginal rate of knowledge acquisition plus the “\(k\)-terms” which capture the impact of uncertainty. To understand the meaning of these terms we analyze whether investment jumps up or down, following a jump of the own or competitor’s technology. Since \(\eta >1\) the term \(\frac{1}{u'(z_1(s))}-\frac{1}{u'(z_1(x_1+1,\cdot ))}=\eta ^{\frac{\eta -1}{\eta }}\left( z_1(s)^{\frac{\eta -1}{\eta }}-z_1(x_1+1,\cdot )^{\frac{\eta -1}{\eta }} \right) \) is negative if \(z_1(s)<z_1(x_1+1,\cdot )\). If this is the case, investment increases slower (or decreases even faster) if the probability of a jump of the own technology due to a higher knowledge stock is high. On the other hand investment increases faster (or decreases slower) if the probability of a jump of the competitor’s technology due to his higher knowledge stock is high.

The \(dq_x{_i}\)-terms give discrete changes in the case of a jump in \(x_i\). When \(x_i\) jumps and \(dq_{x_i(s)}=1\) (\(dq_{x_j(s)}=0\), i.e., there is no contemporaneous jump in \(x_j\)) and \(dt=0\) for this small instant of the jump, Eq. (43) says that \(dz_i(s)\) on the left hand side is given by

$$\begin{aligned} \frac{\eta z_i(s) }{\eta -1}\left( z_i(s)^{\frac{1-\eta }{\eta }}z_i(x_i+1,\cdot )^{\frac{\eta -1}{\eta }}-1\right) \end{aligned}$$
(44)

on the right hand side. This is positive as long as \(z_i(s)<z_i(x_i+1,\cdot )\) which is consistent with the definition of \(dz_i(s)\) given by \(z_i(x_i+1,\cdot )-z_i(s)\). Solving this for \(z_i(s)\) interestingly yields \(z_i(s)=z_i(x_i+1,\cdot )\). Hence, optimal investment does not immediately react to a jump in the industry’s state. This is the result stated in Proposition 3.\(\square \)

C.2 Proof of Lemma 2

Proof of Lemma 2. Using the derived fact, from (43) we can determine the evolution of optimal investment:

$$\begin{aligned} \frac{dz_i(s)}{dt}=z_i(s)\frac{\eta }{\eta -1}\Bigl (\rho +\delta -(\eta z_i(s))^{\frac{1-\eta }{\eta }}\underbrace{\left[ V_i(x_i+1,\cdot )-V_i(s)\right] }_{\equiv \Phi _i (s) }\alpha \gamma k_i^{\alpha -1}\Bigr ). \end{aligned}$$
(45)

We cannot determine the value of \(\Phi _i (s)\), but we know it is always positive and from the first order condition we get \(\frac{\partial \Phi _i (s)}{\partial k_i}=0\). Thus, \(\Phi (\cdot )\) is only a function of the technological gap \(\Delta \) and the same function for both firms, i.e \(\Phi _i (\Delta )=\Phi _j(-\Delta )\). Hence, we can write \(\Phi (\Delta _i)\equiv \Phi _i (\Delta )\). With this result it can directly be seen that the more knowledge a firm has acquired the smaller is the growth rate of optimal investment.

Furthermore, from (28) and with \(\lambda =0\) we can analyze the shape of \(\Phi (\Delta )\). As the value function inherits its shape from the profit function, the value function will be bounded from below and above and will converge to these bound for \(\Delta \rightarrow - \infty \) and \(\Delta \rightarrow \infty \) respectively. Hence, \(\frac{\partial \left( V_i(\Delta +1)-V_i(\Delta )\right) }{\partial \Delta }\) is negative for high values of \(\Delta \) and positive for small (negative) values.

As the slope of the value function measures a leader’s incentive to innovate, this slope is maximal around the neck and neck point since neck-and-neck firms perform R&D at a higher intensity than industry leaders.

The maximized Bellman equations (28) hold for all optimal efforts, especially at steady state as well, i.e., when \(u(z_j)-\delta k_j=0\). In this case, for \(\Delta <0\) we have:

$$\begin{aligned} \rho V_i (s)=-z_i(s)+\big [\Phi (\Delta )\big ]\gamma k_i^{\alpha }-\big [\Phi (\Delta -1)\big ]\gamma k_j^{\alpha }. \end{aligned}$$

The right hand side can only be positive for positive \(\Phi (\Delta )\) if \(\Phi (\Delta )>\Phi (\Delta -1)\) and therefore we have \(\frac{\partial \Phi (\Delta )}{\partial \Delta }>0\) for \(\Delta <0\).

The derivative of the maximized Bellman equation (28) with respect to \(\Delta \) for \(\Delta <0\) using the envelope theorem gives

$$\begin{aligned} \rho \frac{\partial V_i (\Delta , k_i, k_j)}{\partial \Delta }&= \frac{\partial ^2 V_i(\Delta ,\cdot )}{\partial \Delta \partial k_{j}}\left( u(z_j(s))-\delta k_j\right) \nonumber \\&+\left[ \frac{\partial V_i(\Delta +1,\cdot )}{\partial \Delta }-\frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }\right] \gamma k_i^{\alpha } \nonumber \\&+\left[ \frac{\partial V_i(\Delta -1,\cdot )}{\partial \Delta }-\frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }\right] \gamma k_j^{\alpha }. \end{aligned}$$
(46)

For \(\Delta >0\) the derivative yields

$$\begin{aligned} \rho \frac{\partial V_i (\Delta , k_i, k_j)}{\partial \Delta }&= e^{-\Delta }+\frac{\partial ^2 V_i(\Delta ,\cdot )}{\partial \Delta \partial k_{j}}\left( u(z_j(s))-\delta k_j\right) \nonumber \\&+\left[ \frac{\partial V_i(\Delta +1,\cdot )}{\partial \Delta }-\frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }\right] \gamma k_i^{\alpha } \nonumber \\&+\left[ \frac{\partial V_i(\Delta -1,\cdot )}{\partial \Delta }-\frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }\right] \gamma k_j^{\alpha }. \end{aligned}$$
(47)

Here, we can see that the only critical value of \(\Delta \), i.e., a value where signs could possibly change, is indeed at \(\Delta =0\).

We can now determine the derivative of the value effect of a technological step ahead for \(\Delta >0\):

$$\begin{aligned} \rho&\frac{\partial \left( V_i (\Delta +1, k_i, k_j)-V_i (\Delta , k_i, k_j)\right) }{\partial \Delta }\nonumber \\&=e^{-\Delta -1}(1-e)+\left( \frac{\partial ^2 V_i(\Delta +1,\cdot )}{\partial \Delta \partial k_{-i}}-\frac{\partial ^2 V_i(\Delta ,\cdot )}{\partial \Delta \partial k_{-i}}\right) (u(z_j)-\delta k_j) \nonumber \\&\quad +\left[ \frac{\partial V_i(\Delta +2,\cdot )}{\partial \Delta }-2\frac{\partial V_i(\Delta +1,\cdot )}{\partial \Delta }+\frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }\right] \gamma k_i^{\alpha } \nonumber \\&\quad +\left[ \frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }-\frac{\partial V_i(\Delta +1,\cdot )}{\partial \Delta }-\frac{\partial V_i(\Delta -1,\cdot )}{\partial \Delta }+\frac{\partial V_i(\Delta ,\cdot )}{\partial \Delta }\right] \gamma k_j^{\alpha }.\quad \end{aligned}$$
(48)

We know \(V(\cdot )\) is approaching the upper bound for \(\Delta \rightarrow \infty \). Hence, for sufficiently large values of \(\Delta \), the second derivative of the value function and thus the first derivative of \(\Phi (\Delta )\) will be negative. But even more, we see that the right hand side of equation (48) is negative for all \(\Delta >0\) in steady state with sufficiently low knowledge stocks as \(e^{-\Delta -1}(1-e)<0\). Hence, the value function’s inflection point has to be at \(\Delta =0\) and we have \(\frac{\partial \Phi (\Delta )}{\partial \Delta }<0\) for \(\Delta >0\).\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Steinmetz, A. Competition, innovation, and the effect of R&D knowledge. J Econ 115, 199–230 (2015). https://doi.org/10.1007/s00712-014-0415-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00712-014-0415-3

Keywords

JEL classification

Navigation