Skip to main content
Log in

On impatience, temptation and Ramsey’s conjecture

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

Abstract

This article explores the implications of the introduction of self-control costs and temptation motives in intertemporal preferences in an elementary competitive equilibrium. We let heterogeneous agents differ in both their discount parameters and their temptation motives, and the degree of financial-market imperfections vary, and establish their implications for the long-run value of the capital stock and the underlying long-run distribution of consumption and wealth. The results differ from those obtained in a standard Ramsey benchmark model, in that long-run distributions are commonly non-degenerate.

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. This was formally and separately established by Becker (1980) under a nonnegativity constraint on household capital holdings and Bewley (1982) for an economy without borrowing constraints.

  2. For elaborations on related topics, see Becker et al. (2014), Becker and Foias (1994) and Sorger (1994).

  3. \( D u(c):=\partial u(c)/\partial c\).

  4. This property holds for a Cobb–Douglas production function or for a C.E.S. production function when the elasticity of substitution is greater than one. For \(F(K,{1})=K^{\alpha }, \alpha \in \, ]{0},{ 1}[, \lim _{{K}\rightarrow {0}}[ D _{K}F(K, {1})K-\eta K]/ D _{ {L}}F(K, {1})=\alpha /({ 1}-\alpha )\), while \(\lim _{{K}\rightarrow { 0}}[ D _{K}F(K, {1})K-\eta K]/ D _{ {L}}F(K, {1})={0}\) for a C.E.S production function.

References

  • Becker, R.: On the long-run stationary competitive equilibrium in a simple dynamic model of equilibrium with heterogeneous households. Q. J. Econ. 95, 375–382 (1980)

    Article  Google Scholar 

  • Becker, R., Dubey, R., Mitra, T.: On Ramsey equilibrium: capital ownership pattern and inefficiency. Econ. Theory 55, 565–600 (2014)

    Article  Google Scholar 

  • Becker, R., Foias, C.: The local bifurcation of Ramsey equilibrium. Econ. Theory 4, 719–744 (1994)

    Article  Google Scholar 

  • Bewley, T.: An integration of equilibrium theory and turnpike theory. J. Math. Econ. 10, 233–267 (1982)

    Article  Google Scholar 

  • Drugeon, J.-P., Wigniolle, B.: On Time-Consistent Policy Rules for Heterogenous Discounting Programs. Mimeograph (2015)

  • Gul, F., Pesendorfer, W.: Self-control and the theory of consumption. Econometrica 72, 119–158 (2004)

    Article  Google Scholar 

  • Kurz, M.: Optimal economic growth with wealth effects. Int. Econ. Rev. 9, 348–357 (1968)

    Article  Google Scholar 

  • Le Van, C., Vailakis, Y.: Existence of a competitive equilibrium in a one sector growth model with heterogeneous agents and irreversible investment. Econ. Theory 22, 743–771 (2003)

    Article  Google Scholar 

  • Mitra, T., Sorger, G.: On Ramsey’s conjecture. J. Econ. Theory 148, 1953–1973 (2013)

    Article  Google Scholar 

  • Ramsey, F.: A mathematical theory of savings. Econ. J. 38, 543–549 (1928)

    Article  Google Scholar 

  • Sorger, G.: On the structure of Ramsey equilibrium: cycles, indeterminacy, and sunspots. Econ. Theory 4, 745–764 (1994)

    Article  Google Scholar 

  • Zou, H.-F.: The spirit of capitalism and long-run growth. Eur. J. Polit. Econ. 10, 279–293 (1994)

    Article  Google Scholar 

Download references

Acknowledgments

This research was carried out thanks to the Novo Tempus research grant, ANR-12-BSH1-0007, Program BSH1-2012. The authors would like to thank participants at the PET conference, Lisbon, June 2013, and the SAET conference, Paris, July 2013. They are also grateful to two referees for their insightful comments that resulted in a significant improvement in the quality of the exposition.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bertrand Wigniolle.

Appendices

Appendix 1: Proof of Proposition 1

The satisfaction of \(R= D _{K}F(K,{1})+{1}-\eta \) and \(w= D_{{L}}F(K,{1})/n\) directly results from Eqs. (3h) and (3i) when they are considered at the long-run stationary competitive equilibrium. In parallel, the consideration of (3g) allows us to derive \(h=w/(R-{1})\), whereas (3d) will hold in a stationary state. Equations (3a), (3b), (3c), (3e) and (3f) then allow us to obtain Propositions 1(i) and 1(ii) through the following statement:

Claim

Consider the optimal solution:

  1. (i)

    a stationary optimal solution for an unconstrained individual corresponds to stationary values \(c^{i}\) and \(x^{i}\) such that:

    $$\begin{aligned} R\delta _{i}&=\dfrac{ D u(c^{i})+\gamma _{i} D v(c^{i})}{ D u(c^{i})+\gamma _{i}\bigl ( D v(c^{i})- D v(x^{i}+\lambda h)\bigr )}, \end{aligned}$$

    where

    1. a/

      \(x^{i}=(Rc^{i}-w)/(R-{1})\);

    2. b/

      the satisfaction of \(c^{i}<x^{i}+\lambda h\) boils down to \(c^{i}>w({1}-\lambda ) \) or \(\gamma _{i} D v[({ 1}-\lambda )w]<\bigl (\delta _{i}R-{1}\bigr ) D u[({1}-\lambda )w]\);

    3. c/

      the existence of such a solution requires \(R\delta _{i}>{1}\) and implies \(R>{1}\);

  2. (ii)

    a stationary optimal solution for a constrained individual corresponds to stationary values \(c^{i}\) and \(x^{i}\) such that

    1. a/

      \(c^{i}=({1}-\lambda )w\);

    2. b/

      \(x^{i}=[R({1}-\lambda )w-w]/({R-{1}})\);

    3. c/

      the existence of such a solution requires \(\nu ^{i}>{0}\), a condition that can be restated as: \(\gamma _{i} D v(({1}-\lambda )w)>(\delta _{i} R-{1}) D u(({ 1}-\lambda )w)\).

Proof

  1. (i)

    For an unconstrained individual, \(\nu _{t}^{i}={0}\) and from (3b) and (3c):

    $$\begin{aligned} R_{t+{1}}\delta _{i}\bigl [ D u\bigl (c_{t+ { 1}}^{i}\bigr )+\gamma _{i} D v\bigl (c_{t+ { 1}}^{i}\bigr )-\gamma _{i} D v\bigl (x_{t+{1}}^{i}+\lambda h_{t+2}\bigr )\bigr ]= D u\bigl (c_{t}^{i}\bigr )+\gamma _{i} D v\bigl (c_{t}^{i}\bigr ). \end{aligned}$$
    (*)

    Rearranging and considering stationary trajectories in (*), we obtain the conditions of the statement.

  2. (ii)

    The basic form of the condition \(\nu _{i}>{0}\) is:

    $$\begin{aligned} \nu ^{i}= D u(c^{i})+\gamma _{i} D v(c^{i})-\delta _{i}\mu ^{i} R>{ 0}, \end{aligned}$$

    for \(\mu ^{i}= D u(c^{i})+\gamma _{i} D v(c^{i})-\gamma _{i} D v(x^{i}+\lambda h), c^{i}=x^{i}+\lambda h\), \(\mu ^{i}= D u(c^{i})\), that yields, merging and rearranging, the above statement.

Equation (4a) then corresponds to the resource constraints of the economy. It is derived by combining Eqs. (3j), (3h), (3i) and (3a) when considered at the stationary state. Equation (3j) allows us to state \(\sum _{i= {1}}^{n}x^{i}=RK +nw\) while, from Eq. (3a), \(\sum _{i= {1}}^{n}x^{i}(R-{1})=R\sum _{i={1}}^{n}c^{i}-nw\). Finally, eliminating \(\sum _{i={1}}^ {n}x^{i}\) between these equations yields:

$$\begin{aligned} (R-{1})K+nw=\sum _{i={1}}^{n}c^{i}. \end{aligned}$$

Making use of (3h) and (3i), we obtain Eq. (4a). \(\square \)

Appendix 2: Proof of Lemma 1

(i) From (5), the equation defining stationary competitive equilibrium consumption is:

$$\begin{aligned} D u(c^{i})+\gamma _{i} D v(c^{i})=\dfrac{\delta _{i} R}{\delta _{i} R-{1}}\gamma _{i} D v\biggl [\dfrac{Rc^{i}-({1}-\lambda )w}{R-{1}}\biggr ], \end{aligned}$$

We then require a solution to the above equation such that \(c>({ 1}-\lambda )w\). It is worth noting that this can conveniently be restated as \( {H}^{\ell }(c^{i}, \gamma _{i})={H}^{r}(c^{i}, \delta _{i}, \gamma _{i}, R,w,\lambda ), \) where the simultaneous holding of \( D _{c^{i}}{H}^{\ell }(c^{i}, \gamma _{i})<{0}\) and \( D_{c^{i}}{H}^{r}(c^{i},\delta _{i}, \gamma _{i}, R,w,\lambda )>{0}\) first ensures that there exists at most one solution. Turning to existence, and for \(c^{i}\rightarrow ({1}-\lambda )w\), we have:

$$\begin{aligned}&{H}^{\ell }(c^{i},\gamma _{i})\rightarrow D u[({1}-\lambda )w]+\gamma _{i} D v[({1}-\lambda )w],\\&{H}^{r}(c^{i}, \delta _{i}, \gamma _{i}, R,w)\rightarrow \dfrac{\delta _{i} R}{\delta _{i} R-{1}}\gamma _{i} D v[({1}-\lambda )w]. \end{aligned}$$

From the previous assumption, \(\gamma _{i} D v[({ 1}-\lambda )w]<(\delta _{i} R-{1}) D u[({1}-\lambda )w]\), and hence

$$\begin{aligned} {H}^{\ell }(({1}-\lambda )w, \gamma _{i})>{H}^{r}(({1}-\lambda )w, \delta _{i}, \gamma _{i}, R,w,\lambda ). \end{aligned}$$

Conversely, and for \(c^{i}\rightarrow +\infty \), whereas \( D u(c^{i})\rightarrow { 0}, [Rc^{i}-({1}-\lambda )w]/(R-{1})>c^{i}\) as \(c^{i}>({ 1}-\lambda )w\). This gives us

$$\begin{aligned} {H}^{\ell }(+\infty , \gamma _{i})<{H}^{r}(+\infty , \delta _{i}, \gamma _{i}, R,w,\lambda ), \end{aligned}$$

so that we have the existence and uniqueness results for \(c^{i}\).

(ii) Looking then at comparative static properties, it first emerges that \( D _{{ R}}C(\delta _{i}, \gamma _{i}, R, w, \lambda )>{0}\) as \( D _{{ R}}{H}^{r}(c^{i}, \delta _{i}, \gamma _{i}, R,w)<{0}\). We have

$$\begin{aligned} \text{ Sgn }\biggl [ D _{{ R}}\biggl (\dfrac{Rc^{i}-({1}-\lambda )w}{R-{1}} \biggr )\biggr ]&=\text{ Sgn }[c^{i}(R-{1})-Rc^{i} + ({1}-\lambda )w]\\&=\text{ Sgn }[({1}-\lambda )w-c^{i}]\\&<{0}. \end{aligned}$$

Establishing that \( D _{w}C(\delta _{i}, \gamma _{i}, R, w, \lambda )\geqslant {0}\) then comes from \( D _{w}{H}^{r}(c^{i}, \delta _{i}, \gamma _{i}, R,w,\lambda )\,\leqslant \,0\), noting that, for \(\lambda ={1}\), this simplifies to \( D _{w}C(\delta _{i}, \gamma _{i}, R, w, \lambda )= 0\). In the same way, we can easily show that \( D _{\delta _{i}}C(\delta _{i}, \gamma _{i}, R, w, \lambda )>{ 0}\) and \( D _{\lambda }C(\delta _{i}, \gamma _{i}, R, w, \lambda )<{0}\) are direct by-products of \( D _{\delta _{i}}{H}^{r}(c^{i}, \delta _{i}, \gamma _{i}, R,w,\lambda )< 0\) and \( D _{\lambda }{H}^{r}(c^{i}, \delta _{i}, \gamma _{i}, R,w,\lambda )\geqslant {0}\). Finally, showing that \( D _{\gamma _{i}}C(\delta _{i}, \gamma _{i}, R, w, \lambda )<{0}\) proceeds from the restatement of the defining equation of the stationary values for \(c^{i}\) as:

$$\begin{aligned} \dfrac{R\delta _{i}}{R\delta _{i}-{1}}&=\dfrac{ D u(c^{i})+\gamma _{i} D v(c^{i})}{\gamma _{i} D v[(Rc^{i}-({ 1}-\lambda )w)/(R-{1})]}\equiv {G}(c^{i}, \gamma _{i}, w, R). \end{aligned}$$
(**)

As the function \({G}(c^{i}, \gamma _{i}, w, R)\) is decreasing in both \(\gamma _{i}\) and \(c^{i}\), we have \( D _{\gamma _{i}}C(\delta _{i}, \gamma _{i}, R, w, \lambda )<{0}\).

The expression \(C(\delta _{i}, \gamma _{i}, R,w,\lambda )\) being monotone increasing as a function of R, it has a limit that is either finite or infinite. First assuming that \(\ell <+\infty \) and from equation (**), we have that:

$$\begin{aligned} 1=\dfrac{ D u(\ell )+\gamma _{i} D v(\ell )}{\gamma _{i} D v(\ell )}, \end{aligned}$$

which is not possible for \(\ell <+\infty \), so that \(\ell =+\infty \).

Appendix 3: Proof of Proposition 2

(i)–(ii) In order to save on notation, we consider the function \(c^{i}(K)\) as defined by:

$$\begin{aligned} {\left\{ \begin{array}{ll} c^{i}(K)=\mathscr {C}\bigl (\delta _{i}, \gamma _{i},K,{1}\bigr )&{}\quad \text{ for }\quad K\in \left]{0},\widetilde{K}_{i}\right[;\\ c^{i}(K)={0}&{}\quad \text{ for }\quad K\geqslant \widetilde{K}_{i}. \end{array}\right. } \end{aligned}$$

The existence of an equilibrium is associated with a value of K such that:

$$\begin{aligned} F(K,{1})-\eta K=\sum _{i={1}}^{n}c^{i}(K). \end{aligned}$$

Denoting this equation by \({H}^{\ell }\left( K\right) ={H}^{r}\left( K\right) \), as \({1}/\delta _{{1}}>{1}\), we have that \(\widetilde{K}_{{ 1}}<\widehat{K}\). The function \({H}^{\ell }(K)\) is then increasing over \(]{0},\widetilde{K}_{{1}}[\). On the contrary, the function \({H}^{r}(K)\) is decreasing over \(]{0},\widetilde{K}_{ {1}}[\), with \({H}^{r}(\widetilde{K}_{{1}})={0}\) and \(\lim _{{K}\rightarrow {0}}{H}^{r}(K)=+\infty \). These properties ensure the existence and uniqueness of \(K\in \,]{0}, \widetilde{K}_{{ 1}}[\) such that \({H}^{\ell }(K)={H}^{r}(K)\).

Appendix 4: Proof of Lemma 2

(i)–(ii) Uniqueness is established by noting that the L.H.S. of (7) is non-decreasing as a function of K, while the R.H.S. falls as a function of K.

The existence argument in turn results from noting that, for \(K\rightarrow \widetilde{K}_{i}\),

$$\begin{aligned}&\gamma _{i} D v\left[ ({1}-\lambda ) D _{ {L}}F(K,{1})/n\right] \rightarrow \gamma _{i} D v\left[ D _{ {L}}F(\widetilde{K}_{i},{1})/n\right] >{0},\\&\Bigl \{\delta _{i}\bigl [ D _{K}F(K,{1})+({1}-\eta )\bigr ]-{ 1}\Bigr \} D u\left[ ({1}-\lambda ) D _{ {L}}F(K,{1})/n\right] \rightarrow {0}. \end{aligned}$$

At the same time, and for \(K\rightarrow {0}\),

$$\begin{aligned}&\gamma _{i} D v\left[ ({1}-\lambda ) D _{ {L}}F(K,{1})/n\right] \rightarrow \gamma _{i} D v\left[ ({1}-\lambda ) D _{ {L}}F({0},{1})/n\right] ,\\&\Bigl \{\delta _{i}\bigl [ D _{K}F(K,{1})+({1}-\eta )\bigr ]-{ 1}\Bigr \} D u\left[ ({1}-\lambda ) D _{ {L}}F(K,{1})/n\right] \rightarrow +\infty . \end{aligned}$$

This gives us the existence result.

Appendix 5: Proof of Lemma 3

  1. (i)

    By definition, \(\xi ^{i}(K)\) is a solution to:

    $$\begin{aligned}&\dfrac{\delta _{i}R(K)\gamma _{i}}{\delta _{i}R(K)-{1}} =\dfrac{ D u\bigl (\xi ^{i}w(K)\bigr )+\gamma _{i} D v\bigl (\xi ^{i}w(K)\bigr )}{ D v\bigl (\bigl (R(K)\xi ^{i}-({1}-\lambda )\bigr )w(K)/(R(K)-{ 1})\bigr )} \end{aligned}$$

    The L.H.S. of the above equation increases as a function of K, while the R.H.S. may conveniently be reformulated as \({H}^{u}(K, \xi ^{i})/{H}^{\ell }(K,\xi ^{i})\). We then see that \({H}^{u}\) falls with K while \({H}^{\ell }\) increases with both K and \(\xi ^{i}, \left[ R\xi _{i}-({1}-\lambda )\right] /\left( R-{ 1}\right) \) being a decreasing function of R, which yields the monotonicity property for \(\xi ^{i}\).

  2. (ii)

    As the function \(\xi ^{i}(K)\) is decreasing monotone in K, it has a limit at zero that is either \(\ell >{0}\) or \(+\infty \). Assuming first that this is finite and \(\lim _{{K}\rightarrow {0}}\xi ^{i}(K)=\ell \), the above equation yields:

    $$\begin{aligned} \gamma _{i}=\dfrac{ D u\bigl (w({0})\ell \bigr )+\gamma _{i} D v\bigl (w({0})\ell \bigr )}{ D v\bigl (w({0})\ell \bigr )}, \end{aligned}$$

    which is impossible (a contradiction), so that \(\ell =+\infty \).

Appendix 6: Proof of Proposition 3

(i) For a constrained individual i, it is convenient to extend the function \(\xi ^{i}(K)\) over the interval \([\overline{K}_{i},+\infty [\) with \(\xi ^{i}(K)={1}-\lambda \). The existence of an equilibrium is hence ensured by a value of K such that:

$$\begin{aligned} F(K,{1})-\eta K=\sum _{i={1}}^{n}\xi ^{i}(K)\dfrac{ D _{ {L}}F(K,{1})}{n}, \end{aligned}$$

hence, making use of the Euler relationship,

$$\begin{aligned} \dfrac{ D _{K}F(K,{1})K -\eta K}{ D _{ {L}}F(K,{ 1})}=\dfrac{1}{n}\sum _{i={1}}^{n}\bigl [\xi ^{i}(K)-{1}\bigr ]. \end{aligned}$$

or \({H}^{\ell }(K)={H}^{r}(K)\). First, and for \(K\rightarrow {0}, {H}^{\ell }(K)<{H}^{r}(K)\). While \({H}^{\ell }(K)\) is bounded under Assumption T.2, we have that \({H}^{r}(K)\rightarrow \infty \) as a direct corollary of Lemma 3(ii).

Now, for K such that:

$$\begin{aligned} K=\overline{K}_{{1}}>\overline{K}_{{ 2}}>\cdots >\overline{K}_{n}, \end{aligned}$$

\(\xi ^{i}(K)={1}-\lambda \) for any \(i={1}, \ldots ,n\) and \({H}^{r}(\overline{K}_{{1}})=-\lambda \). Further, \( D _{K}F(\overline{K}_{{1}},{1})\overline{K}_{{1}}-\eta \overline{K}_{{1}}>{0}\), since \(\overline{K}_{{ 1}}<\widehat{K}\), for \(\widehat{K}\) that satisfies \( D _{K}F(\widehat{K},{1})=\eta \). This implies that \({H}^{\ell }(\overline{K}_{{1}})>{H}^{r}(\overline{K}_{{ 1}})\). The existence of a \(K\in \,]{0}, \overline{K}_{{ 1}}[\) such that \({H}^{\ell }(K_{{1}})={H}^{r}(K_{{1}})\) is then established.

Appendix 7: Proof of Proposition 4

  1. (i)

    This is straightforward as for \(i>\xi , c^{i}=0\) and the parameter \(\gamma _{i}\) has no impact on the condition \(\delta _{i}R\leqslant 1\), that eventually implies \(c^{i}=0\).

  2. (ii)

    Consider the equation \(\sum _{i=1}^{n}\mathscr {C}(\delta _{i}, \gamma _{i}, K,1)=F(K,1)-\eta K\), that determines the value of K. \(\mathscr {C}\) is a decreasing function of K when \(F(K,1)-\eta K\) is increasing. The result follows from the features of \(\mathscr {C}\), that increases with \(\delta _{i}\) but decreases with \(\gamma _{i}\).

Appendix 8: Proof of Proposition 5

By definition, K is the solution to the following equation:

$$\begin{aligned} F(K, {1})-\eta K=\sum _{i={1}}^{\zeta }c^{i}(K)+(n-\zeta )({ 1}-\lambda )w(K), \end{aligned}$$

or \({H}^{\ell }(K)={H}^{r}(K,\gamma _{i}, \delta _{i}, \lambda )\), for \(i={1},{2},\ldots , k\). The uniqueness of the stationary competitive equilibrium ensures that

$$\begin{aligned} D _{K}{H}^{\ell }\bigl (K\bigr )- D _{K}{H}^{r}\bigl (K, \gamma _{i}, \delta _{i}, \lambda \bigr )>{0}. \end{aligned}$$

Noting then that \({H}^{r}(K,\gamma _{i}, \delta _{i}, \lambda )\) is independent of \(\gamma _{i}\) and \(\delta _{i}\) for any \(i\leqslant \zeta \) establishes (i).

Result (ii) follows from \( D _{\delta _{i}}C(\delta _{i}, \gamma _{i}, R, w, \lambda )>{0}, D _{\gamma _{i}}C(\delta _{i}, \gamma _{i}, R, w, \lambda )<{0}\) from Lemma 1.

Result (iii) in turn follows from \( D _{\lambda }C(\delta _{i}, \gamma _{i}, R, w, \lambda )\leqslant {0}\) from Lemma 1 and \(D_{\lambda }(1-\lambda )w<0\).

Appendix 9: Proof of Corollary 1

The argument of the proof replicates that in Proposition 5(iii), adding the property that, with \(D^{2}v>0, D_{\lambda }C(\delta , \gamma , R, w,\lambda )<0\).

Appendix 10: Proof of Lemma 4

The consumer’s program is given by:

$$\begin{aligned} \max _{\{c_{t}^{i}\}}&\sum _{t={0}}^{+\infty } \bigl (\delta _{i}\bigr )^{t}\Bigl \{u\bigl (c_{t}^{i}\bigr )+\gamma _{i}\bigl [v\bigl (c_{t}^{i}\bigr )-v\bigl (c_{t}^{i*}\bigr )\bigr ]\Bigr \}\\ {s.t}\quad&a_{t+{1}}^{i}= R_{t}a_{t}^{i}+w_{t}-c_{t}^{i},\\&\lim _{t\rightarrow +\infty }a_{t}^{i}\Bigm /\prod _{\tau ={0}}^{t}R_{\tau }\geqslant {0},\\&c_{t}^{i}\leqslant c_{t}^{i*}, \quad a_{{0}}^{i}\ \text{ given }. \end{aligned}$$

The optimality conditions are obtained by first listing the Lagrangian:

$$\begin{aligned} {L}_{t}&=\bigl (\delta _{i}\bigr )^{t}\Bigl \{u\bigl (c_{t}^{i}\bigr ) +\gamma _{i}\bigl [v\bigl (c_{t}^{i}\bigr )-v\bigl (c_{t}^{i*}\bigr )\bigr ]\Bigr \} +\mu _{t+{1}}^{i}\bigl [R_{t}a_{t}^{i}+w_{t}-c_{t}^{i}\bigr ]\\&\quad -\mu _{t}^{i}a_{t}^{i}+\nu _{t}^{i}\bigl (c_{t}^{i*}-c_{t}^{i}\bigr ), \end{aligned}$$

From which:

$$\begin{aligned}&\bigl (\delta _{i}\bigr )^{t}\bigl [ D u\bigl (c_{t}^{i}\bigr )+\gamma _{i} D v\bigl (c_{t}^{i}\bigr )\bigr ]=\mu _{t+{1}}^{i}+\nu _{t}^{i},\\&\mu _{t+{1}}^{i}=\mu _{t}^{i}/R_{t},\\&\lim _{t\rightarrow +\infty }\mu _{t}^{i}a_{t}^{i}={0}. \end{aligned}$$

At the same time, the optimal behavior of the competitive firm yields the equilibrium values of real wages and the rental rate for the capital stock as \(w_{t}= D _{ {L}}F(K_{t},{1})/n\) and \(R_{t}= D _{K}F(K_{t},{1})+{1}-\eta \). The equilibrium on the capital market giving \(\sum _{i={1}}^{n}a_{t}^{i}=K_{t}\) for any \(t\geqslant {0}\) produces the characterization.

Appendix 11: Proof of Proposition 6

The argument proceeds by proving that the sequence \(\{c_{t}^{i*}, K_{t}^{*}\}, i={1},\ldots , n\), satisfies the whole set of conditions that characterize a competitive equilibrium with constraints upon consumption for appropriately chosen values of \(\mu _{t}^{i}\) and \(\nu _{t}^{i}\) and for a suitable distribution of initial wealth \(a_{{0}}^{i}\). Let \(\mu _{t+ { 1}}^{i}=\chi _{t+{1}}/\zeta _{i}\) and \(\nu _{t}^{i}=(\delta _{i})^{t}\gamma _{i} D v(c_{t}^{i})\). Defining then the prices of the factors as \(w_{t}^{*}= D _{ {L}}F(K_{t}^{*},{1})/n\) and \(R_{t}^{*}= D _{K}F(K_{t}^{*},{ 1})+{1}-\eta \), the initial wealth of agents is selected so as to satisfy:

$$\begin{aligned} a_{{0}}^{i}=\sum _{t= { 0}}^{+\infty }\dfrac{c_{t}^{i*}}{\prod _{\tau = { 0}}^{t}R_{\tau }^{*}}-\sum _{t= { 0}}^{+\infty }\dfrac{w_{t}^{*}}{\prod _{\tau = {0}}^{t}R_{\tau }^{*}}, \end{aligned}$$

which is possible since

$$\begin{aligned} \sum _{i={1}}^{n}a_{{0}}^{i}&=\sum _{t={0}}^{+\infty }\dfrac{\sum _{i={1}}^{n}c_{t}^{i*}-nw_{t}^{*}}{\Pi _{\tau = {0}}^{t}R_{\tau }^{*}}\\&=\sum _{t={0}}^{+\infty }\dfrac{F\left( K_{t}^{*},{1}\right) -K_{t+{1}}^{*}+({1}-\eta )K_{t}^{*}-F_{ {L}}\left( K_{t}^{*},{1}\right) }{\Pi _{\tau = {0}}^{t}R_{\tau }^{*}}\\&=\sum _{t={0}}^{+\infty }\dfrac{R_{t}^{*}K_{t}^{*}-K_{t+{1}}^{*}}{\Pi _{\tau = {0}}^{t}R_{\tau }^{*}}\\&=K_{{0}}^{*}. \end{aligned}$$

The sequence \(\bigl \{a_{t}^{i}\bigr \}\) is then recursively defined by

$$\begin{aligned} a_{t+{1}}^{i}= R_{t}a_{t}^{i}+w_{t}-c_{t}^{i*}. \end{aligned}$$

Finally, and by the definition of \(a_{{0}}^{i}\), we have:

$$\begin{aligned} \lim _{t\rightarrow +\infty }a_{t}^{i}\Bigm /\prod _{\tau ={0}}^{t}R_{\tau }={0},\\ \end{aligned}$$

which establishes Proposition 6.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Drugeon, JP., Wigniolle, B. On impatience, temptation and Ramsey’s conjecture. Econ Theory 63, 73–98 (2017). https://doi.org/10.1007/s00199-015-0933-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-015-0933-4

Keywords

JEL Classification

Navigation