You are on page 1of 37

Practical Considerations in Raking Survey Data

Michael P. Battaglia1, David Izrael1, David C. Hoaglin1, and Martin R. Frankel1,2


(1) Abt Associates Inc., (2) Baruch College, CUNY

Contact Author: Michael P. Battaglia


Abt Associates Inc., 55 Wheeler Street, Cambridge, MA 02138
(v) 617-349-2425, (f) 617-349-2605, mike_battaglia@abtassoc.com

Abstract
A survey sample may cover segments of the target population in proportions that do not match the
proportions of those segments in the population itself. The differences may arise from sampling
fluctuations, nonresponse, or because the sample design was not able to cover the entire population.
In such situations one can use raking to improve the relation between the sample and the
population by adjusting the sampling weights of the cases in the sample so that the marginal totals
of the adjusted weights on specified characteristics agree with the corresponding totals for the
population.

The raking procedure is described, and convergence issues and problems are

discussed. The details of several practical aspects of raking are then given. The topics covered
have not received much attention in the literature on raking. Specific aspects of raking are
illustrated with graphical displays of output from a SAS Macro that can be obtained for free from
the authors.
Key Words
Control totals, convergence, raking margins, weights, nonresponse

1. Introduction
A survey sample may cover segments of the target population in proportions that do not match the
proportions of those segments in the population itself. The differences may arise, for example,
from sampling fluctuations, from nonresponse, or because the sample design was not able to cover
the entire population. In such situations one can often improve the relation between the sample and
the population by adjusting the sampling weights of the cases in the sample so that the marginal
totals of the adjusted weights on specified characteristics agree with the corresponding totals for
the population. This operation is known as raking ratio estimation (Kalton 1983), raking, or
sample-balancing, and the population totals are usually referred to as control totals. Raking may
reduce nonresponse and noncoverage biases, as well as sampling variability. The initial sampling
weights in the raking process are often equal to the reciprocal of the probability of selection and
may have undergone some adjustments for unit nonresponse and noncoverage. The weights from
the raking process are used in estimation and analysis.

The adjustment to control totals is sometimes achieved by creating a cross-classification of the


categorical control variables (e.g., age categories x gender x race x family-income categories) and
then matching the total of the weights in each cell to the control total. This approach, however, can
spread the sample thinly over a large number of cells. It also requires control totals for all cells of
the cross-classification. Often this is not feasible (e.g., control totals may be available for age x
gender x race but not when those cells are subdivided by family income). The use of marginal
control totals for single variables (i.e., each margin involves only one control variable) often avoids
many of these difficulties. In return, of course, the two-variable (and higher-order) weighted
distributions of the sample are not required to mimic those of the population.

A somewhat different problem motivated the original development of sample-balancing (Deming


1943). The Census Bureau needed to produce tabulations for the joint distribution of two (or more)
variables in the U.S. population, in situations where information on the joint distribution was
available only from a sample. The marginal totals, however, were available for the full population,
and so the sample counts in the cells of the cross-classification were adjusted to provide an
estimated tabulation that had the correct marginal totals.

Raking (or sample-balancing) usually proceeds one variable at a time, applying a proportional
adjustment to the weights of the cases that belong to the same category of the control variable.
Software for sample-balancing has been available for many years, but not as part of SAS (except
for the CLAMAR macro from France) or most other major software systems (WESVAR includes a
raking algorithm). Older readers may be familiar with a FORTRAN program developed in the
1960s by MarketMath, Inc.

Although that program executed rapidly, it had a variety of

disadvantages. The user had to create an ASCII input data set, painstakingly prepare control
statements (the original program was designed to read input from cards), and then process its
ASCII output data set. It could rake on at most 12 variables. Also, it handled rounding in a way
that could lose precision. Izrael et al. (2000) introduced a SAS macro for raking (sometimes
referred to as the IHB raking macro) that combines simplicity and versatility. More recently, the
IHB raking macro was enhanced to increase its utility and diagnostics (Izrael et al. 2004).

The raking algorithm and issues related to convergence are discussed next. Several practical raking
applications are then covered.

2. Basic Algorithm
The procedure known as raking adjusts a set of data so that its marginal totals match specified
control totals on a specified set of variables. The term raking suggests an analogy with the
process of smoothing the soil in a garden plot by alternately working it back and forth with a rake
in two perpendicular directions.

In a simple 2-variable example the marginal totals in various categories for the two variables are
known from the entire population, but the joint distribution of the two variables is known only from
a sample. In the cross-classification of the sample, arranged in rows and columns, one might begin
with the rows, taking each row in turn and multiplying each entry in the row by the ratio of the
population total to the weighted sample total for that category, so that the row totals of the adjusted
data agree with the population totals for that variable. The weighted column totals of the adjusted
data, however, may not yet agree with the population totals for the column variable. Thus the next
step, taking each column in turn, multiplies each entry in the column by the ratio of the population
total to the current total for that category. Now the weighted column totals of the adjusted data
agree with the population totals for that variable, but the new weighted row totals may no longer
match the corresponding population totals. The process continues, alternating between the rows
and the columns, and agreement on both rows and columns is usually achieved after a few
iterations. The result is a tabulation for the population that reflects the relation of the two variables
in the sample.

The above sketch of the raking procedure focuses on the counts in the cells and on the margins of a
two-variable cross-classification of the sample. In the applications that survey statisticians often
encounter, involving data from complex surveys, it is more common to work with the survey

weights of the n individual respondents. Thus, the basic raking algorithm is described in terms of
those individual weights, wi , i = 1, 2,..., n. For an unweighted (i.e., equally weighted) sample, one
can simply take the initial weights to be wi = 1 for each i.

In a cross-classification that has J rows and K columns, denote the sum of the wi in cell ( j , k ) by

w jk . To indicate further summation, replace a subscript by a + sign. Thus, the initial row totals
and column totals of the sample weights are w j + and w+ k , respectively. Analogously, denote the
corresponding population control totals by T j + and T+ k .

The iterative raking algorithm produces modified weights, whose sums are denoted by a suitably
subscripted m with a parenthesized superscript for the number of the step. Thus, in the twovariable cross-classification m jk (1) denotes the sum of the modified weights in cell (j,k) at the end
of Step 1. If one begins by matching the control totals for the rows, T j + , the initial steps of the
algorithm are

m jk (0) = w jk

(j = 1,...,J; k=1,...,K)

m jk (1) = m jk (0) (T j + / m j + (0) )

(for each k within each j)

m jk (2) = m jk (1) (T+ k / m+ k (1) )

(for each j within each k)

The adjustment factors, T j + / m j + (0) and T+ k / m+ k (1) , are actually applied to the individual weights,
which could be denoted by mi (2) for example. In the iterative process an iteration rakes both rows
and columns. Thus, for iteration s ( s = 0, 1, ...) one may write

m jk (2 s +1) = m jk (2 s ) (T j + / m j + (2 s ) )
m jk (2 s + 2) = m jk (2 s +1) (T+ k / m+ k (2 s +1) )

Bishop et al. (1975) discuss the relationship between iterative proportional fitting and raking. They
point out that raking was originally developed not for fitting an unsaturated model to a data set, but
rather for combining information from two or more data sets. In the two-way table discussed
above, one is in effect fitting a fully saturated log-linear model: the two-factor interaction present
in the sample persists after raking, and the one-factor terms (reflected in the population control
totals) are also fitted. Thus, in some ways raking can thus be thought of as fitting a main effects
model, where the main effects correspond to the given margins.

Raking can also adjust a set of data to control totals on three or more variables. In such situations
the control totals often involve single variables, but they may involve two or more variables. In
one example, in raking on three variables one might have control totals Ta++, T+b+, and T++c. In
another example, the control totals might be Tab+ and T++c --- a two-variable margin and a onevariable margin. In actually carrying out the raking for this second example, it suffices to treat the
two-variable margin as the one-variable margin for a composite variable, whose values simply
index the cells of the underlying two-variable margin.

Ideally, one should rake on variables that exhibit strong associations with the key survey outcome
variables or that are strongly related to nonresponse or noncoverage. This strategy will reduce the
mean squared error of the key outcome variables. In practice, other considerations may enter. A
variable such as gender may not be related to key outcome variables or to nonresponse or
noncoverage, but raking on it may be desirable to preserve the face validity of the sample.

3. Convergence
Convergence of the raking algorithm has received considerable attention in the statistical literature,
especially in the context of iterative proportional fitting for log-linear models, where the number of
variables is at least three and the process begins with a different set of initial values in the fitted
table (often 1 in each cell). For raking survey data it is enough that the iterative raking algorithm
(ordinarily) converges, as one would expect from the fact that (in a suitable scale) the fitted cell
counts produced by the raking are the weighted-least-squares fit to the observed cell counts in the
full cross-classification of the sample by all the raking variables (Deming 1943). As an extreme
example, for the 2 x 2 table shown in Table 1, convergence is impossible.

Convergence may require a large number of iterations. Oh and Scheuren (1978) note that the
available convergence proofs make strong assumptions about the cell counts in the crossclassification of the raking variables that no cells are empty or that some particular combination
of nonempty cells is present. They recommend setting up the raking problem in a sensible
manner to avoid: 1) imposing too many marginal constraints on the sample, 2) defining marginal
categories that contain a small percentage of the sample, and 3) imposing contradictory constraints
on the sample.

The authors experience indicates that, in general, raking on a large number of variables slows the
convergence process. However, other factors also affect convergence. One is the number of
categories of the raking variables. Convergence will typically be slower for raking on 10 variables
each with 5 categories than for 10 variables each with only 2 categories. A second factor is the
number of sample cases in each category of the raking variables. Convergence may be slow if any
categories contain fewer than 5% of the sample cases. A third factor is the size of the difference
between each control total and the corresponding weighted sample total prior to raking. If some
differences are large, the number of iterations will typically be higher. One can guard against the
possibility of nonconvergence or slow convergence by setting an upper limit on the number of
iterations (e.g., 50).

Brick et al. (2003) also discuss problems with convergence. They point out that a large number of
iterations indicate a raking application that is not well-behaved and that problems may exist with
the resulting weights highly variable weights inflate sampling variances and produce unstable
domain estimates. One example of a problem is the use of raking variables that have a strong
association (correlation). In this situation the number of iterations may be large, and convergence
will not occur if there are inconsistencies between the associations in the sample and the control
totals (Table 1 shows such an example). The log-linear models literature on structural zeros in
contingency tables is directly related to this issue. For example, if one rakes on Food Stamps
eligibility and a poverty status variable, the cross-tabulation of these two variables in the sample
will likely result in one or more cells that must be empty by definition.

One simple definition of convergence requires that each marginal total of the raked weights be
within a specified tolerance of the corresponding control total. As noted above, in practice, when a

number of raking variables are involved, one must check for the possibility that the iterations do
not converge (e.g., because of sparseness or some other feature in the full cross-classification of the
sample). As already noted, one can guard against this possibility by setting an upper limit on the
number of iterations. As elsewhere in data analysis, it is sensible to examine the sample (including
its joint distribution with respect to all the raking variables) before doing any raking. For example,
if the sample contains no cases in a category of one of the raking variables, it will be necessary to
revise the set of categories and their control totals (say, by combining categories). The authors
recommend, at a minimum, checking the unweighted percentage of sample cases and the
percentage of control cases in each category of each raking variable. Small categories in the
sample or in the control totals (say under 5%) are potential candidates for collapsing. This step
will reduce the chance of creating very unequal weights in raking. Category collapsing always
needs to be done carefully, and in some instances it may be important to retain a small category in
the raking.

4. The IHB Raking Macro


The IHB SAS macro produces diagnostic output that contains the following information: number
of iterations, name of variable currently being raked on, name of BY-variable if there is one, and
marginal control total and calculated total weight for each level of the current raking variable,
along with their difference and percentage difference. At termination, the macro gives the iteration
number at which termination occurred and the reason, which is either that the tolerance has been
met or that the process did not converge. The macro also writes diagnostics into the SAS LOG,
from several of the checks that it makes.

10

Table 2 illustrates the use of the macro with an example involving two raking variables, Table 2
calls them VARIABLE1 and VARIABLE2, and a BY-variable, AREA, which has two levels. The
marginal percentage and general control total for each level of the BY-variable are obtained outside
the example, from PROC FREQ. Preliminary analyses of the data set showed that all categories of
the raking variables represented in the marginal control data sets exist in the sample as well. Table
2 shows the unweighted distribution of each variable. The actual raking uses the weights of the
individual cases. With the convergence tolerance set to 1, the raking converged after 3 iterations
for Area 1, and also after 3 iterations for Area 2.

5. Sources of Control Totals


The discussion of control totals refers to actual totals as opposed to percents. Surveys that use
demographic and socioeconomic variables for raking must locate a source for the population
control totals. An example of a source of true population control totals is the 2000 U.S. Census
short-form data. The U.S. Census long-form variables, the 2000 U.S. Census 5-Percent Public Use
Microdata Sample (PUMS) files, the Current Population Survey (CPS), U.S. Census Bureau
population projections, the National Health Interview Survey, and private-sector sources such as
Claritas are better viewed as control totals, because they are based either on large samples or on
projection methodologies.

Control totals obtained from a sample such as the CPS estimates are subject to much smaller
sampling variability and nonresponse bias, and may be subject to much lower noncoverage bias,
than a survey sample. For state-specific control totals, say for persons aged 0-17 years, the CPS
estimates will be subject to considerably larger sampling variability; thus they are useful for
national control totals, but potentially less useful for stable state control totals. Combining two
years of CPS data can reduce the sampling variability of the state control totals. For projection

11

methods (e.g., age by sex by race mid-year population projections from the U.S. Census Bureau),
the basic approach is to project information forward from 2000 for the non-censal years. Clearly,
the farther one gets from 2000, the greater the likelihood that the projections will be off. This
happened, for example, with the projection of the size of the Hispanic population for the years
before the 2000 Census results came out. Eventually, the American Community Survey should
provide a new source of information for non-censal years.

It is important to make sure that control totals from different sources all add to the same population
total. If not, the raking will not converge. For example, for a survey in the middle of 2003, one
would use Census Bureau age, sex, and race projections of the civilian noninstitutionalized
population for July 2003, and obtain control totals by household income from the March 2003 CPS.
In this situation one would most likely need to ratio-adjust the CPS income control totals so that
they summed to the Census projection control totals for July 2003.

One must also consider how the variables are measured. A telephone survey may ask a single
question to obtain household income. The source for the control totals, however, may have an
income variable that is constructed from a series of questions about income from several sources
(wages, cash-assistance programs, interest, dividends, etc.).

One needs to consider carefully

whether using income as a raking variable makes sense. If the sample is thought to substantially
under-represent low-income persons, then raking on income may be preferred. If, on the other
hand, there is concern that the survey is measuring income very differently from the source of the
control totals, then consideration should be given to raking on a proxy variable such as educational
attainment or even a dichotomous poverty-status variable.

12

Control totals usually do not come with a missing category. The same variable in the survey
may have a nontrivial percentage of cases that fall in a DK or Refused category. In this situation it
may be possible to impute for item nonresponse in the survey before the raking takes place. When
imputation is not feasible, the following procedure can be used to adjust the control totals. Run a
weighted frequency distribution on the raking variable in order to determine the percentage of
sample cases that have a missing value (e.g., 4.3%). Allocate 4.3% of the control total to a newly
created missing category (e.g., 4.3% of 1,500,000 = 64,500). Reapportion the control totals in the
other categories so that they add to the reduced control total (1,500,000 64,500 = 1,435,500).
After raking, the weighted distribution of the sample will agree with the revised control totals and
will reflect a 4.3% missing- data rate in weighted frequencies and tabulations.

6. Trade-offs Related to Number of Margins and Numbers of Categories


Some raking applications use margins for age, sex, and race, because it is relatively easy to obtain
control totals for these variables. In other situations (especially in surveys with lower response or
important noncoverage issues) one may need to rake on a considerably larger number of variables.
This is feasible if control totals can be assembled. The authors have seen rakings that used well
over ten variables. Raking on many variables will almost always require a large number of
iterations. The authors have also seen rakings that used a smaller number of variables, but with
fairly detailed categories. Again, a large number of iterations may be required. In both situations
the cross-classification of the raking variables often yields an extremely large number of cells. For
example, raking on 12 dichotomous variables yields 4,096 cells. Raking on five variables each
containing six categories yields 7,776 cells. Many of these cells will contain no cases in the
sample. Such cells, by definition, remain empty after raking. However, the two-variable, threevariable, and higher-order interactions in the sample are maintained in the raking to the marginal

13

control totals.

The small cell sizes increase the chance that the raked weights will exhibit

considerable variability, because those weights are maintaining sample interactions that are quite
unstable.

On top of the challenges of the numbers of variables and categories and the resulting number of
underlying cells, large differences, before raking, between the weighted sample totals be and the
marginal control totals will generally increase the number of iterations. These issues point to the
need to closely examine: 1) the variables selected for raking, 2) the number and size of the
categories of those raking variables, and 3) the magnitude of differences between the weighted
sample totals and the control totals. Ideal variables for raking are those related to the key survey
outcome variables and related to nonresponse and/or noncoverage. Variables that do not meet
these conditions are candidates for exclusion from raking when a large number of variables are
being considered. The categories of each candidate raking variable should be examined to see
whether they contain a small proportion of the sample cases (say, under 5%) or whether the control
total percentage is small (also, say, under 5%). Such small categories should be considered for
collapsing. Sometimes the small categories of a nominal categorical variables can be collapsed
into a larger residual category. For ordinal variables, collapsing with an adjacent category is often
the best approach. If one or more weighted sample totals differ by a large amount from the
corresponding control totals, one should first try to determine the source of the difference. Is it
extreme differential nonresponse, or has the variable in the sample been measured in a very
different manner than the corresponding variable used to form the control total? One should
consider whether it is appropriate to use such a variable in raking.

14

7. Examining and Diagnosing Slow Convergence


Sometimes the raking process does not converge in a specified number of iterations. As an aid to
diagnosing such situations and taking appropriate action, the enhanced IHB raking macro
incorporates a module that, in case of non-convergence, uses the data to predict the number of
iterations needed for convergence.

The prediction is based on an empirical observation that the logarithm of the magnitude of the
difference between an adjusted weighted total and its control total declines linearly with the
number of iterations. In the authors experience, this relation holds reasonably well when a slowly
converging raking process approaches the specified number of iterations (50 in most applications).
The enhanced macro extrapolates the last iteration slope and estimates the iteration at which the
slowest converging variable will cross a given tolerance threshold.

One usually considers a raking process to be converging slowly if either it does not converge in a
specified number of iterations or convergence takes substantially more iterations than usual. In the
authors work, convergence usually takes place in 5 to 20 iterations. However, when the number of
raking variables is large (say, more than 8) and some of the raking variables have numerous levels
(the variable State with 51 categories, for instance), the process may take much longer to converge
or may even not converge in an initially set number of iterations. The statistician has options to
proceed with raking. The first one is by using the predicted number of iterations from the
diagnostics to rerake the sample, trying to achieve complete convergence. This option is illustrated
later. However, the predicted number of iterations may be impractically large. Then, as a second
option, one may attempt to preprocess the sample data.

15

A common strategy collapses categories of slowly converging variables. If, for instance, State is
such a variable (with a value for each U.S. state and D.C.), it could be collapsed into, say, Census
Division (9 levels) or even Census Region (4 levels). Of course, the statistician may not always
have flexibility in collapsing. He/she may be required to rake by the original variables, or the
slow variables may already be dichotomous. But if there is some flexibility in the statistical
weighting methods, the authors recommend trying collapsing to accelerate convergence.

How does one determine which raking variables are slow? The most effective way to examine a
convergence process is to draw graphs. Figure 1 displays a plot of a slow raking process involving
12 variables; the x-axis is the iteration number, and the y-axis is log10 of the maximum (taken over
all categories of a given raking variable) of the absolute value of the difference between the
adjusted weighted total and the control total. The reference line indicates the tolerance level, in this
example log10(1) = 0. One can easily construct this kind of graph using standard SAS/GRAPH
facilities.

From the graph, one can easily single out the four slowest converging variables (their traces cluster
distinctly higher): EEE, JJJ, GGG, and AAA. The variables GGG and AAA are dichotomous, so it
is not possible to collapse them. To explore how categories of the variables EEE and JJJ (which
are ordinal) converge and which of them might be collapsed, similar graphs show the individual
categories of those two variables (Figure 2).

Besides visual exploration of convergence of slow categories, one should apply common sense
when combining them. For ordinal variables, for instance, it would be logical to combine adjacent

16

categories. Taking the meaning of values of EEE and JJJ into account, in addition to the graphs in
Figure 2, collapsing combined Categories 1 and 2, and Categories 4 and 5 for both variables
(keeping Category 3 separate). Correspondingly, the respective marginal totals were combined,
after which the raking was rerun and new convergence graphs were constructed for those two
collapsed variables (Figure 3). Because convergence of EEE and JJJ looked promising, a new
overall convergence graph was constructed for all 12 raking variables (Figure 4). Comparing this
graph with Figure 1, one can see that collapsing did play a dramatic role in speeding convergence.
The raking process now converges in 17 iterations.

As already noted, the statistician may not always have the flexibility to collapse categories, or
he/she may still want to achieve convergence without altering the raking variables, i.e., using as
many iterations as required. But how many are required? The enhanced macro calculates a
predicted number of iterations needed for full convergence. The graph in Figure 5 demonstrates a
two-variable raking process that initially did not converge in the default 50 iterations (vertical
reference line) and predicted 65 as the needed number. When rerun, the raking did converge at
exactly the 65th iteration. In a fairly rare situation, rerunning the raking with the predicted number
of iterations could give non-convergence again, with a new and much larger number of predicted
iterations. If this occurs, it makes sense to thoroughly examine sample and population data and
make appropriate changes.

8. Inclusion of Two-Variable Raking Margins


Raking can be viewed as analogous to fitting a main-effects-only model. Because of sample size
limitations and/or availability of only one-variable (factor or dimension) control totals, many
raking applications follow this approach. In some situations it may be important to fit a two-

17

variable interaction to the data. For example, one is planning to rake on Variables A, B, C, and D.
However, control totals for Variable C crossed with Variable D are available and exhibit a strong
interaction (e.g., persons aged 0-17 years are more likely to be Hispanic than persons aged 65+
years). If the cell counts in the C x D margin of the sample are large enough to support fitting a C
x D interaction, one would rake on three margins: A, B, and C x D. It is not necessary also to rake
on separate margins for Variables C and D. If, however, the C x D raking margin involved
collapsing one could consider adding one-variable margins to the raking for Variables C and D
without any collapsing of their categories.

9. Forming Control Totals for Quantity Variables


In a specialized raking situation one is planning on raking a sample of persons on some categorical
variables (e.g., age, sex, and race), but the source of the control totals also has a quantity variable
related, to say, the total number of glasses of milk consumed in a week. The survey has also
measured this same quantity variable; but the survey response rate is, let us assume, only 50%.
One may want to ensure that the weighted total number of glasses of milk consumed per week from
the sample agrees closely with the control total. This can be accomplished by dividing the sample
into groups; each group will have a mean number of glasses of milk consumed in a week and a sum
of weights. In the raking process one can modify the sum of the weights in each group so that the
sum of the weights times the mean, summed over all the groups, adds to the control value of total
glasses of milk consumed in a week. In the simplest application one can divide the sample into two
groups: below versus above the median number of glasses of milk consumed in a week based on
the control total data. For each group one can use the control data to obtain the total number of
glasses of milk consumed in a week. This two-category margin is then added to the raking.
Convergence may not occur making it necessary to shift the group boundary point away from the

18

median in order to achieve convergence. Once convergence is achieved the weighted total number
of glasses of milk consumed in a week will be in close agreement with the control total value. This
procedure may be extended to modify not only the total over the entire sample, but for various
subpopulations as well.

10. Raking at the State Level in a Large National Survey


Some large surveys stratify by state and are designed to yield state estimates. The resulting total
national sample is usually very large. The survey statisticians seek to provide national estimates as
well as state estimates. Often one sets up raking control totals at the state level and carries out 51
individual rakings. Assume those rakings use Variables A, B, and C; but the number of categories
of each variable is limited because of the state sample sizes. For example, one might collapse
Variables A, B, and C differently by state. If Variable A were race/ethnicity, one might be able to
use Hispanic as a separate race/ethnicity category in California, but not in Vermont because of the
small sample size. After the 51 rakings one might compare weighted distribution of Variables A,
B, and C with national control totals and observe some differences that are caused by the state-level
collapsing of categories. If having precise weighted distributions at the national level is important
for analytic or face validity reasons, one can use the IHB raking macro in the following manner.

Set up a single raking that includes margins for State x A, State x B, and State x C (i.e., combine
the 51 individual state rakings into a single raking). Then add detailed national margins for
Variables A, B, and C. Another, similar example would involve adding Variable D as a national
raking margin because its control total is available only at the national level (e.g., household
income). This strategy needs to be implemented carefully. Checks should be made for raking
variables that contain small sample sizes. The coefficient of variation of the weights prior to raking

19

and after raking should be examined in each state to check for large increases in the variability of
the weights.

Finally, the raking diagnostics discussed above should be used if convergence

problems arise.

11. Maintaining Prior Nonresponse and Noncoverage Adjustments in the Final weights
Frankel et al. (2003) have discussed methods based on data on interruptions in telephone service
(of a week or longer in the past 12 months) to compensate for the exclusion of persons in
nontelephone households in random-digit-dialing surveys. One typically adjusts the base sampling
weights of persons with versus without an interruption in telephone service.

The resulting

interruption-based weight adjusts for the noncoverage of nontelephone households. If one then
rakes the sample on age, sex, and race, the impact of the nontelephone adjustment may be diluted
somewhat, even though the raking starts with interruption-based weight. In that case it generally
makes sense to create weighted control totals (using the interruption-based weight) from the sample
for persons residing in households with versus without an interruption in telephone service. These
weighted control totals should be ratio-adjusted so that they have the same sum as the age, sex, and
race control totals. For example, if the age, sex, and race margins sum to 180,000,000 persons,
then the interruption margin needs to be adjusted so that it also sums to 180,000,000. The raking
would use the four variables instead of just three and would ensure that the nontelephone
adjustment is fully reflected in the final weights. This would be appropriate where the interruptionin-telephone-service category could be small (e.g., in states where telephone coverage is very
high), but one still wants to maintain that small category in the raking.

20

12. Raking Surveys that Screen for a Specific Target Population


A common survey model for obtaining interviews with a specific target population is to screen a
sample of households for the presence of members of the target population. An example would be
children with special health care needs. The screening interview collects a roster of children with,
say, their age, sex, and race, and determines whether each child has special health care needs. If
the household contains one child with special health care needs, a detailed interview is conducted
for that child. If the household has two or more such children, one is selected at random for the
detailed interview. Of course, the interview response rate will be less than 100%, because some
parents will not agree to do the detailed interview.

Assume that the survey statisticians need to look at the prevalence of children with special health
care needs, and they will also be analyzing the detailed interview data. In this situation one would
calculate the usual base sampling weights, make adjustments for unit nonresponse and possibly
make a noncoverage adjustment if warranted. One first obtains control totals for age, sex, and race
in the U.S. population aged 0-17 years. One then rakes the entire sample of children in the
screened households to those control totals, because that sample is a sample of children aged 0-17
in the U.S. The resulting screener weights can then be used to estimate the prevalence of children
with special health care needs in the U.S.

That screener weight would typically serve as the input weight in the calculation of weights for the
children with completed detailed interviews. As part of that calculation process one also seeks to
weight the detailed-interview sample by age, sex, and race. Of course, control totals are unlikely to
be available for children with special health care needs. One can, however, use the screener weight

21

and the sample of children with special health care needs identified in the screened households to
form weighted control totals for age, sex, and race and then use those in raking the detailedinterview weights.

This method ensures that the survey analysts do not ask why the age

distribution of children with special health care needs from the screener sample does not agree
exactly with the distribution in the detailed interview data. Some caution needs to be exercised in
using this approach when the screener shows survey evidence of false positives.

13. Raking to Control Totals Expressed as Percentages and Raking with No Input Weight
Frequently, the user working with a weighted or an unweighted sample needs to weight it to fit
marginal population proportions. As an example (Table 3), the authors created an 11-case sample
data set that contains two variables: VAR1, which takes values 1, 2, and 3 with frequencies
27.27%, 45.45% and 27.27%, respectively; and VAR2, which takes values 1 and 2 with
frequencies 45.45% and 54.55%, respectively. The objective was to weight this sample so that the
distributions of VAR1 and VAR2 met the population distributions --- (20%, 35%, 45%) and (60%,
40%), respectively --- within a tolerance of 0.001%.

14. Weight Trimming and Raking


Weight trimming refers to truncation of high or extreme weight values in order to reduce their
impact on the variance of the estimates, especially for subgroup estimates. One consequence of the
truncation of high weight values is that the weights of the entire sample will not add to the
population size. Although weight trimming is a separate topic from raking; they are certainly
related in the sense that weight trimming typically takes place at the last step in the calculations,
which is often raking. Many large surveys use weight trimming (Srinath 2003, Abt Associates
memorandum). Its objective is to reduce the mean squared error of the key outcome estimates. By

22

trimming high weight values one generally lowers sampling variability but may incur some bias.
The MSE will be lower if the reduction in variance is large relative to the increase in bias arising
from weight trimming. There are no established rules for weight trimming; rather most people use
a general set of guidelines. Some common truncation points are: 1) the median weight plus five or
six times the interquartile range (IQR) of the weights, 2) five times the mean weight, 3) the 95th
percentile of the weights.

How can weight trimming be incorporated in raking? The IHB SAS macro can be used for weight
trimming in the following steps (using as an example the median weight plus six times the IQR as
the truncation point)1:

1. Prior to raking i, where i references the number of times the raking is run, examine the
distribution of the raking input weight and calculate the median weight plus six times the
interquartile (IQR) range of the weights.
2. Truncate values of the input weight that are above the median weight plus six times the
IQR plus one to the median weight plus six times the IQR (values at or below the median
weight plus six times the IQR plus one are not altered).
3. Using the truncated input weight, run the raking to obtain raking weight i.
4. Repeat Steps 1 to 3 (i.e., run the raking a second time, third time, etc.) until there are no
weights that are above the median weight plus six times the IQR plus one.

Although the cutoff value equals the median weight plus six times the IQR, weights that exceed the
median weight plus six times the IQR plus one are truncated to the median weight plus six times
1

A somewhat more sophisticated, but computer intensive, procedure is to apply bounds to the weights as the
raking is taking place.

23

the IQR, because the raking may increase the weight values of the cases that have been truncated,
and thus cause the raking steps to repeat endlessly. The approach described above does not
guarantee convergence (i.e., after running the raking several times there could still be weights
above the median weight plus six times the IQR plus one), and one could consider adding a larger
constant to increase the chances of convergence, but the authors have found in their applications
that convergence is often achieved by adding a constant of one. Table 4 shows an example of the
use of weight trimming with raking. Before raking there are four cases with input weights that
exceed the median weight plus six times the IQR plus one of 439.847 (condition). The weights of
those cases are truncated to 438.847 (cutoff) and the raking is run for the first time. After the first
raking the condition equals 444.490. Only one case has a weight that exceeds this value and that
weight is truncated to the cutoff of 443.490. After the second raking no cases have a weight that
exceeds the condition and the process is stopped. The weights from the second raking add to the
population size and meet the raking control totals.

15. Summary
The authors have sought to give some background on how raking works and to discuss the
convergence process. They have also sought to give some warnings of conditions that need to be
checked before and after raking. Brick et al. (2003) discuss other examples of issues that one
should be aware of when using raking. The IHB SAS macro discussed in this paper is available for
free from the first author.

24

References
Bishop YMM, Fienberg SE, and Holland PW. (1975). Discrete Multivariate Analysis: Theory and
Practice. Cambridge, MA: MIT Press.

Brick JM, Montaquila J, and Roth S. (2003). Identifying Problems with Raking Estimators. 2003
Proceedings of the Annual Meeting of the American Statistical Association [CD-ROM],
Alexandria, VA: American Statistical Association, pp. 710-717.

Deming WE. (1943). Statistical Adjustment of Data. New York: Wiley.

Frankel MR, Srinath KP, Hoaglin DC, Battaglia MP, Smith PJ, Wright RA, and Khare M. (2003).
Adjustments for non-telephone bias in random-digit-dialling surveys. Statistics in Medicine,
Volume 22, pp. 1611-1626.

Izrael D, Hoaglin, DC, and Battaglia MP. (2000). A SAS Macro for Balancing a Weighted Sample.
Proceedings of the Twenty-Fifth Annual SAS Users Group International Conference, Cary, NC:
SAS Institute Inc., pp. 1350-1355.

Izrael D, Hoaglin DC, and Battaglia MP. (2004). To Rake or Not To Rake Is Not the Question
Anymore with the Enhanced Raking Macro. May 2004 SUGI Conference, Montreal, Canada.

Kalton G. (1983). Compensating for Missing Survey Data. Survey Research Center, Institute for
Social Research, University of Michigan.

25

Oh HL, and Scheuren F. (1978). Some Unresolved Application Issues in Raking Ratio Estimation.
1978 Proceedings of the Section on Survey Research Methods, Washington, DC: American
Statistical Association, pp. 723-728.

26

Table 1. A 2 x 2 Table for Which Raking Cannot Produce Agreement with the Control Totals

Variable 2

Marginal
Control Total

Variable 1
1
2
Marginal
Control
Total

1
20
0

2
0
10

70
30

50

50

100

27

Table 2. Example of Raking Using the IHB SAS Macro


Raking AREA - 1 VARIABLE1, iteration - 1

VARIABLE1
1
2
3

Calculated
margin

Marginal
Control
Total

15915.87
10912.05
18022.90
==========
44850.82

22154.39
16533.88
24112.03
========
62800.30

Raking AREA - 1

VARIABLE2
1
2

Calculated
margin

Marginal
Control
Total

32684.74
30115.56
==========
62800.30

30697.33
32102.97
========
62800.30

Raking AREA - 1

VARIABLE1
1
2
3

Calculated
margin

Marginal
Control
Total

22102.81
16442.32
24255.17
==========
62800.30

22154.39
16533.88
24112.03
========
62800.30

Raking AREA - 1

VARIABLE2
1
2

Calculated
margin

Marginal
Control
Total

30708.98
32091.32
==========
62800.30

30697.33
32102.97
========
62800.30

Raking AREA - 1

VARIABLE1
1
2
3

Calculated
margin

Marginal
Control
Total

22154.09
16533.34
24112.87
==========
62800.30

22154.39
16533.88
24112.03
========
62800.30

Difference
6238.52
5621.83
6089.13

Calculated
%

Marginal
Control
%

35.486
24.330
40.184
==========
100.00

35.278
26.328
38.395
========
100.00

Difference
in %
0.209
-1.998
1.789

VARIABLE2, iteration - 1

Difference
-1987.40
1987.40

Calculated
%

Marginal
Control
%

52.046
47.954
==========
100.00

48.881
51.119
========
100.00

Difference
in %
3.165
-3.165

VARIABLE1, iteration - 2

Difference
51.586
91.553
-143.139

Calculated
%

Marginal
Control
%

35.195
26.182
38.623
==========
100.00

35.278
26.328
38.395
========
100.00

Difference
in %
-0.082
-0.146
0.228

VARIABLE2, iteration - 2

Difference
-11.6455
11.6455

Calculated
%

Marginal
Control
%

48.899
51.101
==========
100.00

48.881
51.119
========
100.00

Difference
in %
0.019
-0.019

VARIABLE1, iteration - 3

Difference
0.29992
0.53717
-0.83710

28

Calculated
%

Marginal
Control
%

35.277
26.327
38.396
==========
100.00

35.278
26.328
38.395
========
100.00

Difference
in %
-0.000
-0.001
0.001

Raking AREA - 1

VARIABLE2
1
2

Calculated
margin

Marginal
Control
Total

30697.40
32102.90
==========
62800.30

30697.33
32102.97
========
62800.30

VARIABLE2, iteration - 3

Difference
-0.068148
0.068148

Calculated
%

Marginal
Control
%

48.881
51.119
==========
100.00

48.881
51.119
========
100.00

Difference
in %
0.000
-0.000

**** Program for AREA 1 terminated at iteration 3 because all calculated margins
differ from Marginal Control Totals by less than 1

Raking AREA - 2

VARIABLE1
1
2
3

Calculated
margin

Marginal
Control
Total

31377.80
17512.57
26295.48
==========
75185.84

38598.04
29596.11
35307.30
=========
103501.44

Raking AREA - 2

VARIABLE2
1
2

Calculated
margin

Marginal
Control
Total

51930.05
51571.39
==========
103501.44

51902.14
51599.30
=========
103501.44

Raking AREA - 2

VARIABLE1
1
2
3

Calculated
margin

Marginal
Control
Total

38596.66
29599.80
35304.98
==========
103501.44

38598.04
29596.11
35307.30
=========
103501.44

Raking AREA - 2

VARIABLE2
1
2

Calculated
margin

Marginal
Control
Total

51902.75
51598.69
==========
103501.44

51902.14
51599.30
=========
103501.44

VARIABLE1, iteration - 1

Difference
7220.24
12083.54
9011.82

Calculated
%

Marginal
Control
%

41.734
23.292
34.974
==========
100.00

37.292
28.595
34.113
========
100.00

Difference
in %
4.441
-5.303
0.861

VARIABLE2, iteration - 1

Difference
-27.9123
27.9123

Calculated
%

Marginal
Control
%

50.173
49.827
==========
100.00

50.146
49.854
========
100.00

Difference
in %
0.027
-0.027

VARIABLE1, iteration - 2

Difference
1.37510
-3.69114
2.31605

Calculated
%

Marginal
Control
%

37.291
28.598
34.111
==========
100.00

37.292
28.595
34.113
========
100.00

Difference
in %
-0.001
0.004
-0.002

VARIABLE2, iteration - 2

Difference
-0.61296
0.61296

29

Calculated
%

Marginal
Control
%

50.147
49.853
==========
100.00

50.146
49.854
========
100.00

Difference
in %
0.001
-0.001

Raking AREA - 2

VARIABLE1
1
2
3

Calculated
margin

Marginal
Control
Total

38598.01
29596.19
35307.25
==========
103501.44

38598.04
29596.11
35307.30
=========
103501.44

Raking AREA - 2

VARIABLE2
1
2

Calculated
margin

Marginal
Control
Total

51902.15
51599.29
==========
103501.44

51902.14
51599.30
=========
103501.44

VARIABLE1, iteration - 3

Difference
0.030193
-0.081052
0.050859

Calculated
%

Marginal
Control
%

37.292
28.595
34.113
==========
100.00

37.292
28.595
34.113
========
100.00

Difference
in %
-0.000
0.000
-0.000

VARIABLE2, iteration - 3

Difference
-0.013460
0.013460

Calculated
%

Marginal
Control
%

50.146
49.854
==========
100.00

50.146
49.854
========
100.00

**** Program for AREA 2 terminated at iteration 3 because all calculated margins
differ from Marginal Control Totals by less than 1

30

Difference
in %
0.000
-0.000

Figure 1. Convergence of a Raking Process Involving 12 Variables

-1

-2
0

variable

10

11

12

AAA
GGG

13

14

15

16

17

18

19

20

21

BBB
HHH

22

23

24

25

26

CCC
III

31

27

28

29

30

31

32

33

DDD
JJJ

34

35

36

37

38

39

40

EEE
KKK

41

42

43

44

45

46

47

FFF
LLL

48

49

50

51

Figure 2. Convergence of Variables EEE and JJJ before Collapsing

Variable EEE

Variable JJJ

4
3

3
2

0
0

-1
-1

-2

-2

-3
0

10

category

12

14

16

18

20

22

24

26

28

30

32

34

36

38

40

42

44

46

48

50

10

category

32

12

14

16

18

20

22

24

26

28

30

32

34

36

38

40

42

44

46

48

50

Figure 3. Convergence of Variables EEE and JJJ after Collapsing

Variable JJJ

Variable EEE

2
2

1
1

-1

-1
-2

-2
0

10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42

category

44 46 48 50

10

12

14

16

category

Category 2 and 4 collapsed into 1 and 5 respectively

18

20

22

24

26

28

30

32

34

36

38

40

Category 2 and 4 collapsed into 1 and 5 respectively

33

42

44

46

48

50

Figure 4. Convergence of All 12 Variables in the Raking Process after collapsing Variables EEE
and JJJ.

-1

-2

-3
0

variable

10

11

12

AAA
GGG

13

14

15

16

17

18

19

20

21

BBB
HHH

22

23

24

25

26

CCC
III

34

27

28

29

30

31

32

33

DDD
JJJ

34

35

36

37

38

39

40

EEE
KKK

41

42

43

44

45

46

47

FFF
LLL

48

49

50

51

Figure 5. Prediction of the Number of Iterations Needed for Convergence

-1
0

10

12

14

16

18

20

22

24

variable

26

28

30

32

34

AAA

36

38

40

42

44

46

48

BBB

predicted number of iterations for convergence - 65

35

50

52

54

56

58

60

62

64

66

Table 3. Raking Using Marginal Percentage Controls and No Input Weight (first and last
iteration shown).
The FREQ Procedure
Cumulative
Cumulative
VAR1
Frequency
Percent
Frequency
Percent
--------------------------------------------------------1
3
27.27
3
27.27
2
5
45.45
8
72.73
3
3
27.27
11
100.00

Cumulative
Cumulative
VAR2
Frequency
Percent
Frequency
Percent
--------------------------------------------------------1
5
45.45
5
45.45
2
6
54.55
11
100.00

Raking

VAR1
1
2
3

Calculated
margin

Marginal
Control
Total

3
5
3
==========
11

20
35
45
========
100
Raking

VAR2
1
2

Calculated
margin

Marginal
Control
Total

42.667
57.333
==========
100.000

60
40
========
100
Raking

VAR1
1
2
3

Calculated
margin

Marginal
Control
Total

20.000
35.001
44.999
==========
100.000

20
35
45
========
100
Raking

VAR2
1
2

Calculated
margin

Marginal
Control
Total

60.000
40.000
==========
100.000

60
40
========
100

VAR1, iteration - 1

Difference
17
30
42

Calculated
%

Marginal
Control
%

27.273
45.455
27.273
==========
100.00

20.000
35.000
45.000
========
100.00

Difference
in %
7.273
10.455
-17.727

VAR2, iteration - 1

Difference
17.3333
-17.3333

Calculated
%

Marginal
Control
%

42.667
57.333
==========
100.00

60.000
40.000
========
100.00

Difference
in %
-17.333
17.333

VAR1, iteration - 5

Difference
0.000256716
-.000834329
0.000577612

Calculated
%

Marginal
Control
%

20.000
35.001
44.999
==========
100.00

20.000
35.000
45.000
========
100.00

Difference
in %
-0.000
0.001
-0.001

VAR2, iteration - 5

Difference
0.000205597
-.000205597

Calculated
%

Marginal
Control
%

60.000
40.000
==========
100.00

60.000
40.000
========
100.00

Difference
in %
-0.000
0.000

**** Program terminated at iteration 5 because all Calculated Percents differ from Marginal
Percents by less than 0.001

36

Table 4: Example of Weight Trimming During Raking

OBSERVATIONS IN ORIGINAL DATASET TO BE TRUNCATED


CUTOFF: MEDIAN+6*IQR
CONDITION: MEDIAN+6*IQR +1

weight_to_
id

truncate

mean

median

IQR

cutoff

condition

715

477.576

144.250

132.491

51.0592

438.847

439.847

651

509.018

144.250

132.491

51.0592

438.847

439.847

1085

690.762

144.250

132.491

51.0592

438.847

439.847

770

515.720

144.250

132.491

51.0592

438.847

439.847

OBSERVATIONS TO BE TRUNCATED AFTER ITERATION = 1


CUTOFF: MEDIAN+6*IQR
CONDITION: MEDIAN+6*IQR + 1

truncated_
id

1085

weight

451.059

mean

144.250

median

133.108

IQR

51.7302

cutoff

443.490

condition

444.490

OBSERVATIONS TO BE TRUNCATED AFTER ITERATION = 2


CUTOFF: MEDIAN+6*IQR
CONDITION: MEDIAN+6*IQR + 1

THERE ARE NO WEIGHTS TO TRUNCATE

RERAKING-TRUNCATION PROCESS CONVERGED IN 2 ITERATIONS WITH CONDITION MEDIAN+6*IQR+1.

37

You might also like