Skip to main content

Advertisement

Log in

Endogenous structural change, aggregate balanced growth, and optimality

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

Abstract

Developed countries have experienced labor employment shifts from the manufacturing sector to the services sector, which is a process termed structural change. We construct a two-sector endogenous growth model of structural change in which knowledge spillover sustains growth. Our model is consistent with the empirically observed pattern of structural change as well as aggregate balanced growth. In both centralized and decentralized economies, labor employment shifts from a high productivity growth sector to a low productivity growth sector. Because of knowledge spillover, excessive labor is allocated to productivity improvements in the high productivity growth sector. Hence, the decentralized economy is characterized by rapid structural change compared to the optimal allocation.

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

Similar content being viewed by others

Notes

  1. For example, see Kuznets (1971), Maddison (1980), Baumol et al. (1985), and Jorgenson and Timmer (2011). Some authors note that services are heterogeneous and that the service sector contains some subsectors that show the economy’s highest productivity growth and some that show the lowest productivity growth (Baumol et al. 1985; Jorgenson and Timmer 2011). However, for simplicity, we do not consider such heterogeneity of services.

  2. For example, see Baumol (1967), Echevarria (1997), Laitner (2000), Kongsamut et al. (2001), Ngai and Pissarides (2007), Acemoglu and Guerrieri (2008), Foellmi and Zweimüller (2008), Buera and Kaboski (2009, 2012), İşcan (2010), Guilló et al. (2011), and Hori et al. (2015).

  3. Some authors introduce both approaches into unified frameworks. İşcan (2010) finds that although both approaches are important, productivity growth differentials played a relatively important role in structural change in the second half of the twentieth century in the USA. Boppart (2014) examines postwar US structural change and finds that both approaches are equally important. Guilló et al. (2011) examine structural change from agricultural to non-agricultural sectors considering both approaches. The authors conclude that the second approach is important. Echevarria (1997) and Buera and Kaboski (2009) also introduce both approaches into unified frameworks.

  4. See Theorem 5 of Acemoglu and Guerrieri (2006). Equation (74) shows the employment shift from Sector 2 to Sector 1. Equations (71), (73), and (75) imply that Sector 1 experiences higher productivity growth than Sector 2.

  5. The first generation of the innovation-based endogenous growth model, such as Romer (1990) and Grossman and Helpman (1991), has been empirically criticized because it exhibits “scale effects”: The growth rate of the economy is proportional to resources employed in the R&D sector (see Jones 1995). Broadly speaking, there are two types of endogenous growth models without scale effects. The first type is the semi-endogenous growth model, where the long-run growth rate is proportional to the population growth rate. The second is characterized by the presence of both horizontal and vertical innovations. Authors such as Lainez and Peretto (2006) provide evidence that supports the second type.

  6. The second-order condition for firms’ profit maximization, \(\partial ^2 \Pi _{i,j}(t-1)/\partial b_{i,j}(t)^2 <0\), where \( \Pi _{i,j}(t-1)\) is defined later, requires the elasticity \(za_i^\prime (z)/a_i(z)\) to increase with \(z= b_{i,j}(t) /\overline{b}_{i}(t-1) \). The specification of Young (1998) satisfies this property. However, a constant elasticity function, \(a(z)=z^d\) (\(d>1)\), fails to satisfy this property.

  7. Ngai and Samaniego (2011) study sources of productivity differential for US industries. They find that the main determinant of difference in productivity growth among production sectors is technology of knowledge production.

  8. Industries such as research and business services have large shares of R&D employment. Under the NACE industry classification scheme REV. 1, these are classified as 73 and 74. Both are included in Market Services in Table 1.

  9. Even if \( \varepsilon >1 \), (10) and (25) hold. A comparison between (12) and (13) shows that if \( \varepsilon >1 \) (and \( \varepsilon \ne 1+1/\sigma \)), \( A_2(t)/A_1(t) \) and \( N(t) ({=}N_2(t)/N_1(t)) \) move in the same direction. Thus, employment shifts from the low productivity growth sector to the high productivity growth sector. When \( 1<\varepsilon <1+1/\sigma \), the productivity growth in Sector 2 is lower than that in Sector 1, while the productivity growth in Sector 2 is higher than that in Sector 1 when \( \varepsilon >1+1/\sigma \).

  10. (10) and (11) show that without firm-level productivity growth differential, \( A_2(t)/A_1(t) \) moves in the same direction as \( n_2(t)/n_1(t) \) and \( E_2(t)/E_1(t) \).

  11. We can show that if the instantaneous utility takes a CRRA form, \( D(t)^{1-\zeta }/(1-\zeta ) \) (\( \zeta >0 \) and \( \zeta \ne 1\)), structural change does not satisfy aggregate balanced growth in our model.

  12. See footnote 4 in Introduction of this paper.

  13. To obtain a constant growth rate, Boppart (2014) and Herrendorf and Valentinyi (2015) introduce an investment good sector with an AK technology, of which product does not directly enter utility function.

  14. Monopolistic competition also creates inefficient allocation. A small \( \alpha \) indicates a strong market power. In both centralized and decentralized economies, the rates of structural change are independent of \( \alpha \). Then, monopolistic competition does not generate disparity in the rates of structural change between the centralized and decentralized economies.

  15. See footnote 12 in Young (1998) for more details.

  16. Proposition 2 does not prove the uniqueness of the solution for the central planner’s problem. However, we reasonably guess that the solution is unique because of the concavity of utility function and the convexity of \( a_i(\cdot ) \).

  17. This result holds even under a Dixit–Stiglitz-type consumption index because our specification of composite good, (2), includes the Dixit–Stiglitz type.

  18. Besides \( \tau ^I_i(t) \) and \( \tau ^E_i(t) \), production subsides are also needed to realize the optimal allocation because of monopolistic competition. Let us denote the subsidy rate as \( \tau ^P_i(t) \). The operating profit of firm j in Sector i is given by \( \pi _{i,j}(t) = [p_{i,j}(t) -(1-\tau ^P_i(t) )w(t)/b_{i,j}(t) ] x_{i,j}(t) \). We can show that the optimal subsidy rate satisfies \( 1-\tau ^P_i(t)=\sigma \alpha /(1-\alpha ) \). Because the presence of \( \tau ^P_i(t) \) does not affect the characteristics of the optimal \( \tau ^I_i(t) \) and \( \tau ^E_i(t) \), we do not consider \( \tau ^P_i(t) \) for simplicity.

  19. Even under (22), (38) and (41) hold. The other two first-order conditions are modified as

    $$\begin{aligned}&\lambda _P(t) \frac{n_i^*(t)x_i^*(t)}{b_i^*(t)} =v(t)^\frac{1-\varepsilon }{\varepsilon } \gamma _i \left( n_i^*(t)^{\sigma +1} x_i^*(t) -s_i\right) ^\frac{-1}{\varepsilon } n_i^*(t)^{\sigma +1} x_i^*(t) , \\&v(t)^\frac{1-\varepsilon }{\varepsilon } (\sigma +1) \gamma _i \left( n_i^*(t)^{\sigma +1} x_i^*(t) -s_i \right) ^\frac{-1}{\varepsilon }n_i^*(t)^{\sigma +1} x_i^*(t)\\&\quad = n_i^*(t) \left[ \lambda _P(t) \frac{x_i^*(t)}{b_i^*(t)} + \frac{\lambda _I(t-1)}{\beta } a_i\left( \frac{b_i^*(t)}{b_i^*(t-1)} \right) \right] . \end{aligned}$$

    From these equations and \( \lambda (t) \equiv \lambda _P(t) = \lambda _I(t) \), we obtain \( \sigma \lambda (t) \frac{n_i^*(t)x_i^*(t)}{b_i^*(t)} = \frac{\lambda (t-1)}{\beta } a_i(t)n_i^*(t) \). From the last equation, (41), and \( a_i(g)= f_i e^{\xi _i g} \), we derive (23).

References

  • Acemoglu, D., Guerrieri, V.: Capital deepening and nonbalanced economic growth. NBER Working Paper 12475 (2006)

  • Acemoglu, D., Guerrieri, V.: Capital deepening and nonbalanced economic growth. J. Polit. Econ. 116, 467–498 (2008)

    Article  Google Scholar 

  • Baumol, W.J.: Macroeconomics of unbalanced growth: the anatomy of urban crisis. Am. Econ. Rev. 57, 415–426 (1967)

    Google Scholar 

  • Baumol, W.J., Blackman, S.N.B., Wolff, E.N.: Unbalanced growth revisited: asymptotic stagnancy and new evidence. Am. Econ. Rev. 75, 806–817 (1985)

    Google Scholar 

  • Benassy, J.-P.: Taste for variety and optimum production patterns in monopolistic competition. Econ. Lett. 52, 41–47 (1996)

    Article  Google Scholar 

  • Boppart, T.: Structural change and the Kaldor facts in a growth model with relative price effects and non-Gorman preferences. Working Paper Series, Department of Economics, University of Zurich (2011)

  • Boppart, T.: Structural change and the Kaldor facts in a growth model with relative price effects and non-Gorman preferences. Econometrica 82, 2167–2196 (2014)

    Article  Google Scholar 

  • Buera, F.J., Kaboski, J.P.: Can traditional theories of structural change fit the data? J. Eur. Econ. Assoc. 7, 469–477 (2009)

    Article  Google Scholar 

  • Buera, F.J., Kaboski, J.P.: Scale and the origins of structural change. J. Econ. Theory 147, 684–712 (2012)

    Article  Google Scholar 

  • Dinopoulos, E., Thompson, P.: Schumpeterian growth without scale effects. J. Econ. Growth 3, 313–335 (1998)

    Article  Google Scholar 

  • Echevarria, C.: Changes in sectoral composition associated with economic growth. Int. Econ. Rev. 38, 431–452 (1997)

    Article  Google Scholar 

  • Foellmi, R., Zweimüller, J.: Structural change, Engel’s consumption cycles and Kaldor’s facts of economic growth. J. Monet. Econ. 55, 1317–1328 (2008)

    Article  Google Scholar 

  • Grossman, G.M., Helpman, E.: Innovation and Growth in the Global Economy. MIT Press, Cambridge (1991)

    Google Scholar 

  • Guilló, M.D., Papageorgiou, C., Perez-Sebastian, F.: A unified theory of structural change. J. Econ. Dyn. Control 35, 1393–1404 (2011)

    Article  Google Scholar 

  • Herrendorf, B., Rogerson, R., Valentinyi, Á.: Two perspectives on preferences and structural transformation. Am. Econ. Rev. 103, 2752–2789 (2013)

    Article  Google Scholar 

  • Herrendorf, B., Valentinyi, Á.: Endogenous sector-biased technological change and industrial policy. Working Paper (2015)

  • Hori, T., Ikefuji, M., Mino, K.: Structural change and conformism. Int. Econ. Rev. 56, 939–961 (2015)

    Article  Google Scholar 

  • Howitt, P.: Steady endogenous growth with population and R&D inputs growing. J. Polit. Econ. 107, 715–730 (1999)

    Article  Google Scholar 

  • İşcan, T.B.: How much can Engel’s law and Baumol’s disease explain the rise of the service employment in the United States? BE. J. Macroecon. 10(1), Article 26 (2010)

  • Jones, C.I.: R&D-based models of economic growth. J. Polit. Econ. 103, 759–794 (1995)

    Article  Google Scholar 

  • Jorgenson, D.W., Timmer, M.P.: Structural change in advanced nations: a new set of stylised facts. Scand. J. Econ. 113, 1–29 (2011)

    Article  Google Scholar 

  • Kongsamut, P., Rebelo, S., Xie, D.: Beyond balanced growth. Rev. Econ. Stud. 68, 869–882 (2001)

    Article  Google Scholar 

  • Kuznets, S.: Economic Growth of Nations: Total Output and Production Structure. Harvard University Press, Cambridge (1971)

    Book  Google Scholar 

  • Lainez, C.A., Peretto, P.F.: Scale effects in endogenous growth theory: an error of aggregation not specification. J. Econ. Growth 11, 263–288 (2006)

    Article  Google Scholar 

  • Laitner, J.: Structural change and economic growth. Rev. Econ. Stud. 67, 545–561 (2000)

    Article  Google Scholar 

  • Maddison, A.: Economic growth and structural change in advanced countries. In: Levenson, I., Wheeler, J. (eds.) Western Economies in Transition: Structural Change and Adjustment Policies in Industrial Countries, pp. 41–60. Westview Press, Boulder (1980)

    Google Scholar 

  • Matsuyama, K.: Structural change in an interdependent world: a global view of manufacturing decline. J. Eur. Econ. Assoc. 7, 478–486 (2009)

    Article  Google Scholar 

  • Ngai, L.R., Pissarides, C.A.: Structural change in a multisector model of growth. Am. Econ. Rev. 97, 429–445 (2007)

    Article  Google Scholar 

  • Ngai, L.R., Samaniego, R.M.: Accounting for research and productivity growth across industries. Rev. Econ. Dyn. 14, 475–495 (2011)

    Article  Google Scholar 

  • Peretto, P.F.: Technological change and population growth. J. Econ. Growth 3, 283–311 (1998)

    Article  Google Scholar 

  • Romer, P.: Endogenous Technological Change. J. Polit. Econ. 70, 65–94 (1990)

    Google Scholar 

  • Uy, T., Yi, K.-M., Zhang, J.: Structural change in an open economy. J. Monet. Econ 60, 667–682 (2013)

    Article  Google Scholar 

  • Young, A.: Growth without scale effects. J. Polit. Econ. 106, 41–63 (1998)

    Article  Google Scholar 

Download references

Acknowledgments

The authors are grateful to two anonymous referees and the editor of this journal, Nicholas Yannelis, for their constructive comments. We also thank Shuhei Aoki, Yosuke Furukawa, Koichi Futagami, Hirokazu Ishise, Tatsuro Iwaisako, Munechika Katayama, Yoko Konishi, Kazuo Mino, Akira Momota, Mototsugu Shintani, Shuhei Takahashi, Takayuki Tsuruga, Katsunori Yamada, and seminar participants at Aoyamagakuin University, Kyoto University, and Osaka University. Hori’s research was financially supported by JSPS KAKENHI Grant No. 15K17025.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Takeo Hori.

Appendices

Appendices

1.1 Appendix 1: Optimization problem of the representative household

We solve the optimization problem of the representative household in three steps. We first maximize (2) subject to (3), which yields the following demand function for brand j of good i:

$$\begin{aligned} x_{i,j}(t) = \frac{E_i(t) {p_{i,j}(t)}^{-\frac{1}{1-\alpha }} }{ \int ^{n_i(t)}_0 { p_{i,j^\prime }(t) }^{-\frac{\alpha }{1-\alpha }} \mathrm{d}j^\prime }. \end{aligned}$$
(24)

Substituting (24) into (2) yields \(D_i(t) = E_i(t)/P_i(t)\), where \(P_i(t) \equiv {n_i(t)}^{\frac{1-\alpha }{\alpha }-\sigma } \left[ \int ^{n_i(t)}_0 { p_{i,j}(t) }^{-\frac{\alpha }{1-\alpha } } \mathrm{d}j \right] ^{-\frac{1-\alpha }{\alpha } }\) is the price index of the composite good i. The next step maximizes (1) subject to \(E(t) = E_1(t) + E_2(t)\) and \(D_i(t) = E_i(t)/P_i(t)\) to obtain

$$\begin{aligned} \frac{D_2(t)}{D_1(t)} = \left( \frac{1-\gamma }{\gamma } \right) ^\varepsilon \left( \frac{P_1(t)}{P_2(t)} \right) ^\varepsilon , \quad \text{ or } \quad \frac{E_2(t)}{E_1(t)} = \left( \frac{1-\gamma }{\gamma } \right) ^\varepsilon \left( \frac{P_2(t)}{P_1(t)} \right) ^{1-\varepsilon }. \end{aligned}$$
(25)

Using \(E(t)=P_1(t)D_1(t)+P_2(t)D_2(t)\) and the first equation of (25), we can rewrite (1) as \(D(t) = E(t)/P(t)\), where \(P(t) \equiv [ \gamma ^\varepsilon {P_1(t)}^{1-\varepsilon } + (1-\gamma )^\varepsilon {P_2(t)}^{1-\varepsilon } ]^\frac{1}{1-\varepsilon }\) represents the price index for all consumption goods. We normalize \( P_t=1 \). We finally maximize U(t) subject to (4) and \(D(t) = E(t)/P(t)\), which yields

$$\begin{aligned} \frac{D(t+1)}{D(t)}=\frac{E(t+1)}{E(t)} = \beta ( 1+ r(t) ). \end{aligned}$$
(26)

The transversality condition is \( \lim _{t \rightarrow +\infty } \beta ^t W(t)/E(t)=0 \).

1.2 Appendix 2: Solution for decentralized economy

We begin with the derivation of (12). Using (7a) and the definition of \( P_i(t) \) [see the sentence just below (24)], we obtain

$$\begin{aligned} \frac{P_2(t)}{P_1(t)} = \frac{ b_1(t) }{ b_2(t) } N(t)^{-\sigma }. \end{aligned}$$
(27)

We substitute the first equation of (10) and (27) into the second equation of (25) to obtain

$$\begin{aligned} \frac{a_2(g_2)}{a_1(g_1) }N(t) = \left( \frac{1-\gamma }{\gamma } \right) \left[ \frac{b_1(t)}{b_2(t)}N(t)^{-\sigma } \right] ^{1-\varepsilon } = \left( \frac{1-\gamma }{\gamma } \right) \left[ \frac{b_{1,0}}{b_{2,0}} \left( \frac{g_{1}}{g_{2}} \right) ^t N(t)^{-\sigma } \right] ^{1-\varepsilon }. \end{aligned}$$

In the second equality, we use \( b_i(t)=b_{i,0}{g_i}^t \). Solving this equation for N(t) yields (12).

The remainder of this appendix provides the solutions of variables other than N(t) and checks the transversality condition. We first derive \( L^I(t) \) and \( L^P(t) \). We use (26) and (8) to obtain

$$\begin{aligned} L^I(t)=a_1(g_1)n_1(t+1) + a_2(g_2)n_2(t+1) = \frac{(1-\alpha ) E(t+1)}{ (1+r(t))w(t) }=\frac{(1-\alpha ) \beta E(t)}{ w(t) }. \end{aligned}$$
(28)

Using (9) and \( E(t) = E_1(t) + E_2(t) \), we obtain

$$\begin{aligned} L^P(t) = \alpha \frac{E_1(t)+ E_2(t)}{w(t)} =\alpha \frac{E(t)}{w(t)}. \end{aligned}$$
(29)

From (28) and (29), we have \( L^I(t)/L^P(t)= (1-\alpha )\beta /\alpha \). Using this equation and \( L=L^I(t)+L^P(t) \), we can derive

$$\begin{aligned} L^I(t)=\frac{(1-\alpha )\beta }{\alpha + (1-\alpha )\beta }L \equiv L^I \quad \text{ and } \quad L^P(t)=\frac{\alpha }{\alpha + (1-\alpha )\beta }L \equiv L^P. \end{aligned}$$
(30)

We next derive E(t) , \( E_1(t) \), and \( E_2(t) \). (29) and (30) imply

$$\begin{aligned} E(t)=\frac{Lw(t)}{\alpha + (1-\alpha )\beta }. \end{aligned}$$
(31)

From the first equation of (10), (12), and \( E(t) = E_1(t) + E_2(t) \), we have

$$\begin{aligned} E_1(t) = \frac{ 1 }{ 1+ \frac{a_2(g_2) }{a_1(g_1) } \Phi G^{t} } E(t) \quad \hbox { and }\quad E_2(t) = \frac{ \frac{a_2(g_2) }{ a_1(g_1) } \Phi {G}^{t} }{1+ \frac{a_2(g_2) }{a_1(g_1) } \Phi {G}^t } E(t). \end{aligned}$$
(32)

Because \( L^I(t) \) is constant, W(t) is also constant.

$$\begin{aligned} W(t) = a_(g_1) n_1(t+1) w(t) + a(g_2) n_2(t+1) w(t) = L^I w(t). \end{aligned}$$
(33)

Here, we use the first equality of (28). From (31) and (33), W(t) / E(t) becomes constant. Then, the transversality condition, \( \lim _{t \rightarrow +\infty } \beta ^t W(t)/E(t)=0 \), is satisfied.

From (12) and the first equality of (28), we derive

$$\begin{aligned} n_1(t) = \frac{L^I}{a_1(g_1)+a_2(g_2) \Phi {G}^t} \quad \text{ and } \quad n_2(t) = \frac{\Phi {G}^tL^I}{a_1(g_1)+a_2(g_2) \Phi {G}^t}. \end{aligned}$$
(34)

These two equations give the sequences of \( n_1(t) \) and \( n_2(t) \). If \( G=1 \), \( n_1(t) \) and \( n_2(t) \) are constant over time. If \( G>1 \), \( n_1(t) \) decreases over time and satisfies \( \lim _{t \rightarrow \infty } n_1(t)=0 \). In contrast, \( n_2(t) \) increases over time and satisfies \( \lim _{t \rightarrow \infty } n_2(t)=L^I/a_2(g_2) \). Because of \( L^I_i(t)=a_i(g_i)n_i(t+1) \) [see (8)], (34) drives the dynamics of \( L^I_i(t) \). Substituting (34) and \( b_{i,t}= b_{i,0} {g_i}^t \) into \( A_i(t) = b_i(t) n_i(t)^\sigma \), we obtain the dynamics of \( A_i(t) \).

The sequence of \( p_i(t) \) is given by \( p_i(t) =w(t) [\alpha b_i(t)]^{-1} = w(t) [\alpha b_{i,0} {g_i}^t]^{-1}\) [see (7a)]. The sequence of \( P_i(t) \) is given by \( P_i(t) =w(t) [\alpha b_i(t) {n_i(t)}^\sigma ]^{-1}= w(t)[\alpha b_{i,0} {g_i}^t {n_i(t)}^\sigma ]^{-1}\). Substituting \( P_i(t) \) into \(P(t) = [ \gamma ^\varepsilon {P_1(t)}^{1-\varepsilon } + (1-\gamma )^\varepsilon {P_2(t)}^{1-\varepsilon } ]^\frac{1}{1-\varepsilon } =1 \) yields

$$\begin{aligned} w(t) = [ \gamma ^\varepsilon \{\alpha b_1(t) n_1(t)^\sigma \}^{\varepsilon -1} + (1-\gamma )^\varepsilon \{\alpha b_2(t) n_2(t)^\sigma \}^{\varepsilon -1} ]^\frac{1}{\varepsilon -1}. \end{aligned}$$
(35)

Given the sequences of \( b_i(t) \) and \( n_i(t) \), the above equation gives the sequence of w(t) . (31) and (33) give the sequences of E(t) and W(t) , respectively. The sequences of \( E_1(t) \) and \( E_2(t) \) are given by (32). The sequences of \( p_i(t) \) and \( P_i(t) \) are given by \( p_i(t) = w(t) [\alpha b_{i,0} {g_i}^t]^{-1}\) and \( P_i(t) = w(t)[\alpha b_{i,0} {g_i}^t {n_i(t)}^\sigma ]^{-1}\), respectively. Now, we have the sequences of \(n_i(t)\), \( E_i(t) \), and \( p_i(t) \). Then, the sequences of \( x_i(t) \) and \( \pi _i(t) \) are given by (7b) and (7c), respectively.

From (26) and (31), we have \( D(t+1)/D(t)=E(t+1)/E(t) = w(t+1)/w(t) \). Using \( P(t) = [ \gamma ^\varepsilon {P_1(t)}^{1-\varepsilon } + (1-\gamma )^\varepsilon {P_2(t)}^{1-\varepsilon } ]^\frac{1}{1-\varepsilon } =1 \), we obtain the following relation:

$$\begin{aligned} 1=\left( \frac{P(t+1)}{P(t)} \right) ^{1-\varepsilon }&=\frac{ 1+\left( \frac{1-\gamma }{\gamma } \right) ^\varepsilon \left( \frac{P_2(t+1)}{P_1(t+1)} \right) ^{1-\varepsilon } }{ 1+\left( \frac{1-\gamma }{\gamma } \right) ^\varepsilon \left( \frac{P_2(t)}{P_1(t)} \right) ^{1-\varepsilon } } \times \left( \frac{P_1(t+1)}{P_1(t)} \right) ^{1-\varepsilon } \nonumber \\&=\frac{ 1+ \frac{E_2(t+1)}{E_1(t+1)} }{ 1+\frac{E_2(t)}{E_1(t)} } \times \left( \frac{P_1(t+1)}{P_1(t)} \right) ^{1-\varepsilon } \nonumber \\&=\frac{ 1+ \frac{a_2(g_2)}{a_1(g_1)} \Phi {G}^{t+1} }{ 1+\frac{a_2(g_2)}{a_1(g_1)} \Phi {G}^{t} } \times \left[ \frac{1}{g_1} \left( \frac{n_1(t)}{n_1(t+1)} \right) ^\sigma \frac{w(t+1)}{w(t)} \right] ^{1-\varepsilon } \nonumber \\&= \frac{1}{g_1^{1-\varepsilon }} \left( \frac{ 1+ \frac{a_2(g_2)}{a_1(g_1)} \Phi {G}^{t+1} }{ 1+\frac{a_2(g_2)}{a_1(g_1)} \Phi {G}^{t} } \right) ^{1+\sigma (1-\varepsilon )} \left[ \frac{w(t+1)}{w(t)} \right] ^{1-\varepsilon } . \end{aligned}$$
(36)

The second equality uses the second equation of (25). The third equality uses the first equation of (10), (12), and \( P_1(t) = w(t) [\alpha b_{1,0} {g_1}^t {n_1(t)}^\sigma ]^{-1}\). The last equality uses the first equation of (34). From (36) and \( D(t+1)/D(t)=w(t+1)/w(t) \), we obtain

$$\begin{aligned} \frac{D(t+1)}{D(t)} =\frac{w(t+1)}{w(t)} = g_1 \times \left( \frac{ 1+ \frac{a_2(g_2)}{a_1(g_1)}\Phi {G}^{t} }{ 1+\frac{a_2(g_2)}{a_1(g_1)}\Phi {G}^{t+1} }\right) ^{1/\phi }, \end{aligned}$$
(37)

where \( \phi = \frac{ 1-\varepsilon }{1+\sigma (1-\varepsilon ) } \) is defined just below (12). If \(G=1\), \( D(t+1)/D(t) \) is constant. In contrast, if \({G}>1\), \( D(t+1)/D(t) \) decreases with t and satisfies \( \lim _{t \rightarrow +\infty } D(t+1)/D(t) = g_2 \). As in Acemoglu and Guerrieri (2008), the asymptotic constant growth path exists.

1.3 Appendix 3: Centralized economy

The first-order conditions for the central planner’s problems are

$$\begin{aligned}&\lambda _P(t) = \lambda _I(t), \end{aligned}$$
(38)
$$\begin{aligned}&\lambda _P(t) \frac{n_i^*(t)x_i^*(t)}{b_i^*(t)}=v(t)^\frac{1-\varepsilon }{\varepsilon } \gamma _i \left( n_i^*(t)^{\sigma +1} x_i^*(t) \right) ^\frac{\varepsilon -1}{\varepsilon }, \end{aligned}$$
(39)
$$\begin{aligned}&v(t)^\frac{1-\varepsilon }{\varepsilon } (\sigma +1) \gamma _i \left( n_i^*(t)^{\sigma +1} x_i^*(t) \right) ^\frac{\varepsilon -1}{\varepsilon }\nonumber \\&\quad = n_i^*(t) \left[ \lambda _P(t) \frac{x_i^*(t)}{b_i^*(t)} + \frac{\lambda _I(t-1)}{\beta } a_i\left( \frac{b_i^*(t)}{b_i^*(t-1)} \right) \right] , \end{aligned}$$
(40)
$$\begin{aligned}&\frac{\lambda _I(t-1)}{\beta } \frac{b_i^*(t)}{b_i^*(t-1)} a_i^\prime \left( \frac{b_i^*(t)}{b_i^*(t-1)} \right) n_i^*(t)\nonumber \\&\quad =\lambda _P(t) \frac{n_i^*(t)x_i^*(t)}{b_i^*(t)} + \lambda _I(t) \frac{b_i^*(t+1)}{b_i^*(t)} a_i^\prime \left( \frac{b_i^*(t+1)}{b_i^*(t)} \right) n_i^*(t+1), \end{aligned}$$
(41)

where \( i= \)1 or 2, \( \gamma _1=\gamma \), \( \gamma _2=1-\gamma \), and \( \lambda _P(t) \) and \( \lambda _I(t) \) are the Lagrangian multipliers associated with (15) and (16), respectively. We must have \( L^{P*}(t) \in (0, L) \) because if \( L^{P*}(t) = L \), no intermediate firms produce in period t and if \( L^{P*}(t) = 0 \), there are no intermediate goods firms in period \( t+1 \). Then, (38) holds. We denote \( \lambda (t) \equiv \lambda _P(t) = \lambda _I(t) \).

Using (15), (39), and the definition of v(t) , we obtain \( {L^P}^*(t) = \frac{v(t)^\frac{1-\varepsilon }{\varepsilon }}{\lambda (t)} \sum _{i} \gamma _i \left( n_i^*(t)^{\sigma +1} x_i^*(t) \right) ^\frac{\varepsilon -1}{\varepsilon } = 1/\lambda (t) \).

From (15), (16), (40), the definition of v(t) , and \( {L^P}^*(t)= 1/\lambda (t) \), we can derive \( \sigma \beta = \lambda (t) {L^I}^*(t) \).

We substitute \( {L^P}^*(t)= 1/\lambda (t) \) and \( \sigma \beta = \lambda (t) {L^I}^*(t) \) into \( L={L^P}^*(t) + {L^I}^*(t) \) to obtain

$$\begin{aligned} \lambda (t) = \frac{1+\sigma \beta }{L} \equiv \lambda , \quad {L^P}^*(t) = \frac{L}{1+\sigma \beta } \equiv {L^P}^*, \quad \text{ and } \quad {L^I(t)}^* = \frac{\sigma \beta L}{1+\sigma \beta } \equiv {L^I}^*. \end{aligned}$$
(42)

We denote \( g_{i}^*(t) \equiv b_i^*(t)/b_i^*(t-1) \) and \( G^*(t) =[g_{1}^*(t)/g_{2}^*(t)]^\phi \). To simplify our expressions, we denote \( a_i \left( g_{i}^*(t) \right) =a_{i}^*(t) \). From (39) and (40), the following two relations are obtained:

$$\begin{aligned} \beta \sigma \frac{n_i^*(t)x_i^*(t)}{b_i^*(t)}&= a_{i}^*(t)n_i^*(t), \end{aligned}$$
(43)
$$\begin{aligned} \frac{x_2^*(t)}{x_1^*(t)}&= \frac{a_{2}^*(t)}{a_{1}^*(t)}\frac{b_2^*(t)}{b_1^*(t)}. \end{aligned}$$
(44)

In addition, from (39), (40), and (43), we obtain

$$\begin{aligned} \frac{1-\gamma }{\gamma } \left[ \frac{ n_2^*(t)^{\sigma +1} x_2^*(t) }{ n_1^*(t)^{\sigma +1} x_1^*(t) } \right] ^\frac{\varepsilon -1}{\varepsilon } = \frac{a_{2}^*(t)}{a_{1}^*(t)}\frac{n_2^*(t)}{n_1^*(t)}. \end{aligned}$$
(45)

Using (44) and (45), we obtain

$$\begin{aligned} \frac{n_2^*(t)}{n_1^*(t)} = \left[ \left( \frac{1-\gamma }{\gamma }\right) ^\varepsilon \frac{a_{1}^*(t)}{a_{2}^*(t)}\right] ^\frac{1}{1+\sigma (1-\varepsilon )} \left( \frac{b_1^*(t)}{b_2^*(t)}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} . \end{aligned}$$
(46)

From (16) and (46), we have

$$\begin{aligned} n_1^*(t)&=\frac{ {L^I}^* }{ a_{1}^*(t) + a_{2}^*(t)\left[ \left( \frac{1-\gamma }{\gamma }\right) ^\varepsilon \frac{a_{1}^*(t)}{a_{2}^*(t)}\right] ^\frac{1}{1+\sigma (1-\varepsilon )} \left( \frac{b_1^*(t)}{b_2^*(t)}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} }, \end{aligned}$$
(47)
$$\begin{aligned} n_2^*(t)&=\frac{\left[ \left( \frac{1-\gamma }{\gamma }\right) ^\varepsilon \frac{a_{1}^*(t)}{a_{2}^*(t)}\right] ^\frac{1}{1+\sigma (1-\varepsilon )} \left( \frac{b_1^*(t)}{b_2^*(t)}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} {L^I}^* }{ a_{1}^*(t) + a_{2}^*(t)\left[ \left( \frac{1-\gamma }{\gamma }\right) ^\varepsilon \frac{a_{1}^*(t)}{a_{2}^*(t)}\right] ^\frac{1}{1+\sigma (1-\varepsilon )} \left( \frac{b_1^*(t)}{b_2^*(t)}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} }. \end{aligned}$$
(48)

Using (41), (43), and \( a_i(g)= f_i e^{\xi _i g} \), we derive

$$\begin{aligned} g_{i}^*(t) = \frac{1}{\sigma \xi _i} + \beta \frac{a_{i}^*(t+1) n_i^*(t+1) }{a_{i}^*(t) n_i^*(t)}g_{i}^*(t+1). \end{aligned}$$
(49)

This equation shows \( g_{i}^*(t) \ge (\sigma \xi _i)^{-1} >0 \). (47) and (48) show \( n_i^*(t)>0 \). If \( g_{i}^*(t)=+\infty \), the resource constraint implies that \( n_i^*(t) \) must be strictly equal to zero. To ensure \( n_i^*(t)>0 \), \( g_{i}^*(t)<+\infty \) must hold.

We now prove a series of lemmas to prove Proposition 2 in Sect. 4.

Lemma 1

Suppose that Assumption (i) is satisfied (\( 0<\varepsilon <1 \)) and that \( \xi _1=\xi _2 \). There exists a solution for the central planner’s problem such that \( g^*_1(t)=g^*_2(t)= [ (1-\beta ) \sigma \xi ]^{-1} \) and \( G^*(t) =1 \) for all \( t=0,1,\ldots \).

Proof

We guess that \( g_{1}^*(t) =g_{2}^*(t) =g^*(t) \) holds for all \( t=0,1,\ldots \). Then, we verify this guess and derive the expression for \( g^*(t) \).

Because \( g_{1}^*(t) =g_{2}^*(t) =g^*(t) \) holds for all \( t=0,1,\ldots \), we have \( a_{1}^*(t)/a_{2}^*(t) = f_1/f_2 \) and \( b_{1}^*(t)/b_{2}^*(t) = b_{1,0}/b_{2,0} \). Consequently, (47) and (48) show that \( a_{i}^*(t)n_i^*(t) \) becomes constant. We can rewrite (49) as \( g_{i}^*(t) = 1/(\sigma \xi ) + \beta g_{i}^*(t+1) \). This difference equation has a unique and unstable steady state, where \( g_{i}^*(t)= [ (1-\beta ) \sigma \xi ]^{-1} \). Because \( 0<g_{i}^*(t) < +\infty \), \( g_{i}^*(t) \) always stays at the steady state, which implies \( g_{1}^*(t) =g_{2}^*(t)= [ (1-\beta ) \sigma \xi ]^{-1} \) and \( G^*(t) =1 \) for all \( t=0,1,\ldots \). \(\square \)

Lemma 2

Suppose that Assumption (i) is satisfied (\( 0<\varepsilon <1 \)) and \( \xi _1=\xi _2 \). There exists a solution for the central planner’s problem such that \( G_{n}^*(t)=G_{L^I}^*(t)=G_{L^P}^*(t) = G^*(t) =1 \) for all \( t=0,1,\ldots \).

Proof

We have \( {L^I}^*_i(t) = a_i^*(t+1)n_i^*(t+1) \) and \( {L^p}^*_i(t) = n_i^*(t)x_i^*(t)/b_i^*(t) \). From (43), we have \( {L^I_2}^*(t-1)/{L^I_1}^*(t-1)={L^P_2}^*(t)/{L^P_1}^*(t) = a_2^*(t)n_2^*(t) / [ a_1^*(t)n_1^*(t) ] \). If \( \xi _1=\xi _2 \), we have \( g_{1}^*(t) =g_{2}^*(t)= [ (1-\beta ) \sigma \xi ]^{-1} \) and \( G^*(t)=1 \) from Lemma 1. Then, \( a_i^*(t) \) and \( n_2^*(t)/n_1^*(t) \) become constant. Consequently, \( {L^I_2}^*(t-1)/{L^I_1}^*(t-1) \) and \( {L^P_2}^*(t)/{L^P_1}^*(t) \) are constant. \(\square \)

Lemma 3

Suppose that Assumption (i) is satisfied (\( 0<\varepsilon <1 \)) and that \( \xi _1<\xi _2 \). There exists a solution for the central planner’s problem such that \( \lim _{t\rightarrow \infty } g^*_1(t)=g^*_1 \) and \( \lim _{t\rightarrow \infty } g^*_2(t)= [ (1-\beta ) \sigma \xi _2 ]^{-1} \equiv g^*_2 \), where \( g^*_1 \) is defined by (17) and satisfies \( g^*_2< g_{1}^* < [ (1-\beta ) \sigma \xi _1 ]^{-1}\).

Proof

We guess \( \lim _{t\rightarrow \infty } g^*_1(t)=g^*_1> g^*_2=\lim _{t\rightarrow \infty } g^*_2(t) \). Then, we verify this guess and derive the expressions for \( g^*_1 \) and \( g^*_2 \).

Because of \( \lim _{t\rightarrow \infty } g^*_1(t)=g^*_1> g^*_2=\lim _{t\rightarrow \infty } g^*_2(t) \), we have \( \lim _{t\rightarrow \infty } a_{1}^*(t)/a_{2}^*(t) \) is constant and \( \lim _{t\rightarrow \infty } b_{1}^*(t)/b_{2}^*(t) = +\infty \). Consequently, (48) shows \( \lim _{t\rightarrow \infty } a_{2}^*(t)n_2^*(t)= {L^I}^* \). Then, for \( i=2 \) and a sufficiently large t, (49) can be approximated as \( g_{2}^*(t) = 1/(\sigma \xi _2) + \beta g_{2}^*(t+1) \). This difference equation has a unique and unstable steady state, where \( g_2^*(t)= [ (1-\beta ) \sigma \xi _2 ]^{-1} \). Because \( 0<g_2^*(t) < +\infty \), \( g_2^*(t) \) must stay at the steady state for large t, which implies \( \lim _{t\rightarrow \infty } g^*_2(t)= [ (1-\beta ) \sigma \xi _2 ]^{-1} \equiv g^*_2 \).

We rewrite (46) as

$$\begin{aligned} \frac{a_{2}^*(t) n_2^*(t)}{a_{1}^*(t) n_1^*(t)} = \left( \frac{1-\gamma }{\gamma }\right) ^\frac{\varepsilon }{1+\sigma (1-\varepsilon )} \left[ \frac{a_{2}^*(t)}{a_{1}^*(t)}\right] ^\frac{\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )} \left( \frac{b_1^*(t)}{b_2^*(t)}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )}. \end{aligned}$$
(50)

For sufficiently large t, we have \( a_{2}^*(t) n_2^*(t)=a_{2}^*(t+1) n_2^*(t+1)= {L^I}^* \) and \( g_{2}^*(t) = [ (1-\beta ) \sigma \xi _2 ]^{-1} \). Furthermore, we have \( b_i^*(t+1)=g_{i}^*(t+1)b_i^*(t) \). Using these equations and (50), we obtain

$$\begin{aligned} \frac{a_{1}^*(t+1) n_1^*(t+1)}{a_{1}^*(t) n_1^*(t)} = \left[ \frac{a_{1}^*(t+1)}{a_{1}^*(t)}\right] ^\frac{\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )} \left( \frac{[ (1-\beta ) \sigma \xi _2 ]^{-1}}{g_{1}^*(t+1)}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )}. \end{aligned}$$

for a sufficiently large t. Using this equation and (49), we obtain

$$\begin{aligned}&a_{1}^*(t)^\frac{\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )} \left( g_{1}^*(t)-\frac{1}{\sigma \xi _1} \right) \\&\quad = \beta a_{1}^*(t+1)^\frac{\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )} \left( [ (1-\beta ) \sigma \xi _2 ]^{-1}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} {g_{1}^*(t+1)}^\frac{\varepsilon +\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )}, \end{aligned}$$

for sufficiently large t. Because \( a_{1}^*(t)=a_1 \left( g_{1}^*(t) \right) = f_1 e^{\xi _1 g_{1}^*(t) } \), the above equation is the first-order difference equation for \( g_{1}^*(t) \) and it has a unique and unstable steady state. Because \( 0<g_{1}^*(t) < +\infty \), \( g_{1}^*(t) \) must stay at the steady state for large t. Because the above equation reduces to (17) in the steady state, \( g^*_1 \) is defined by (17).

We evaluate the RHS of (17) at \( g_{1}^* = [ (1-\beta ) \sigma \xi _2 ]^{-1} \) and \( g_{1}^* = [ (1-\beta ) \sigma \xi _1 ]^{-1} \), where \( [ (1-\beta ) \sigma \xi _2 ]^{-1} < [ (1-\beta ) \sigma \xi _1 ]^{-1} \).

$$\begin{aligned} \left. RHS \right| _{g_{1}^* = [ (1-\beta ) \sigma \xi _2 ]^{-1}}&=\frac{1}{\sigma \xi _1} + \beta \left( [ (1-\beta ) \sigma \xi _1 ]^{-1}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} \left( [ (1-\beta ) \sigma \xi _2 ]^{-1} \right) ^\frac{\varepsilon +\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )} \\&>\frac{1}{\sigma \xi _2} + \beta \left( [ (1-\beta ) \sigma \xi _2 ]^{-1}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} \left( [ (1-\beta ) \sigma \xi _2 ]^{-1} \right) ^\frac{\varepsilon +\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )}\\&= [ (1-\beta ) \sigma \xi _2 ]^{-1} = g^*_2, \\ \left. RHS \right| _{g_{1}^* = [ (1-\beta ) \sigma \xi _1 ]^{-1}}&=\frac{1}{\sigma \xi _1} + \beta \left( [ (1-\beta ) \sigma \xi _1 ]^{-1}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} \left( [ (1-\beta ) \sigma \xi _2 ]^{-1} \right) ^\frac{\varepsilon +\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )}\\&<\frac{1}{\sigma \xi _1} + \beta \left( [ (1-\beta ) \sigma \xi _1 ]^{-1}\right) ^\frac{1-\varepsilon }{1+\sigma (1-\varepsilon )} \left( [ (1-\beta ) \sigma \xi _1 ]^{-1} \right) ^\frac{\varepsilon +\sigma (1-\varepsilon )}{1+\sigma (1-\varepsilon )}\\&= [ (1-\beta ) \sigma \xi _1 ]^{-1}, \end{aligned}$$

Then, \( g_{1}^* \) is between \( [ (1-\beta ) \sigma \xi _2 ]^{-1}\) and \( [ (1-\beta ) \sigma \xi _1 ]^{-1} \). \(\square \)

Lemma 4

Suppose that Assumption (i) is satisfied (\( 0<\varepsilon <1 \)) and that \( \xi _1<\xi _2 \). There exists a solution for the central planner’s problem such that \( \lim _{t\rightarrow \infty } G^*(t)=\lim _{t\rightarrow \infty } G_{n}^*(t)=\lim _{t\rightarrow \infty } G_{L^I}^*(t)=\lim _{t\rightarrow \infty } G_{L^P}^*(t) =\left( g^*_1/g^*_2 \right) ^\phi \equiv G^* >1 \).

Proof

The proof is similar to that of Lemma 2. Lemma 3 shows that if \( \xi _1<\xi _2 \), we have \( \lim _{t\rightarrow \infty } G^*(t)=\left( g^*_1/g^*_2 \right) ^\phi \equiv G^* >1 \). We know from (46) that \( \lim _{t\rightarrow \infty } G_{n}^*(t)= \lim _{t\rightarrow \infty } G^*(t) \). In addition, (43) implies that \( \lim _{t\rightarrow \infty } G_{n}^*(t)=\lim _{t\rightarrow \infty } G_{L^I}^*(t)=\lim _{t\rightarrow \infty } G_{L^P}^*(t) \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hori, T., Mizutani, N. & Uchino, T. Endogenous structural change, aggregate balanced growth, and optimality. Econ Theory 65, 125–153 (2018). https://doi.org/10.1007/s00199-016-1012-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-016-1012-1

Keywords

JEL Classification

Navigation