Skip to main content
Log in

Estimating a model of inefficient cooperation and consumption in collective households

  • Published:
Review of Economics of the Household Aims and scope Submit manuscript

Abstract

Lewbel and Pendakur (2021) propose a model of consumption inefficiency in collective households, based on “cooperation factors". We simplify that model to make it empirically tractable, and apply it to identify and estimate household member resource shares, and to measure the dollar cost of inefficient levels of cooperation. Using data from Bangladesh, we find that increased cooperation among household members yields the equivalent of a 13% gain in total expenditures, with most of the benefit of this gain going towards men.

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. Some results in DLP go through if these goods are only assignable but not private. So, e.g., when food is the assignable good, it could still have a coefficient in the A matrix that doesn’t equal one (and so technically isn’t private). This could arise if, e.g., food waste is lower in larger households. For simplicity, we follow DLP, but our results could also be generalized to allow the assignable good to be non-private. See Lechene et al. (2022). This would mainly entail extra notation, and adding some restrictions to Assumptions A5 and A6 in the Appendix.

  2. In practice, the private assignable goods may have the same price for each member, making π1 = . . . = πJ. For example, the private assignable good could be rice if we observed how much rice each household member eats, and rice has the same market price for all household members. As with DLP, some of the formal assumptions of our model will be easier to satisfy when the private assignable goods all have the same price.

  3. Equation (8) also implies restrictions on Af relative to the range of possible vectors p. These restrictions are comparable to those imposed by other empirical consumer demand models. See Lewbel and Pendakur (2008) and the Appendix for details.

  4. We have food consumption for each household member, and so could in theory estimate resource shares for each, rather than for total men, total women, and total children. However, that would then require estimating a separate model for every possible household composition, e.g., a separate model for households with 2 children vs those with 3.

  5. Module O1 (Food Consumption) and Module X2 (Intra-Household Food Distribution) actually collect food quantities and intakes, respectively, in nearly 300 categories. We aggregate these to 7 higher-level categories to make more sensible unit-values (described below). Module O1 gathers information from the female enumerator (who responds to most of the survey instrument); Module X2 gathers information from the female responsible for cooking that day. From Module X2, we use the weight of ingredients, rather than cooked weights, in our aggregation procedure.

  6. These are: jakat, fitra, daan, sodka, kurbani, milad, and other religious offerings.

  7. A previous version of this paper included an indicator of domestic abuse as a cooperation factor and log-wealth as a regressor. In Appendix B Table 6, we include these variables in the covariate list z. Their inclusion does not affect our major conclusions.

  8. This is reassuring for previous applications of similar models like DLP that don’t allow for inefficiency, suggesting that those models will still do a good a job of estimating resource shares, even if they miss the effects of inefficiency.

  9. This is a very strong dependence on household size, but well within the bounds allowable by the model. Specifically, BCL implies Barten scales between 1/n and 1, which we can use to calculate an approximate maximum value for δ of \(\frac{1}{2}\ln n\).

  10. Most more recent alternatives, like so-called “rank three" demand systems, are used for data from countries where the distribution of y is large, and more complicated budget responses are needed to capture behavior at both low and high income levels. Other popular demand models, like the multinomial logit based models widely used in the industrial organization literature, are designed for use with discrete demand data and are unsuitable for the type of continuous consumer demand data we analyze here.

  11. For example, one way Af drops out is if Af is block diagonal, with one block that does not vary by f, and with sj only depending on πj and the prices in that block. Alternatively, linear constraints could be imposed on the elements of Af, with sj depending only on the corresponding functions of prices, that, by these constraints, do not vary with Af. Analogous restrictions are often imposed on demand systems. For example as shown in Lewbel (1991), the Translog demand system as implemented by Jorgenson, and Slesnick (1987) imposes a linear constraint on its Barten (1964) scales, that results in a restriction like this on its equivalence scales. Note that BCL refer to the diagonal elements of Af as Barten technology parameters, due to their equivalence to Barten scales.

References

  • Abraham, A., & Laczo, S. (2017). Efficient risk sharing with limited commitment and storage. The Review of Economic Studies, 85(3), 1389–1424.

    Article  Google Scholar 

  • Apps, P.F. & Rees, R. (1988). Taxation and the Household. Journal of Public Economics, 35(3), 355–369.

    Article  Google Scholar 

  • Ashraf, N. (2009). Spousal control and intra-household decision making: an experimental study in the Philippines. American Economic Review, 99(4), 1245–1277.

    Article  Google Scholar 

  • Banks, J., Blundell, R., & Lewbel, A. (1997). Quadratic engel curves and consumer demand. Review of Economics and Statistics, 79(4), 527–539.

    Article  Google Scholar 

  • Barten, A. P. (1964). Family composition, prices, and expenditure patterns. In P. Hart, G. Mills & J. D. Whitaker (Eds.), Econometric Analysis for National Economic Planning: 16th Symposium of the Colston Society (277–292). Butterworth.

  • Basu, K. (2006). Gender and say: a model of household behaviour with endogenously determined balance of power. The Economic Journal, 116, 558–580.

    Article  Google Scholar 

  • Becker, G. S. (1965). A theory of the allocation of time. The Economic Journal, 75(299), 493–517.

    Article  Google Scholar 

  • Becker, G. S. (1981). A treatise on the family. Harvard University Press.

  • Bloch, F., & Rao, V. (2002). Terror as a bargaining instrument: a case study of dowry violence in rural India. American Economic Review, 92(4), 1029–1043.

    Article  Google Scholar 

  • Blundell, R., Chiappori, P.-A., & Meghir, C. (2005). Collective labor supply with children. Journal of Political Economy, 113(6), 1277–1306.

    Article  Google Scholar 

  • Brown, C., Calvi, R. & Penglase, J. (2021). Sharing the pie: An analysis of undernutrition and individual consumption in Bangladesh. Journal of Public Economics, 200, 104460.

    Article  Google Scholar 

  • Browning, M., & Chiappori, P.-A. (1998). Efficient intra-household allocations: a general characterization and empirical tests. Econometrica, 66(6), 1241–1278.

    Article  Google Scholar 

  • Browning, M., Chiappori, P.-A., & Lewbel, A. (2013). Estimating consumption economies of scale, adult equivalence scales, and household bargaining power. Review of Economic Studies, 80(4), 1267–1303.

    Article  Google Scholar 

  • Calvi, R. (2019). Why are older women missing in India? The age profile of bargaining power and poverty. Journal of Political Economy, 128(7), 2453–2501.

    Article  Google Scholar 

  • Castilla, C., & Walker, T. (2013). Is ignorance bliss? The effect of asymmetric information between spouses on intra-household allocations. American Economic Review: Papers & Proceedings, 103(3), 263–268.

    Article  Google Scholar 

  • Cherchye, L., De Rock, B., Lewbel, A. & Vermeulen, F. (2015). Sharing rule identification for general collective consumption models. Econometrica, 83(5), 2001–2041.

    Article  Google Scholar 

  • Chiappori, P.-A. (1988). Nash-bargained households decisions: a comment. International Economic Review, 29, 791–796.

    Article  Google Scholar 

  • Chiappori, P.-A. (1992). Collective labor supply and welfare. Journal of Political Economy, 100(3), 437–467.

    Article  Google Scholar 

  • Chiappori, P.-A., & Mazzocco, M. (2017). Static and intertemporal household decisions. Journal of Economic Literature, 55(3), 985–1045.

    Article  Google Scholar 

  • Christensen, L.R., Jorgenson, D.W. & Lau, L.J. (1975). Transcendental logarithmic utility functions. The American Economic Review, 65(3), 367–383.

    Google Scholar 

  • Deaton, A., & Muellbauer, J. (1980). An almost ideal demand system. American Economic Review, 70(3), 312–326.

    Google Scholar 

  • Deaton, A. (1997). The analysis of household surveys: a microeconometric approach to development policy. World Bank Publications.

  • Doepke, M., & Kindermann, F. (2019). Bargaining over babies: theory, evidence, and policy implications. American Economic Review, 109(9), 3264–3306.

    Article  Google Scholar 

  • Dunbar, G., Lewbel, A., & Pendakur, K. (2013). Children’s resources in collective households: identification, estimation and an application to child poverty in Malawi. American Economic Review, 103, 438–471.

    Article  Google Scholar 

  • Eswaran, M., & Malhotra, N. (2011). Domestic violence and women’s autonomy in developing countries: theory and evidence. Canadian Journal of Economics, 44(4), 1222–1263.

    Article  Google Scholar 

  • Gobbi, P. E. (2018). Childcare and commitment within households. Journal of Economic Theory, 176(C), 503–551.

    Article  Google Scholar 

  • Gutierrez, F.H. (2018). The intra-household allocation of resources in the presence of common-pool goods: theory and application to the measurement of indifference scales. Vanderbilt University unpublished manuscript.

  • Hansen, L. P. (1982). Large sample properties of generalized method of moments estimators. Econometrica: Journal of the Econometric Society, 50, 1029–1054.

  • Hidrobo, M., Peterman, A., & Heise, L. (2016). The effect of cash, vouchers, and food transfers on intimate partner violence: evidence from a randomized experiment in Northern Ecuador. American Economic Journal: Applied Economics, 8(3), 284–303.

    Google Scholar 

  • Hughes, C., Bolis, M., Fries, R., & Finigan, S. (2015). Women’s economic inequality and domestic violence: exploring the links and empowering women. Gender and Development, 23(2), 279–297.

    Article  Google Scholar 

  • Jorgenson, D. W., & Slesnick, D. T. (1987). Aggregate consumer behavior and household equivalence scales. Journal of Business & Economic Statistics, 5(2), 219–232.

  • Iyigun, M., & Walsh, R. P. (2007). Endogenous gender power, household labor supply and the demographic transition. Journal of Development Economics, 82, 138–155.

    Article  Google Scholar 

  • Koç, A. & Erkin, C. (2011). Development, women’s resources and domestic violence. Globalization, Religion and Development, 129–148.

  • Lechene, V., Pendakur, K. & Wolf, A. (2022). Ordinary least squares estimation of the intrahousehold distribution of expenditure. Journal of Political Economy, 130(3), 681–731.

    Article  Google Scholar 

  • Lewbel, A., & Pendakur, K. (2008). Estimation of collective household models with engel curves. Journal of Econometrics, 147, 350–358.

    Article  Google Scholar 

  • Lewbel, A. & Lin, X. (2022). Identification of semiparametric model coefficients, with an application to collective households. Journal of Econometrics, 226(2), 205–223.

    Article  Google Scholar 

  • Lewbel, A. & Pendakur, K. (2022). Inefficient collective households: cooperation and consumption. The Economic Journal, 132(645), 1882–1893.

    Article  Google Scholar 

  • Lewbel, A. (1991). The rank of demand systems: theory and nonparametric estimation. Econometrica: Journal of the Econometric Society, pp. 711–730.

  • Lise, J., & Seitz, S. (2011). Consumption inequality and intra-household allocations. Review of Economic Studies (2011), 78, 328–355.

    Article  Google Scholar 

  • Lise, J., & Yamada, K. (2019). Household sharing and commitment: evidence from panel data on individual expenditures and time use. The Review of Economic Studies, 86(5), 2184–2219.

    Article  Google Scholar 

  • Lundberg, S. & Pollak, R.A. (1993). Separate spheres bargaining and the marriage market. Journal of political Economy, 101(6), 988–1010.

    Article  Google Scholar 

  • Lundberg, S. & Pollak, R.A. (2003). Efficiency in marriage. Review of Economics of the Household, 1, 153–167.

    Article  Google Scholar 

  • Mazzocco, M. (2007). Household intertemporal behaviour: a collective characterization and a test of commitment. The Review of Economic Studies, 74(3), 857–895.

    Article  Google Scholar 

  • Menon, M., Pendakur, K., & Perali, F. (2012). On the expenditure-dependence of resource shares. Economics Letters, 117, 739–742.

    Article  Google Scholar 

  • Muellbauer, J. (1974). Household composition, Engel curves and welfare comparisons between households: A duality approach. European Economic Review, 5(2), 103–122.

    Article  Google Scholar 

  • Ramos, A. (2016). Household decision making with violence: implications for transfer programs. Universitat Autonoma de Barcelona.

  • Schaner, S. (2015). Do opposites detract? Intrahousehold preference heterogeneity and inefficient strategic savings. American Economic Journal: Applied Economics, 7(2), 135–174.

    Google Scholar 

  • Schaner, S. (2017). The cost of convenience? Transaction costs, bargaining power, and savings account use in Kenya. Journal of Human Resources, 52(4), 919–945.

    Article  Google Scholar 

  • Vogler, C. & Pahl, J. (1994). Money, power and inequality within marriage. The Sociological Review, 42(2), 263–288.

    Article  Google Scholar 

Download references

Acknowledgements

Pendakur acknowledges the financial support of the Social Sciences and Humanities Research Council of Canada. We are also grateful for the valuable research assistance provided by Shirleen Manzur, and helpful comment and feedback from our colleagues: Dave Freeman, Valerie Lechene, and Chris Muris.

Author contributions

A.L. and K.P. developed the model. Lewbel proved the theorems. Pendakur did the data work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Krishna Pendakur.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix

See Table 68

Formal assumptions and proofs

Here we formally derive our model, and prove that it is semiparametrically point identified. To simplify the derivations and assumptions, we first prove results without unobserved random utility parameters (as would apply if, e.g., our data consisted of many observations of a single household, or of many households with no unobserved variation in tastes). We then later add unobserved error terms to the model, corresponding to unobserved preference heterogeneity.

Let f, r, y, p, π, and z be as defined in the main text. Note that the first few Lemmas below will not impose the restriction that f only equal two values.

Assumption 1

Conditional on f, r, y, p, π, and z, the household chooses quantities to consume using the program given by equation (6) in the main text.

Assumption A1 describes the collective household’s conditionally efficient behavior. For each household member j, Uj is that member’s utility function over consumption goods, uj is that members additional utility or disutility associated with f, and ωj is that member’s Pareto weight.

As can be seen by Eq. (6) in the main text, the way that private assignable goods qj differ from other goods g is that each qj only appears in the utility function of individual j (which makes it assignable to that member) and these goods are unaffected by the matrix Af in the budget constraints, meaning that they are not shared or consumed jointly (which makes them private goods).

We next assume some regularity conditions. These assumptions ensure sensible and convenient restrictions on economic behavior like no money illusion, preferring larger consumption bundles to smaller ones, and the absence of corner solutions in the household’s maximization problem.

Assumption 2

Each \({\omega }_{j}\left(f,z,p,\pi ,y\right)\) function is differentiable and homogeneous of degree zero in \(\left(p,\pi ,y\right)\). Each \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) function is concave, strictly increasing, and twice continously differentiable in gj and qj. For each f, the matrix Af is nonsingular with all nonnegative elements and a strictly positive diagonal. The variable y and each element of p and π are all strictly positive, and the maximizing values of g1, q1, . . . gJ, qJ in Assumption A1 are all strictly positive.

Lemma 1

Let Assumptions A1 and A2 hold. Then there exist positive resource share functions \({\eta }_{j}\left(p,\pi ,y,f,z\right)\) such that \(\mathop{\sum }\nolimits_{j = 1}^{J}{\eta }_{j}\left(p,\pi ,y,f,z\right)=1\), and the household’s demand function for goods is given by each member j solving the program

$$\mathop{\max }\limits_{{g}_{j},{q}_{j}}{U}_{j}\left({q}_{j},{g}_{j},z\right)$$
(13)
$${{{\rm{such}}}}\,{{{\rm{that}}}}\,{p}^{{\prime} }{A}_{f}{g}_{j}+{\pi }_{j}{q}_{j}={\eta }_{j}\left(p,\pi ,y,z,f\right)y\,{{{\rm{and}}}}\,g={A}_{f}\mathop{\sum }\limits_{j=1}^{J}{g}_{j}.$$

To prove Lemma 1, first observe that the values of g1, q1, . . . gJ, qJ that maximize Eq. (6) in the main text are equivalent to the values that maximize

$$\mathop{\max }\limits_{{g}_{1},{q}_{1},...{g}_{J},{q}_{J}}\mathop{\sum }\limits_{j=1}^{J}{U}_{j}\left({q}_{j},{g}_{j},z\right){\omega }_{j}\left(p,\pi ,y,f\right)$$
(14)

given the same budget constraint. because the terms in Eq. (6) in the main text that are not in (14) do not depend on g1, q1, . . . gJ, qJ. With that replacement, the proof of Lemma 1 then follows immediately from the results derived in BCL. BCL only considered J = 2, but the extension of this Lemma to more than two household members, and to carrying the additional covariates, is straightforward. Note that the resource share functions ηj in Lemma 1 do not depend on r, because r, including the component v, does not appear in either Eq. (14) or in the budget constraint, and so cannot affect the outcome quantities.

Our empirical work will make use of cross section data, where price variation is not observed. Most of the remaining assumptions we make about resource shares and about the Uj component of utility are the same, or similar, to those made by DLP, and for the same reason: to ensure identification of the model without requiring price variation.

Assumption 3

The resource share functions \({\eta }_{j}\left(p,\pi ,y,f,z\right)\) do not depend on y.

DLP give many arguments, both theoretical and empirical, supporting the assumption that resource shares do not vary with y. Given Assumption A3, we hereafter write the resource share function as \({\eta }_{j}\left(\pi ,p,f,z\right)\).

For the next assumption, recall that an indirect utility function is defined as the function of prices and the budget that is obtained when one substitutes an individual’s demand functions into their direct utility function.

Assumption 4

For each household member j, the direct utility function \({U}_{j}\left({g}_{j},{q}_{j},z\right)\), when facing prices p and π and having the budget y, has the associated indirect utility function

$${V}_{j}\left({\pi }_{j},p,y,z\right)=\left[\ln y-\ln {S}_{j}\left({\pi }_{j},p,z\right)\right]{M}_{j}\left({\pi }_{j},p,z\right)$$
(15)

For some functions Sj and Mj.

Assumption A4 says that household members each have utility functions in the class that Muellbauer (1974) called PIGLOG (price independent, generalized logarithmic) preferences. As noted in the main text, this is a class of functional forms that is widely known to fit empirical continuous consumer demand data well. Examples of popular models in this class include the Christensen et al. (1975) Translog demand system and Deaton and Muellbauer’s (1980) AIDS (Almost Ideal Demand System) model.Footnote 10

Lemma 2

Let Assumptions A1, A2, A3, and A4 hold. Then the value of \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) attained by household member j is given by

$${U}_{j}=\left[\ln {\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)+\ln y-\ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)\right]{M}_{j}\left({\pi }_{j},{A}_{f}p,z\right)$$
(16)

and the household’s demand functions for the private assignable goods qj are

$${q}_{j}={\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)y\left(\frac{\partial \ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)}{\partial {\pi }_{j}}-\frac{\partial \ln {M}_{j}\left({\pi }_{j},{A}_{f}p,z\right)}{\partial {\pi }_{j}}\ln \left(\frac{{\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)y}{{S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)}\right)\right)$$
(17)

To prove Lemma 2, observe that by Lemma 1, household member j maximizes the utility function \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) facing shadow prices \({A}_{f}^{{\prime} }p\) and πj and having the shadow budget \({\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)y\). Therefore, using the definition of indirect utility, member j’s attained utility level \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) is given by \({V}_{j}\left({\pi }_{j},{A}_{f}^{{\prime} }p,{\eta }_{j}\left(\pi ,{A}_{f}p,f\right)y\right)\), which by Assumption A4 equals Eq. (16). Next, a property of regular indirect utility functions is that the corresponding demand functions can be obtained by Roy’s identity. Equation (17) is obtained by applying Roy’s identity to Eq. (15) for the private assignable goods qj, and then replacing p and y in the result with \({A}_{f}^{{\prime} }p\) and \({\eta }_{j}\left(\pi ,{A}_{f}p,f\right)y\).

We could similarly obtain the demand functions for other goods g, as in BCL, but these will be more complicated due to the sharing, with Roy’s identity being applied to each member to obtain each gj demand function, and substituting the results into \(g={A}_{f}\mathop{\sum }\nolimits_{j = 1}^{J}{g}_{j}\). However, our empirical analyses will only make use of the private assignable goods qj with demands given by Eq. (17).

Assumption 5

Let \(\ln {M}_{j}\left({\pi }_{j},{A}_{f}p,z\right)={m}_{j}\left({A}_{f}p,z\right)-\beta \left(z\right)\ln {\pi }_{j}\) for some functions mj and β.

There are two restrictions embodied in Assumption A5. One is that the functional form of \(\ln {M}_{j}\) in terms of prices is linear and additive in \(\ln {\pi }_{j}\), and the other is that the function \(\beta \left(z\right)\) does not vary by j. The functional form restriction of log linearity in log prices is a common one in consumer demand models, e.g., the function Mj in Deaton and Muellbauer’s (1980) AIDS (Almost Ideal Demand System) satisfies this restriction. Assumption A5 could be further relaxed by letting β depend on p (though not on Af) without affecting later results.

To identify their model, DLP define and use a property of preferences called similarity across people (SAP), and provide empirical evidence in support of SAP. The restriction that β not vary by j suffices to make SAP hold for the private assignable goods (but not necessarily for other goods).

Assumption 6

Let \(\ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)=\ln {s}_{j}\left({\pi }_{j},p,z\right)-\ln \delta \left({A}_{f}p,z\right)\) for some functions sj and δ. Without loss of generality, let \(\ln \delta \left({A}_{0}p,z\right)=0\).

Assumption A6 assumes separability of the effects of πj and f on the function Sj. DLP discuss various ways in which the matrix Af can drop out of a function of prices, as required in the function sj.Footnote 11 This assumption is not vital, but will be helpful for making the cost of an inefficient choice of f identifiable. Assuming \(\ln \delta \left({A}_{0}p,z\right)=0\) in Assumption A6 is without loss of generality, because if it does not hold then one can make it hold if one redefines δ and sj by subtracting \(\ln \delta \left({A}_{0}p,z\right)\) from both \(\ln \delta \left(f,p,z\right)\) and \(\ln {s}_{j}\left({\pi }_{j},p,z\right)\).

It will be convenient to express our demand functions in budget share form. Define wj = qjπj/y. This budget share is the fraction of the household’s budget y that is spent on buying person j’s assignable good qj.

Lemma 3

Given Assumptions A1 to A6, the value of \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) attained by household member j is given by

$$\left[\ln {\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)+\ln y-\ln {s}_{j}\left({\pi }_{j},p,z\right)+\ln \delta \left({A}_{f}p,z\right)\right]\left[{m}_{j}\left({A}_{f}p,z\right)-\beta \left(z\right)\ln {\pi }_{j}\right]$$
(18)

and the budget share demand functions for each private assignable good are given by

$${w}_{j}={\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)\left[{\gamma }_{j}\left({\pi }_{j},p,z\right)+\beta \left(z\right)\left(\ln y+\ln {\eta }_{j}\left(\pi ,{A}_{f}p,f,z\right)+\ln \delta \left({A}_{f}p,z\right)\right)\right].$$
(19)

where the function γj is defined by

$${\gamma }_{j}\left({\pi }_{j},p,z\right)=\frac{\partial \ln {s}_{j}\left({\pi }_{j},p,z\right)}{\partial \ln {\pi }_{j}}-\beta \left(z\right)\ln {s}_{j}\left({\pi }_{j},p,z\right)$$

The proof of Lemma 3 consists of substituting the expressions for Mj and Sj given by Assumptions A5 and A6 into the equations given by Lemma 2, and converting the quantity qj into the budget share wj.

Assumption 7

Market prices p and π are the same for all households.

Our data come from a single time period, which (assuming the law of one price) justifies assuming p and π are the same across all households. This assumption makes our demand functions reduce to Engel curves. For simplicity, we abuse notation here and redefine objects that were functions of Afp as just functions of f, since with fixed prices the only source of variation of Afp is just variation in f).

Lemma 4

Given Assumptions A1 to A7, the value of \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) attained by household member j is given by

$$\left[\ln {\eta }_{j}\left(f,z\right)+\ln y-\ln {s}_{j}\left(z\right)+\ln \delta \left(f,z\right)\right]{M}_{j}\left(f,z\right)$$
(20)

and the budget share Engel curve functions \({w}_{j}={W}_{j}\left(f,z,y\right)\) for each private assignable good are given by

$${W}_{j}\left(f,z,y\right)={\eta }_{j}\left(f,z\right)\left[{\gamma }_{j}\left(z\right)+\beta \left(z\right)\left(\ln y+\ln {\eta }_{j}\left(f,z\right)+\ln \delta \left(f,z\right)\right)\right].$$
(21)

Lemma 4 entails a small abuse of notation, where we have absorbed the values of p and π into the definitions of all of our functions, noting that any function of Afp remains a function of f even if p is a constant. Lemma 4 is just rewriting Lemma 3 after dropping the prices.

Lemma 5

Let Assumptions A1 to A7 hold. Let \({W}_{j}\left(f,z,y\right)\) be defined by Eq. (21) for j = 1, . . . , J. Given functions \({W}_{j}\left(f,z,y\right)\), the functions \({\eta }_{j}\left(f,z\right)\), \(\delta \left(f,z\right)\),\({\gamma }_{j}\left(z\right)\), and \(\beta \left(z\right)\) are all point identified.

To prove Lemma 5, observe first by Eq. (21) that \({\eta }_{j}\left(f,z\right)\beta \left(z\right)=\partial {W}_{j}\left(f,z,y\right)/\partial \ln y\). Next, since resource shares sum to one, we can identify \(\beta \left(z\right)\) and \({\eta }_{j}\left(f,z\right)\) by

$$\beta \left(z\right)=\mathop{\sum }\limits_{j=1}^{J}\frac{\partial {W}_{j}\left(f,z,y\right)}{\partial \ln y}\,{{{\rm{and}}}}\,{\eta }_{j}\left(f,z\right)=\frac{1}{\beta \left(z\right)}\frac{\partial {W}_{j}\left(f,z,y\right)}{\partial \ln y}$$

Next, define \({\rho }_{j}\left(f,z,y\right)\) by

$${\rho }_{j}\left(f,z,y\right)=\frac{{W}_{j}\left(f,z,y\right)}{{\eta }_{j}\left(f,z\right)}-\beta \left(z\right)\left(\ln y+\ln {\eta }_{j}\left(f,z\right)\right)$$

The function \({\rho }_{j}\left(f,z,y\right)\) is identified because it is defined entirely in terms of identified functions. By Eq. (21), \({\rho }_{j}\left(f,z,y\right)={\gamma }_{j}\left(z\right)-\beta \left(z\right)\ln \delta \left(f,z\right)\). It follows from Assumption A6 that \(\ln \delta \left(0,z\right)=0\), so \({\gamma }_{j}\left(z\right)\) and \(\delta \left(f,z\right)\) are identified by

$${\gamma }_{j}\left(z\right)={\rho }_{j}\left(0,z,y\right)\,{{{\rm{and}}}}\,\ln \delta \left(f,z\right)=\frac{{\rho }_{j}\left(f,z,y\right)-{\rho }_{j}\left(0,z,y\right)}{\beta \left(z\right)}$$

evaluated at any value of y (or, e.g., averaged over y).

Lemma 5 shows that, given the household demand functions, the resource share functions \({\eta }_{j}\left(f,z\right)\) are identified, so our model, like DLP, overcomes the problem in the earlier collective household literature of (the levels of) resource shares not being identified. Lemma 5 also shows identification of the preference related functions \({\gamma }_{j}\left(z\right)\) and \(\beta \left(z\right)\), and identification of our new cost of inefficiency function \(\delta \left(f,z\right)\).

Lemma 6

Let Assumptions A1 to A7 hold. Assume f is determined by maximizing \({{\Psi }}\left({U}_{1}+{u}_{1},...,{U}_{J}+{u}_{J}\right)\) for some function Ψ. Then \(f=\arg \max {{\Psi }}\left({R}_{1}\left(p,y,f,v\right),...{R}_{J}\left(p,y,f,v\right)\right)\) where \({R}_{j}\left(f,y,v,z\right)\) is given by

$${R}_{j}\left(f,y,v,z\right)=\left(\ln {\eta }_{j}\left(f,z\right)+\ln y-\ln {s}_{j}\left(z\right)+\ln \delta \left(f,z\right)\right){M}_{j}\left(f,z\right)+{u}_{j}\left(f,v,z\right)$$

The proof of Lemma 6 is then that, by Eq. (20) and the definition of uj, for any f the level of Uj + uj attained by member j is given by the function \({R}_{j}\left(f,y,v,z\right)\).

The above analyses apply to a single household. Our data will actually consist of a cross section of households, each only observed once. To allow for unobserved variation in tastes across households in a conveniently tractible form, replace the function \(\ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)\) with \(\ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)-{\widetilde{\varepsilon }}_{j}\) where \({\widetilde{\varepsilon }}_{j}\) is a random utility parameter representing unobserved variation in preferences for goods. This means that \({\widetilde{\varepsilon }}_{j}\) appears in member j’s utility function Uj. We assume these taste parameters vary randomly across households, so \(E\left({\widetilde{\varepsilon }}_{j}| r,z\right)=0\). Similarly, replace \({u}_{j}\left(f,r,z\right)\) with \({u}_{j}\left(f,r,z\right)+{\widetilde{e}}_{jf}\) where \({\widetilde{e}}_{jf}\) represents variation in the utility or disutility associated with the choice of f. The errors \({\widetilde{e}}_{jf}\) and \({\widetilde{\varepsilon }}_{j}\) can be correlated with each other and across household members.

Substituting these definitions into the above equations, we get

$${w}_{j}={\eta }_{j}\left(f,z\right)\left[{\gamma }_{j}\left(z\right)+\beta \left(z\right)\left(\ln y+\ln {\eta }_{j}\left(f,z\right)+\ln \delta \left(f,z\right)\right)+{\varepsilon }_{j}\right]$$
(22)

where \({\varepsilon }_{j}=\beta \left(z\right)\,{\widetilde{\varepsilon }}_{j}\) so \(E\left({\varepsilon }_{j}| r,z\right)=0\), and f is now determined by

$$f=\arg \max {{\Psi }}\left({\widetilde{R}}_{1f},...{\widetilde{R}}_{Jf}\right),{{{\rm{where}}}}\,{\widetilde{R}}_{jf}={R}_{j}\left(f,y,r,z\right)+\left({M}_{j}\left(f,z\right)/\beta \left(z\right)\right){\varepsilon }_{j}+{\widetilde{e}}_{jf}$$
(23)

We will want to estimate the Engel curve Eq. (22) for j = 1, . . . , J. Equation (23) shows that f is an endogenous regressor in these equations, because f depends on both εj and \({\widetilde{e}}_{jf}\). As discussed in the main text, we do not try to empirically identify or estimate Eq. (23), because both the functions Rj and errors \({\widetilde{e}}_{1f}\) depend on uj, and there may be important determinents of uj (the direct utility or disutility from cooperation) that we cannot observe. However, we will require at least one instrument for f.

Another source of error in our model is that, in our data, y is a constructed variable (including imputations from home production), and so may suffer from measurement error. We will therefore require instruments for y. Our current collective household model is static. This is justified by a standard two stage budgeting (time separability) assumption, in which households first decide how much of their income and assets to save versus how much to spend in each time period, and then allocate their expenditures to the various goods they purchase. The total they spend in the time period is y, and the household’s allocation of y to the goods they purchase is given by Eq. (6) in the main text. These means that variables associated with household income and wealth will correlate with y and so are potential instruments for y.

This time separability applies to the utility functions over goods, \({U}_{j}\left({q}_{j},{g}_{j},z\right)\) for each member j, but need not apply to the utility or disutility associated with f, that is, \({u}_{j}\left(f,v,z\right)\). So at least some of these income and wealth variables could be components of v. Let \(\widetilde{r}\) denote a vector of potential instruments for y. These are measures related to income or wealth that are not already included in v.

Assume there exists values v0 and v1 such that \({u}_{j}\left(f,{v}_{0},z\right)\,\ne \,{u}_{j}\left(f,{v}_{1},z\right)\) for some member j who’s utility appears in Ψ. Then it follows from Eq. (23) that f varies with v, so v can serve as an instrument for f. Similarly, assume that \(\ln y\) correlates with \(\widetilde{r}\), which can serve as instruments for \(\ln y\) (elements of v could also be instruments for y). Based on Eq. (22), we then have conditional moments

$$E\left[\left(\frac{{w}_{j}}{{\eta }_{j}\left(f,z\right)}-{\gamma }_{j}\left(z\right)-\beta \left(z\right)\left(\ln y+\ln {\eta }_{j}\left(f,z\right)+\ln \delta \left(f,z\right)\right)\right)| \widetilde{r},v,z\right]=0$$
(24)

Later in this Appendix we consider nonparametric identification of the functions in this expression based on these moments, but for now consider using these moments parametrically. If we parameterize each of the unknown functions using a parameter vector θ, then Eq. (24) implies unconditional moments

$$E\left[\left(\frac{{w}_{j}}{{\eta }_{j}\left(f,z,\theta \right)}-{\gamma }_{j}\left(z,\theta \right)-\beta \left(z,\theta \right)\left(\ln y+\ln {\eta }_{j}\left(f,z,\theta \right)+\ln \delta \left(f,z,\theta \right)\right)\right)\phi\, \left(\widetilde{r},v,z\right)\right]=0$$
(25)

for any suitably bounded functions \(\phi \left(\widetilde{r},v,z\right)\). Our actual estimator will consist of parameterizing the unknown functions in this expression, choosing a set of functions \(\phi \left(\widetilde{r},v,z\right)\), and estimating the parameters by GMM (the generalized method of moments) based on these moments. At the end of this Appendix we discuss choice of the ϕ functions.

Equation (25) can suffice for parametric identification and estimation, but is it still possible to nonparametrically identify the functions in this model in the presence of unobserved heterogeneity? The following Theorem shows that the answer is yes, if we make some additional assumptions. Theorem 1 shows these additional assumptions are sufficient for nonparametric identification of these functions, These additional assumptions, which are not required for parametric identification, are listed in Assumption A8.

Assumption 8

Add unobservable heterogeneity terms \({\widetilde{\varepsilon }}_{j}\) and \({\widetilde{e}}_{jf}\) to the model by replacing the function \(\ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)\) with \(\ln {S}_{j}\left({\pi }_{j},{A}_{f}p,z\right)-{\widetilde{\varepsilon }}_{j}\) and \({u}_{j}\left(f,v,z\right)\) with \({u}_{j}\left(f,v,z\right)+{\widetilde{e}}_{jf}\), for j = 1, . . , J. Assume f is determined by maximizing Ψ, where Ψ is linear, so \({{\Psi }}\left({\widetilde{R}}_{1f},...{\widetilde{R}}_{Jf}\right)=\mathop{\sum }\nolimits_{j = 1}^{J}{\widetilde{c}}_{j}{\widetilde{R}}_{jf}\) for some constants \({\widetilde{c}}_{1}\),...,\({\widetilde{c}}_{J}\). Let \(\widetilde{e}=\mathop{\sum }\nolimits_{j = 1}^{J}{\widetilde{c}}_{j}\left({\widetilde{e}}_{j1}-{\widetilde{e}}_{j0}\right)\). Define \(\widetilde{y}\left(\widetilde{r},v,z\right)\) by \(\ln \widetilde{y}\left(\widetilde{r},v,z\right)=E\left(\ln y| \widetilde{r},v,z\right)\). Assume the following: The function \(\widetilde{y}\left(\widetilde{r},v,z\right)\) is differentiable in a scalar \(\widetilde{r}\) with a nonzero derivative. The error \(\widetilde{e}\) is independent of \(y,\widetilde{r},v,z\) and \(\left({\varepsilon }_{j},\widetilde{e}\right)\) is independent of \(\widetilde{r}\) conditional on \(\left(v,z\right)\). \(E\left({\varepsilon }_{j}| \widetilde{r},v,z\right)=0\). The functions \({M}_{j}\left(f,z\right)\) do not depend on f. There exist values v1 and v0 of v such that \(\mathop{\sum }\nolimits_{j = 1}^{J}{\widetilde{c}}_{j}{u}_{j}\left(f,{v}_{1},z\right)\ne \mathop{\sum }\nolimits_{j = 1}^{J}{\widetilde{c}}_{j}{u}_{j}\left(f,{v}_{0},z\right)\).

Theorem 1

Let Assumptions A1 to A8 hold. Then the functions \({\eta }_{j}\left(f,z\right)\), \(\delta \left(f,z\right)\),\({\gamma }_{j}\left(z\right)\), and \(\beta \left(z\right)\) are identified.

To prove Theorem 1, first observe that, with f binary, it follows from Eq. (23) that f = 1 if \(\mathop{\sum }\nolimits_{j = 1}^{J}{\widetilde{c}}_{j}\left[{R}_{j}\left(1,y,r,z\right)+\left({M}_{j}\left(1,z\right)/\beta \left(z\right)\right){\varepsilon }_{j}+{\widetilde{e}}_{j1}\right]\) is greater than

\(\mathop{\sum }\nolimits_{j = 1}^{J}{\widetilde{c}}_{j}\left[{R}_{j}\left(0,y,r,z\right)+\left({M}_{j}\left(0,z\right)/\beta \left(z\right)\right){\varepsilon }_{j}+{\widetilde{e}}_{j0}\right]\), where the function Rj is given by Lemma 6. Taking the difference in these expressions, and using the assumption that \({M}_{j}\left(f,z\right)\) doesn’t depend on f, we get that f = 1 if and only if

$$\begin{array}{ll}&\mathop{\sum }\limits_{j=1}^{J}{\widetilde{c}}_{j}\left[\left(\ln {\eta }_{j}\left(1,z\right)+\ln \delta \left(1,z\right)\right){M}_{j}\left(z\right)+{\mu }_{j}\left(1,v,z\right)\right.\\ &\left.-\left(\ln {\eta }_{j}\left(0,z\right)+\ln \delta \left(0,z\right)\right){M}_{j}\left(z\right)-{\mu }_{j}\left(0,v,z\right)\right]+\widetilde{e}\end{array}$$

is positive. This means that \(f=\widetilde{f}\left(v,z,\widetilde{e}\right)\) for some function \(\widetilde{f}\). More precisely, f obeys a threshold crossing model where f is one if a function of v and z given by the above expression is greater than \(-\widetilde{e}\), otherwise f is zero.

Now, again exploiting that f is binary,

$$E\left({w}_{j}| \widetilde{r},v,z,y\right)=E\left[{W}_{j}\left(f,z,y\right)+\beta \left(z\right)\ln \delta \left(f,z\right)\,{\widetilde{\varepsilon }}_{j}\,| \widetilde{r},v,z,y\right]$$
$$=E[{W}_{j}\left(1,z,y\right)f+\beta \left(z\right)\ln \delta \left(1,z\right)\,f{\widetilde{\varepsilon }}_{j}+{W}_{j}\left(0,z,y\right)\left(1-f\right)+\beta \left(z\right)\ln \delta \left(0,z\right)\left(1-f\right){\widetilde{\varepsilon }}_{j}| \widetilde{r},v,z,y]\,$$
$$\begin{array}{ll}={W}_{j}\left(0,z,y\right)+\left[{W}_{j}\left(1,z,y\right)-{W}_{j}\left(0,z,y\right)\right]E\left(f| \widetilde{r},v,z,y\right)\\ \quad+\,\beta \left(z\right)\left[\ln \delta \left(1,z\right)-\ln \delta \left(0,z\right)\right]E\left(f{\widetilde{\varepsilon }}_{j}| \widetilde{r},v,z,y\right).\end{array}$$

Next, observe that, since \({W}_{j}\left(f,z,y\right)\) is linear in \(\ln y\), \(E\left[{W}_{j}\left(0,z,y\right)| \widetilde{r},v,z\right]={W}_{j}\left(0,z,\widetilde{y}\right)\) and \(E\left[{W}_{j}\left(1,z,y\right)| \widetilde{r},v,z\right]={W}_{j}\left(1,z,\widetilde{y}\right)\) where \(\widetilde{y}=\widetilde{y}\left(\widetilde{r},v,z\right)\). Averaging the above expression over y, and noting that \(f=\widetilde{f}\left(v,z,{\widetilde{e}}_{1}\right)\), we get

$$\begin{array}{ll}E\left({w}_{j}| \widetilde{r},v,z\right)={W}_{j}\left(0,z,\widetilde{y}\right)+\left[{W}_{j}\left(1,z,\widetilde{y}\right)-{W}_{j}\left(0,z,\widetilde{y}\right)\right]E\left(f| \widetilde{r},v,z\right)\\ \qquad\qquad\qquad\quad+\,\beta \left(z\right)\left[\ln \delta \left(1,z\right)-\ln \delta \left(0,z\right)\right]E\left(f{\widetilde{\varepsilon }}_{j}| \widetilde{r},v,z\right).\end{array}$$

and by the conditional independence assumptions regarding \({\widetilde{\varepsilon }}_{j}\) and \({\widetilde{e}}_{1}\),

$$\begin{array}{rc}E\left({w}_{j}| \,\widetilde{r},v,z\right)&={W}_{j}\left(0,z,\widetilde{y}\right)+\left[{W}_{j}\left(1,z,\widetilde{y}\right)-{W}_{j}\left(0,z,\widetilde{y}\right)\right]E\left(f| v,z\right)\\ &+\beta \left(z\right)\left[\ln \delta \left(1,z\right)-\ln \delta \left(0,z\right)\right]E\left(f{\widetilde{\varepsilon }}_{j}| v,z\right).\end{array}$$

Now the functions \(E\left({w}_{j}| \widetilde{r},v,z\right)\) and \(\widetilde{y}\left(\widetilde{r},v,z\right)\) (the latter defined by \(\ln \widetilde{y}\left(\widetilde{r},v,z\right)=E\left(\ln y| \widetilde{r},v,z\right)\)) are both identified from data (and could, e.g., be consistently estimated by nonparametric regressions. So the derivatives of these expressions with respect to \(\widetilde{r}\) are identified. This means that the following expression is identified.

$$\frac{\partial E\left({w}_{j}| \widetilde{r},v,z\right)}{\partial \ln \widetilde{r}}/\frac{\partial \ln \widetilde{y}\left(\widetilde{r},v,z\right)}{\partial \ln \widetilde{r}}=\frac{\partial {W}_{j}\left(0,z,\widetilde{y}\right)}{\partial \ln \widetilde{y}}+\frac{\partial \left[{W}_{j}\left(1,z,\widetilde{y}\right)-{W}_{j}\left(0,z,\widetilde{y}\right)\right]}{\partial \ln \widetilde{y}}E\left(f| v,z\right)$$
(26)

Taking the difference between the above expression evaluated at v = v1 and at v = v0 then gives (and so identifies)

$$\frac{\partial \left[{W}_{j}\left(1,z,\widetilde{y}\right)-{W}_{j}\left(0,z,\widetilde{y}\right)\right]}{\partial \ln \widetilde{y}}\left[E\left(f| {v}_{1},z\right)-E\left(f| {v}_{0},z\right)\right]$$

and, since \(E\left(f| v,z\right)\) is also identified, this identifies \(\partial \left[{W}_{j}\left(1,z,\widetilde{y}\right)-{W}_{j}\left(0,z,\widetilde{y}\right)\right]/\partial \ln \widetilde{y}\). We can then solve Eq. (26) for \(\partial {W}_{j}\left(0,z,\widetilde{y}\right)/\partial \ln \widetilde{y}\) where all the terms defining this derivative are identified. Taken together, the last two steps identify \(\partial {W}_{j}\left(f,z,\widetilde{y}\right)/\partial \ln \widetilde{y}\) for f = 0 and for f = 1.

Given these identified functions and derivatives, we may then duplicate the proof of Lemma 5, (replacing y with \(\widetilde{y}\), to show that the functions \(\beta \left(z\right)\), \({\eta }_{j}\left(f,z\right)\), \({\gamma }_{j}\left(z\right)\), and \(\delta \left(f,z\right)\) are identified.

2.1 Instrument validity

To more formally define conditions under which village-level average f is a valid instrument, assume that the household h random utility parameters \({\widetilde{e}}_{1fh}\) and \({\widetilde{\varepsilon }}_{jh}\) defined in the Appendix are independent across households. Let \({\overline{f}}_{h}\) equal the expected value of fh conditional on being a household other than h in the village. Then \({\overline{f}}_{h}\) is the probability that a randomly chosen household in the village, other than household h, cooperates. Assume that we include \({\overline{f}}_{h}\) in the function Rj (equals Uj + uj, formally defined in the Appendix). Taking the conditional mean of Eq. (23) across households other than household h in the village then shows that \({\overline{f}}_{h}\) equals a function of the joint distribution of \({y}_{{h}^{{\prime} }}\), \({{{{\boldsymbol{r}}}}}_{{h}^{{\prime} }}\), \({{{{\boldsymbol{z}}}}}_{{h}^{{\prime} }}\), \({\widetilde{e}}_{1f{h}^{{\prime} }}\) and \({\widetilde{\varepsilon }}_{1{h}^{{\prime} }}\) across all households \({h}^{{\prime} }\) other than h in the village. It follows that \({\overline{f}}_{h}\) is a relevent instrument in that it affects the choice of f (by being in Rj) and that it is a valid instrument in the quantity demand equations because \({\overline{f}}_{h}\) is independent of household h’s specific value of \({\widetilde{\varepsilon }}_{jh}\) and hence of εjh.

2.2 Instruments

Our model has two endogenous regressors: the log of household total expenditures, \(\ln\)yh, and the cooperation factor fh. As discussed earlier, if we assume that the consumption allocation decision in our model is separable from the decision of how to allocate household income between total consumption and savings, then functions of household wealth are valid instruments for \(\ln\)yh. This time separability is a standard assumption in the consumer demand literature, including in collective household models (see, e.g., Lewbel & Pendakur, 2008). We discuss time separability formally in the Appendix. Another reason yh could potentially be endogenous is measurement error, stemming from, e.g., purchase mismeasurement, or infrequency of expenditures on some consumption items. Functions of wealth are also valid instruments for dealing with expenditure measurement issues (see, e.g., Banks et al., 1997).

Now consider instruments for fh. We do not attempt to specify and estimate this equation, so we need an instrument vh for fh. This instrument does not need to be randomly assigned, but it does need to correlate with the choice of fh, while not (after conditioning on other covariates) directly affecting the household’s food consumption decisions (in terms of the model, vh must appear in one or more of the uj functions, but not appear in the functions Uj and ωj for j = 1, . . . , J).

Our primary instrument for fh is the leave-one-out village level average value of f (the average excluding household h). The idea is that variation in the local prevalence of families whose members cooperate on consumption decisions is likely to correlate with an individual’s own decision to likewise cooperate. Roughly, village level average f (leaving out household h) is a valid instrument in our model if the choice of f in households other than household h is unrelated to the unobserved preference heterogeneity in member’s demand functions for food in household h. See the Appendix for a formal definition of conditions under which this instrument is is valid.

For estimation, we do not need to distinguish which elements of the instrument list rh are intended to be specifically instruments for fh vs for yh (i.e., elements of v vs elements of \(\widetilde{{{{\boldsymbol{r}}}}}\) in the Appendix). In particular, though we argue that \({\overline{f}}_{h}\) should primarily correlate with fh and wealth should primarily correlate with yh, either or both could affect both. Moreover, since we do not know the functional forms by which fh and yh depend on \({\overline{f}}_{h}\) and wealth, we let our instrument list rh consist of r1h and r2h, where r1h consists of the first through fourth powers of \({\overline{f}}_{h}\) and r2h consists of the first through fourth powers of log wealth. We use these powers to flexibly capture how fh and yh might depend on these instruments. Descriptive statistics for our instruments are given at the bottom of Table 1b.

If our model were linear, then our nonlinear GMM estimator would (apart from weighting matrix) reduce to a linear two stage least squares. The first stage of that two stage least squares would consist of regressing the endogenous f and \(\ln y\) on the instruments and exogenous regressors.

To assess the strength of our instruments, we ran those first stage linear regressions. In Table 2 we give regression estimates and associated standard errors from a linear regression of our endogenous regressors, fh and \(\ln {y}_{h}\) on our 18 demographic variables zh and our 8 instruments rh. Standard errors are clustered at the village (i.e., the Upazila) level. Table 2 shows that fh is difficult to predict, with an R2 of just 0.17, but the instruments collectively appear strong, in that the F-statistic for the relevance of the instruments (conditional on covariates) is 62. As expected, the village-level average instruments do most of the work here, with an F-statistic of 121, and the log-wealth instruments are also jointly insignificant in this equation. The low R2 of this regression emphasizes the point that we can’t (and don’t try to) actually model the decision to cooperate. All we need are sufficiently strong instruments, which our F-statistic indicates is the case (being, e.g., much larger than the rule of thumb level of 10).

Although we can’t treat this regression as a formal model of cooperation, it is suggestive regarding covariates. The regression shows that village level average cooperation is positively correlated with a household’s individual decision to cooperate fh, as expected. It is also positively correlated with the education of women and age of children, and negatively correlated with the age of women, suggesting that it may respond to women’s bargaining power.

The household log budget \(\ln {y}_{h}\) is fitted with an R2 of 0.43 and an F-statistic of the instruments of 101. Here, the log-wealth instruments do most of the work, with an F-statistic of 186. But, the cooperation instruments are also relevant in this equation, with an F-statistic of 9.

The above results provide evidence for the relevance our instruments. For further reassurance that the instruments are valid for our model, we later test the exogeneity of the instruments via overidentification tests.

Table 6 GMM estimates, varying covariates
Table 7 ▓

2.3 Discussion of Table 8

In Table 8, we consider three alternatives regarding data construction in our baseline model. In the leftmost column, labeled (7), we retain the previously dropped 238 households that had zero food intake for any member type (adult males, adult females or children). We dropped these households because they likely indicate measurement issues. However, if these zeroes result instead from infrequency that is correlated with regressors (e.g., if significant numbers of households are so poor that some members don’t eat every day), then excluding these households could lead to bias. In comparison with the baseline estimates, we see slightly smaller estimates of resource shares for men in the reference household type, and slightly larger estimates of resource shares for women and children. However, the estimated marginal effect of cooperation f on resource shares is very similar to the baseline estimates: cooperation increases male resource shares by about 3 percentage points, has roughly no effect on female resource shares and reduces children’s resource shares by 3 percentage points. The estimated efficiency gain from cooperation is also very similar to baseline, with \(\delta =\exp (0.135)\), or about 14 percent. The resulting money metric utility gains Δj from efficiency are 28 percent for men, 14 percent for women, and 5 percent for children, compared to the baseline estimates of 23, 11, and 6 percent.

Table 8 Estimated efficiency and resource shares, varying samples

In the middle panel, we consider some additional sample restrictions that may be sensible. In this model we exclude: a) households where the female respondent is not married; b) households in the top or bottom percentile of the distribution of budgets; and c) households that report zero wealth. The restriction a) is relevant because our cooperation indicator specifically refers to husbands, and respondents in households where the female respondent is not married may not consider the response “self and husband" to be valid. The restriction b) is used because outliers in the budget may have excessive influence on the slopes of estimated Engel curves. The restriction c) is because reported zero wealth may actually be mismeasured wealth. These restrictions result in the loss of 302 observations (roughly 10 percent of the sample).

The resulting estimates in column (8) are somewhat different from the baseline. The estimated value of \(\ln \delta\) is 0.173, which is a bit larger than that in the baseline. The associated efficiency gain δ is about 19 percent. Cooperation now increases male and female resource shares by roughly 4 and 1 percentage points, respectively, and decreases children’s resource shares by roughly 5 percentage points. At a gross level, these results are qualitatively the same as the baseline (men gain a lot, women a little and children’s money metric change is insignificant), but the estimated magnitudes are somewhat larger.

The nuclear households in our data have 1 adult man and 1 adult woman and one to four children. We also have 1325 non-nuclear households, having either more than 1 adult man or more than 1 adult woman. These non-nuclear households are a mix of polygamous and multi-generational households. Our model and data might be less appropriate for these non-nuclear households, since our cooperation factor f is only reported by “the main" adult female in the household, and primarily refers to joint decision making by the main adult woman and her husband. Column (9) in Table 4 reports result from estimating the model just with nuclear households, which greatly reduces the sample size.

Again, we see similar patterns as in the baseline case. Cooperation increases efficiency, and induces a shift in resource shares from children towards adult men, with a statistically insigificant impact on the resource shares of women. However, the estimated marginal effect of cooperation f on \(\ln \delta\) is much smaller in these nuclear households than in the baseline model, with an estimated value of 0.078. This means that cooperation induces an efficiency gain of only (\(\exp 0.078\)) 8 percent in nuclear households, compared with 13 percent for all households.

The main difference between nuclear households and the full sample is that the nuclear subsample has smaller households on average. This suggests that the efficiency gains δ may depend on household size. Having more people in a household means goods can be jointly consumed by more household members, leading to greater efficiency. In our model, this can arise because more people sharing a good means a smaller element of A for that good, and hence a lower shadow price.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lewbel, A., Pendakur, K. Estimating a model of inefficient cooperation and consumption in collective households. Rev Econ Household (2024). https://doi.org/10.1007/s11150-023-09688-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11150-023-09688-0

Keywords

JEL classification

Navigation