Lazier Than Lazy Greedy
Baharan Mirzasoleiman
ETH Zurich
Ashwinkumar Badanidiyuru
Google Research Mountain View
Amin Karbasi
Yale University
Jan Vondr
´
ak
IBM Almaden
Andreas Krause
ETH Zurich
Abstract
Is it possible to maximize a monotone submodular function
faster than the widely used lazy greedy algorithm (also known
as accelerated greedy), both in theory and practice? In this
paper, we develop the first linear-time algorithm for maxi-
mizing a general monotone submodular function subject to
a cardinality constraint. We show that our randomized algo-
rithm, STOCHASTIC-GREEDY, can achieve a (1 1/e ε)
approximation guarantee, in expectation, to the optimum so-
lution in time linear in the size of the data and independent of
the cardinality constraint. We empirically demonstrate the ef-
fectiveness of our algorithm on submodular functions arising
in data summarization, including training large-scale kernel
methods, exemplar-based clustering, and sensor placement.
We observe that STOCHASTIC-GREEDY practically achieves
the same utility value as lazy greedy but runs much faster.
More surprisingly, we observe that in many practical scenar-
ios STOCHASTIC-GREEDY does not evaluate the whole frac-
tion of data points even once and still achieves indistinguish-
able results compared to lazy greedy.
Introduction
For the last several years, we have witnessed the emergence
of datasets of an unprecedented scale across different scien-
tific disciplines. The large volume of such datasets presents
new computational challenges as the diverse, feature-rich,
unstructured and usually high-resolution data does not allow
for effective data-intensive inference. In this regard, data
summarization is a compelling (and sometimes the only)
approach that aims at both exploiting the richness of large-
scale data and being computationally tractable. Instead of
operating on complex and large data directly, carefully con-
structed summaries not only enable the execution of vari-
ous data analytics tasks but also improve their efficiency and
scalability.
In order to effectively summarize the data, we need to
define a measure for the amount of representativeness that
lies within a selected set. If we think of representative el-
ements as the ones that cover best, or are most informative
w.r.t. the items in a dataset then naturally adding a new el-
ement to a set of representatives, say A, is more beneficial
than adding it to its superset, say B A, as the new element
Copyright
c
2015, Association for the Advancement of Artificial
Intelligence (www.aaai.org). All rights reserved.
can potentially enclose more uncovered items when consid-
ered with elements in A rather than B. This intuitive di-
minishing returns property can be systematically formalized
through submodularity (c.f., Nemhauser, Wolsey, and Fisher
(1978)). More precisely, a submodular function f : 2
V
R
assigns a subset A V a utility value f(A) –measuring the
representativeness of the set A– such that
f(A {i}) f(A) f(B {i}) f (B)
for any A B V and i V \ B. Note that
∆(i|A)
.
= f (A {i}) f(A) measures the marginal gain
of adding a new element i to a summary A. Of course,
the meaning of representativeness (or utility value) depends
very much on the underlying application; for a collection
of random variables, the utility of a subset can be mea-
sured in terms of entropy, and for a collection of vectors, the
utility of a subset can be measured in terms of the dimen-
sion of a subspace spanned by them. In fact, summariza-
tion through submodular functions has gained a lot of inter-
est in recent years with application ranging from exemplar-
based clustering (Gomes and Krause 2010), to document
(Lin and Bilmes 2011; Dasgupta, Kumar, and Ravi 2013)
and corpus summarization (Sipos et al. 2012), to recom-
mender systems (Leskovec et al. 2007; El-Arini et al. 2009;
El-Arini and Guestrin 2011).
Since we would like to choose a summary of a manage-
able size, a natural optimization problem is to find a sum-
mary A
of size at most k that maximizes the utility, i.e.,
A
= argmax
A:|A|≤k
f(A). (1)
Unfortunately, this optimization problem is NP-hard for
many classes of submodular functions (Nemhauser and
Wolsey 1978; Feige 1998). We say a submodular function is
monotone if for any A B V we have f(A) f(B). A
celebrated result of Nemhauser, Wolsey, and Fisher (1978)
–with great importance in artificial intelligence and machine
learning– states that for non-negative monotone submodu-
lar functions a simple greedy algorithm provides a solution
with (1 1/e) approximation guarantee to the optimal (in-
tractable) solution. This greedy algorithm starts with the
empty set A
0
and in iteration i, adds an element maximiz-
ing the marginal gain ∆(e|A
i1
). For a ground set V of
size n, this greedy algorithm needs O(n · k) function eval-
uations in order to find a summarization of size k. How-
ever, in many data intensive applications, evaluating f is
Proceedings of the Twenty-Ninth AAAI Conference on Artificial Intelligence
1812
expensive and running the standard greedy algorithm is in-
feasible. Fortunately, submodularity can be exploited to im-
plement an accelerated version of the classical greedy algo-
rithm, usually called LAZY-GREEDY (Minoux 1978a). In-
stead of computing ∆(e|A
i1
) for each element e V , the
LAZY-GREEDY algorithm keeps an upper bound ρ(e) (ini-
tially ) on the marginal gain sorted in decreasing order. In
each iteration i, the LAZY-GREEDY algorithm evaluates the
element on top of the list, say e, and updates its upper bound,
ρ(e) ∆(e|A
i1
). If after the update ρ(e) ρ(e
0
) for all
e
0
6= e, submodularity guarantees that e is the element with
the largest marginal gain. Even though the exact cost (i.e.,
number of function evaluations) of LAZY-GREEDY is un-
known, this algorithm leads to orders of magnitude speedups
in practice. As a result, it has been used as the state-of-the-
art implementation in numerous applications including net-
work monitoring (Leskovec et al. 2007), network inference
(Rodriguez, Leskovec, and Krause 2012), document sum-
marization (Lin and Bilmes 2011), and speech data subset
selection (Wei et al. 2013), to name a few. However, as the
size of the data increases, even for small values of k, run-
ning LAZY-GREEDY is infeasible. A natural question to ask
is whether it is possible to further accelerate LAZY-GREEDY
by a procedure with a weaker dependency on k. Or even bet-
ter, is it possible to have an algorithm that does not depend
on k at all and scales linearly with the data size n?
In this paper, we propose the first linear-time algorithm,
STOCHASTIC-GREEDY, for maximizing a non-negative
monotone submodular function subject to a cardinality con-
straint k. We show that STOCHASTIC-GREEDY achieves
a (1 1/e ) approximation guarantee to the opti-
mum solution with running time O(n log(1/)) (measured
in terms of function evaluations) that is independent of
k. Our experimental results on exemplar-based clustering
and active set selection in nonparametric learning also con-
firms that STOCHASTIC-GREEDY consistently outperforms
LAZY-GREEDY by a large margin while achieving practi-
cally the same utility value. More surprisingly, in our exper-
iments we observe that STOCHASTIC-GREEDY sometimes
does not even evaluate all the items and shows a running
time that is less than n while still providing solutions close
to the ones returned by LAZY-GREEDY. Due to its inde-
pendence of k, STOCHASTIC-GREEDY is the first algorithm
that truly scales to voluminous datasets.
Related Work
Submodularity is a property of set functions with deep the-
oretical and practical consequences. For instance, sub-
modular maximization generalizes many well-known com-
binatorial problems including maximum weighted match-
ing, max coverage, and facility location, to name a few. It
has also found numerous applications in machine learning
and artificial intelligence such as influence maximization
(Kempe, Kleinberg, and Tardos 2003), information gath-
ering (Krause and Guestrin 2011), document summariza-
tion (Lin and Bilmes 2011) and active learning (Guillory
and Bilmes 2011; Golovin and Krause 2011). In most of
these applications one needs to handle increasingly larger
quantities of data. For this purpose, accelerated/lazy vari-
ants (Minoux 1978a; Leskovec et al. 2007) of the celebrated
greedy algorithm of Nemhauser, Wolsey, and Fisher (1978)
have been extensively used.
Scaling Up: To solve the optimization problem (1) at
scale, there have been very recent efforts to either make
use of parallel computing methods or treat data in a
streaming fashion. In particular, Chierichetti, Kumar, and
Tomkins (2010) and Blelloch, Peng, and Tangwongsan
(2011) addressed a particular instance of submodular
functions, namely, maximum coverage and provided a
distributed method with a constant factor approximation
to the centralized algorithm. More generally, Kumar et al.
(2013) provided a constant approximation guarantee for
general submodular functions with bounded marginal gains.
Contemporarily, Mirzasoleiman et al. (2013) developed a
two-stage distributed algorithm that guarantees solutions
close to the optimum if the dataset is massive and the
submodular function is smooth.
Similarly, Gomes and Krause (2010) presented a heuristic
streaming algorithm for submodular function maximization
and showed that under strong assumptions about the way
the data stream is generated their method is effective. Very
recently, Badanidiyuru et al. (2014) provided the first one-
pass streaming algorithm with a constant factor approxima-
tion guarantee for general submodular functions without any
assumption about the data stream.
Even though the goal of this paper is quite different
and complementary in nature to the aforementioned work,
STOCHASTIC-GREEDY can be easily integrated into ex-
isting distributed methods. For instance, STOCHASTIC-
GREEDY can replace LAZY-GREEDY for solving each sub-
problem in the approach of Mirzasoleiman et al. (2013).
More generally, any distributed algorithm that uses LAZY-
GREEDY as a sub-routine, can directly benefit from our
method and provide even more efficient large-scale algorith-
mic frameworks.
Our Approach: In this paper, we develop the first central-
ized algorithm whose cost (i.e., number of function evalua-
tions) is independent of the cardinality constraint, which in
turn directly addresses the shortcoming of LAZY-GREEDY.
Perhaps the closest, in spirit, to our efforts are approaches
by Wei, Iyer, and Bilmes (2014) and Badanidiyuru and
Vondr
´
ak (2014). Concretely, Wei, Iyer, and Bilmes (2014)
proposed a multistage algorithm, MULTI-GREEDY, that
tries to decrease the running time of LAZY-GREEDY by
approximating the underlying submodular function with a
set of (sub)modular functions that can be potentially eval-
uated less expensively. This approach is effective only
for those submodular functions that can be easily decom-
posed and approximated. Note again that STOCHASTIC-
GREEDY can be used for solving the sub-problems in each
stage of MULTI-GREEDY to develop a faster multistage
method. Badanidiyuru and Vondr
´
ak (2014) proposed a dif-
ferent centralized algorithm that achieves a (1 1/e
) approximation guarantee for general submodular func-
tions using O(n/ log(n/)) function evaluations. However,
1813
STOCHASTIC-GREEDY consistently outperforms their algo-
rithm in practice in terms of cost, and returns higher util-
ity value. In addition, STOCHASTIC-GREEDY uses only
O(n log(1/)) function evaluations in theory, and thus pro-
vides a stronger analytical guarantee.
STOCHASTIC-GREEDY Algorithm
In this section, we present our randomized greedy algorithm
STOCHASTIC-GREEDY and then show how to combine it
with lazy evaluations. We will show that STOCHASTIC-
GREEDY has provably linear running time independent of
k, while simultaneously having the same approximation ra-
tio guarantee (in expectation). In the following section we
will further demonstrate through experiments that this is also
reflected in practice, i.e., STOCHASTIC-GREEDY is substan-
tially faster than LAZY-GREEDY, while being practically
identical to it in terms of the utility.
The main idea behind STOCHASTIC-GREEDY is to pro-
duce an element which improves the value of the solution
roughly the same as greedy, but in a fast manner. This is
achieved by a sub-sampling step. At a very high level this
is similar to how stochastic gradient descent improves the
running time of gradient descent for convex optimization.
Random Sampling
The key reason that the classic greedy algorithm works is
that at each iteration i, an element is identified that reduces
the gap to the optimal solution by a significant amount, i.e.,
by at least (f(A
) f(A
i1
))/k. This requires n oracle
calls per step, the main bottleneck of the classic greedy al-
gorithm. Our main observation here is that by submodu-
larity, we can achieve the same improvement by adding a
uniformly random element from A
to our current set A. To
get this improvement, we will see that it is enough to ran-
domly sample a set R of size (n/k) log(1/), which in turn
overlaps with A
with probability 1 . This is the main
reason we are able to achieve a boost in performance.
The algorithm is formally presented in Algorithm 1. Sim-
ilar to the greedy algorithm, our algorithm starts with an
empty set and adds one element at each iteration. But in
each step it first samples a set R of size (n/k) log(1/) uni-
formly at random and then adds the element from R to A
which increases its value the most.
Algorithm 1 STOCHASTIC-GREEDY
Input: f : 2
V
R
+
, k {1, . . . , n}.
Output: A set A V satisfying |A| k.
1: A .
2: for (i 1; i k; i i + 1) do
3: R a random subset obtained by sampling s random
elements from V \ A.
4: a
i
argmax
aR
∆(a|A).
5: A A {a
i
}
6: return A.
Our main theoretical result is the following. It shows that
STOCHASTIC-GREEDY achieves a near-optimal solution for
general monotone submodular functions, with computa-
tional complexity independent of the cardinality constraint.
Theorem 1. Let f be a non-negative monotone submoduar
function. Let us also set s =
n
k
log
1
. Then STOCHASTIC-
GREEDY achieves a (1 1/e ) approximation guarantee
in expectation to the optimum solution of problem (1) with
only O(n log
1
) function evaluations.
Since there are k iterations in total and at each it-
eration we have (n/k) log(1/) elements, the total
number of function evaluations cannot be more than
k × (n/k) log(1/) = n log(1/). The proof of the
approximation guarantee is given in the analysis section.
Random Sampling with Lazy Evaluation
While our theoretical results show a provably linear time al-
gorithm, we can combine the random sampling procedure
with lazy evaluation to boost its performance. There are
mainly two reasons why lazy evaluation helps. First, the
randomly sampled sets can overlap and we can exploit the
previously evaluated marginal gains. Second, as in LAZY-
GREEDY although the marginal values of the elements might
change in each step of the greedy algorithm, often their or-
dering does not change (Minoux 1978a). Hence in line 4
of Algorithm 1 we can apply directly lazy evaluation as fol-
lows. We maintain an upper bound ρ(e) (initially ) on
the marginal gain of all elements sorted in decreasing order.
In each iteration i, STOCHASTIC-GREEDY samples a set R.
From this set R it evaluates the element that comes on top
of the list. Let’s denote this element by e. It then updates
the upper bound for e, i.e., ρ(e) ∆(e|A
i1
). If after the
update ρ(e) ρ(e
0
) for all e
0
6= e where e, e
0
R, sub-
modularity guarantees that e is the element with the largest
marginal gain in the set R. Hence, lazy evaluation helps us
reduce function evaluation in each round.
Experimental Results
In this section, we address the following questions: 1)
how well does STOCHASTIC-GREEDY perform compared
to previous art and in particular LAZY-GREEDY, and 2)
How does STOCHASTIC-GREEDY help us get near optimal
solutions on large datasets by reducing the computational
complexity? To this end, we compare the performance of
our STOCHASTIC-GREEDY method to the following bench-
marks: RANDOM-SELECTION, where the output is k ran-
domly selected data points from V ; LAZY-GREEDY, where
the output is the k data points produced by the accelerated
greedy method (Minoux 1978a); SAMPLE-GREEDY, where
the output is the k data points produced by applying LAZY-
GREEDY on a subset of data points parametrized by sam-
pling probability p; and THRESHOLD-GREEDY, where the
output is the k data points provided by the algorithm of
Badanidiyuru et al. (2014). In order to compare the com-
putational cost of different methods independently of the
concrete implementation and platform, in our experiments
we measure the computational cost in terms of the number
of function evaluations used. Moreover, to implement the
SAMPLE-GREEDY method, random subsamples are gener-
ated geometrically using different values for probability p.
1814
Higher values of p result in subsamples of larger size from
the original dataset. To maximize fairness, we implemented
an accelerated version of THRESHOLD-GREEDY with lazy
evaluations (not specified in the paper) and report the best
results in terms of function evaluations. Among all bench-
marks, RANDOM-SELECTION has the lowest computational
cost (namely, one) as we need to only evaluate the selected
set at the end of the sampling process. However, it provides
the lowest utility. On the other side of the spectrum, LAZY-
GREEDY makes k passes over the full ground set, providing
typically the best solution in terms of utility. The lazy eval-
uation eliminates a large fraction of the function evaluations
in each pass. Nonetheless, it is still computationally pro-
hibitive for large values of k.
In our experimental setup, we focus on three important
and classic machine learning applications: nonparametric
learning, exemplar-based clustering, and sensor placement.
Nonparametric Learning. Our first application is data
subset selection in nonparametric learning. We focus on the
special case of Gaussian Processes (GPs) below, but simi-
lar problems arise in large-scale kernelized SVMs and other
kernel machines. Let X
V
, be a set of random variables in-
dexed by the ground set V . In a Gaussian Process (GP)
we assume that every subset X
S
, for S = {e
1
, . . . , e
s
},
is distributed according to a multivariate normal distribu-
tion, i.e., P (X
S
= x
S
) = N (x
S
; µ
S
, Σ
S,S
), where µ
S
=
(µ
e
1
, . . . , µ
e
s
) and Σ
S,S
= [K
e
i
,e
j
](1 i, j k) are
the prior mean vector and prior covariance matrix, respec-
tively. The covariance matrix is given in terms of a posi-
tive definite kernel K, e.g., the squared exponential kernel
K
e
i
,e
j
= exp(−|e
i
e
j
|
2
2
/h
2
) is a common choice in prac-
tice. In GP regression, each data point e V is considered a
random variable. Upon observations y
A
= x
A
+ n
A
(where
n
A
is a vector of independent Gaussian noise with variance
σ
2
), the predictive distribution of a new data point e V is
a normal distribution P(X
e
| y
A
) = N (µ
e|A
, Σ
2
e|A
), where
µ
e|A
= µ
e
+ Σ
e,A
A,A
+ σ
2
I)
1
(x
A
µ
A
),
σ
2
e|A
= σ
2
e
Σ
e,A
A,A
+ σ
2
I)
1
Σ
A,e
. (2)
Note that evaluating (2) is computationally expensive as
it requires a matrix inversion. Instead, most efficient ap-
proaches for making predictions in GPs rely on choosing a
small – so called active set of data points. For instance, in
the Informative Vector Machine (IVM) we seek a summary
A such that the information gain, f(A) = I(Y
A
; X
V
) =
H(X
V
) H(X
V
|Y
A
) =
1
2
log det(I + σ
2
Σ
A,A
) is max-
imized. It can be shown that this choice of f is monotone
submodular (Krause and Guestrin 2005). For small values
of |A|, running LAZY-GREEDY is possible. However, we
see that as the size of the active set or summary A increases,
the only viable option in practice is STOCHASTIC-GREEDY.
In our experiment we chose a Gaussian kernel with h =
0.75 and σ = 1. We used the Parkinsons Telemonitor-
ing dataset (Tsanas et al. 2010) consisting of 5,875 bio-
medical voice measurements with 22 attributes from peo-
ple with early-stage Parkinsons disease. We normalized the
vectors to zero mean and unit norm. Fig. 1a and 1b com-
pare the utility and computational cost of STOCHASTIC-
GREEDY to the benchmarks for different values of k. For
THRESHOLD-GREEDY, different values of have been cho-
sen such that a performance close to that of LAZY-GREEDY
is obtained. Moreover, different values of p have been cho-
sen such that the cost of SAMPLE-GREEDY is almost equal
to that of STOCHASTIC-GREEDY for different values of .
As we can see, STOCHASTIC-GREEDY provides the closest
(practically identical) utility to that of LAZY-GREEDY with
much lower computational cost. Decreasing the value of ε
results in higher utility at the price of higher computational
cost. Fig. 1c shows the utility versus cost of STOCHASTIC-
GREEDY along with the other benchmarks for a fixed k =
200 and different values of . STOCHASTIC-GREEDY pro-
vides very compelling tradeoffs between utility and cost
compared to all benchmarks, including LAZY-GREEDY.
Exemplar-based clustering. A classic way to select a set
of exemplars that best represent a massive dataset is to solve
the k-medoid problem (Kaufman and Rousseeuw 2009) by
minimizing the sum of pairwise dissimilarities between ex-
emplars A and elements of the dataset V as follows: L(A) =
1
V
P
eV
min
vA
d(e, v), where d : V × V R is a dis-
tance function, encoding the dissimilarity between elements.
By introducing an appropriate auxiliary element e
0
we can
turn L into a monotone submodular function (Gomes and
Krause 2010): f (A) = L({e
0
}) L(A {e
0
}). Thus max-
imizing f is equivalent to minimizing L which provides a
very good solution. But the problem becomes computation-
ally challenging as the size of the summary A increases.
In our experiment we chose d(x, x
0
) = ||x x
0
||
2
for the
dissimilarity measure. We used a set of 10,000 Tiny Images
(Torralba, Fergus, and Freeman 2008) where each 32 × 32
RGB image was represented by a 3,072 dimensional vector.
We subtracted from each vector the mean value, normalized
it to unit norm, and used the origin as the auxiliary exem-
plar. Fig. 1d and 1e compare the utility and computational
cost of STOCHASTIC-GREEDY to the benchmarks for differ-
ent values of k. It can be seen that STOCHASTIC-GREEDY
outperforms the benchmarks with significantly lower com-
putational cost. Fig. 1f compares the utility versus cost of
different methods for a fixed k = 200 and various p and .
Similar to the previous experiment, STOCHASTIC-GREEDY
achieves near-maximal utility at substantially lower cost
compared to the other benchmarks.
Large scale experiment. We also performed a simi-
lar experiment on a larger set of 50,000 Tiny Images. For
this dataset, we were not able to run LAZY-GREEDY and
THRESHOLD-GREEDY. Hence, we compared the utility and
cost of STOCHASTIC-GREEDY with RANDOM-SELECTION
using different values of p. As shown in Fig. 1j and Fig. 1k,
STOCHASTIC-GREEDY outperforms SAMPLE-GREEDY in
terms of both utility and cost for different values of k. Fi-
nally, as Fig. 1l shows that STOCHASTIC-GREEDY achieves
the highest utility but performs much faster compare to
SAMPLE-GREEDY which is the only practical solution for
this larger dataset.
1815
Sensor Placement. When monitoring spatial phenomena,
we want to deploy a limited number of sensors in an area
in order to quickly detect contaminants. Thus, the problem
would be to select a subset of all possible locations A V to
place sensors. Consider a set of intrusion scenarios I where
each scenario i I defines the introduction of a contam-
inant at a specified point in time. For each sensor s S
and scenario i, the detection time, T (s, i), is defined as the
time it takes for s to detect i. If s never detects i, we set
T (s, i) = . For a set of sensors A, detection time for
scenario i could be defined as T (A, i) = min
sA
T (s, i).
Depending on the time of detection, we incur penalty π
i
(t)
for detecting scenario i at time t. Let π
i
() be the max-
imum penalty incurred if the scenario i is not detected at
all. Then, the penalty reduction for scenario i can be de-
fined as R(A, i) = π
i
() π
i
(T (A, i)). Having a proba-
bility distribution over possible scenarios, we can calculate
the expected penalty reduction for a sensor placement A as
R(A) =
P
i∈I
P (i)R(A, i). This function is montone sub-
modular (Krause et al. 2008) and for which the greedy algo-
rithm gives us a good solution. For massive data however,
we may need to resort to STOCHASTIC-GREEDY.
In our experiments we used the 12,527 node distribu-
tion network provided as part of the Battle of Water Sen-
sor Networks (BWSN) challenge (Ostfeld et al. 2008). Fig.
1g and 1h compare the utility and computational cost of
STOCHASTIC-GREEDY to the benchmarks for different val-
ues of k. It can be seen that STOCHASTIC-GREEDY out-
performs the benchmarks with significantly lower computa-
tional cost. Fig. 1i compares the utility versus cost of differ-
ent methods for a fixed k = 200 and various p and . Again
STOCHASTIC-GREEDY shows similar behavior to the previ-
ous experiments by achieving near-maximal utility at much
lower cost compared to the other benchmarks.
Conclusion
We have developed the first linear time algorithm
STOCHASTIC-GREEDY with no dependence on k
for cardinality constrained submodular maximization.
STOCHASTIC-GREEDY provides a 1 1/e approx-
imation guarantee to the optimum solution with only
n log
1
function evaluations. We have also demonstrated
the effectiveness of our algorithm through an extensive set
of experiments. As these show, STOCHASTIC-GREEDY
achieves a major fraction of the function utility with much
less computational cost. This improvement is useful even
in approaches that make use of parallel computing or de-
compose the submodular function into simpler functions for
faster evaluation. The properties of STOCHASTIC-GREEDY
make it very appealing and necessary for solving very
large scale problems. Given the importance of submodular
optimization to numerous AI and machine learning appli-
cations, we believe our results provide an important step
towards addressing such problems at scale.
Appendix, Analysis
The following lemma gives us the approximation guarantee.
Lemma 2. Given a current solution A, the expected
gain of STOCHASTIC-GREEDY in one step is at least
1
k
P
aA
\A
∆(a|A).
Proof. Let us estimate the probability that R(A
\A) 6= .
The set R consists of s =
n
k
log
1
random samples from
V \ A (w.l.o.g. with repetition), and hence
Pr[R (A
\ A) = ] =
1
|A
\ A|
|V \ A|
s
e
s
|A
\A|
|V \A|
e
s
n
|A
\A|
.
Therefore, by using the concavity of 1 e
s
n
x
as a function
of x and the fact that x = |A
\ A| [0, k], we have
Pr[R(A
\A) 6=]1e
s
n
|A
\A|
(1e
sk
n
)
|A
\ A|
k
.
Recall that we chose s =
n
k
log
1
, which gives
Pr[R (A
\ A) 6= ] (1 )
|A
\ A|
k
. (3)
Now consider STOCHASTIC-GREEDY: it picks an element
a R maximizing the marginal value ∆(a|A). This is
clearly as much as the marginal value of an element ran-
domly chosen from R (A
\ A) (if nonempty). Overall,
R is equally likely to contain each element of A
\ A, so
a uniformly random element of R (A
\ A) is actually a
uniformly random element of A
\ A. Thus, we obtain
E[∆(a|A)] Pr[R(A
\A) 6= ]×
1
|A
\ A|
X
aA
\A
∆(a|A).
Using (3), we conclude that E[∆(a|A)]
1
k
P
aA
\A
∆(a|A).
Now it is straightforward to finish the proof of Theo-
rem 1. Let A
i
= {a
1
, . . . , a
i
} denote the solution returned
by STOCHASTIC-GREEDY after i steps. From Lemma 2,
E[∆(a
i+1
|A
i
) | A
i
]
1
k
X
aA
\A
i
∆(a|A
i
).
By submodularity,
X
aA
\A
i
∆(a|A
i
) ∆(A
|A
i
) f(A
) f(A
i
).
Therefore,
E[f(A
i+1
) f(A
i
) | A
i
] = E[∆(a
i+1
|A
i
) | A
i
]
1
k
(f(A
) f(A
i
)).
By taking expectation over A
i
,
E[f(A
i+1
) f(A
i
)]
1
k
E[f(A
) f(A
i
)].
By induction, this implies that
E[f(A
k
)]
1
1
1
k
k
!
f(A
)
1 e
(1)
f(A
) (1 1/e )f(A
).
Acknowledgment. This research was supported by SNF
200021-137971, ERC StG 307036, a Microsoft Faculty Fel-
lowship, and an ETH Fellowship.
1816
Figure 1: Performance comparisons. a), d) g) and j) show the performance of all the algorithms for different values of k on
Parkinsons Telemonitoring, a set of 10,000 Tiny Images, Water Network, and a set of 50,000 Tiny Images respectively. b), e) h)
and k) show the cost of all the algorithms for different values of k on the same datasets. c), f), i), l) show the utility obtained
versus cost for a fixed k = 200.
1817
References
Badanidiyuru, A., and Vondr
´
ak, J. 2014. Fast algorithms for
maximizing submodular functions. In SODA, 1497–1514.
Badanidiyuru, A.; Mirzasoleiman, B.; Karbasi, A.; and
Krause, A. 2014. Streaming submodular optimization: Mas-
sive data summarization on the fly. In KDD.
Blelloch, G. E.; Peng, R.; and Tangwongsan, K. 2011.
Linear-work greedy parallel approximate set cover and vari-
ants. In SPAA.
Chierichetti, F.; Kumar, R.; and Tomkins, A. 2010. Max-
cover in map-reduce. In Proceedings of the 19th interna-
tional conference on World wide web.
Dasgupta, A.; Kumar, R.; and Ravi, S. 2013. Summarization
through submodularity and dispersion. In ACL.
Dueck, D., and Frey, B. J. 2007. Non-metric affinity propa-
gation for unsupervised image categorization. In ICCV.
El-Arini, K., and Guestrin, C. 2011. Beyond keyword
search: Discovering relevant scientific literature. In KDD.
El-Arini, K.; Veda, G.; Shahaf, D.; and Guestrin, C. 2009.
Turning down the noise in the blogosphere. In KDD.
Feige, U. 1998. A threshold of ln n for approximating set
cover. Journal of the ACM.
Golovin, D., and Krause, A. 2011. Adaptive submodularity:
Theory and applications in active learning and stochastic op-
timization. Journal of Artificial Intelligence Research.
Gomes, R., and Krause, A. 2010. Budgeted nonparametric
learning from data streams. In ICML.
Guillory, A., and Bilmes, J. 2011. Active semi-supervised
learning using submodular functions. In UAI. AUAI.
Kaufman, L., and Rousseeuw, P. J. 2009. Finding groups
in data: an introduction to cluster analysis, volume 344.
Wiley-Interscience.
Kempe, D.; Kleinberg, J.; and Tardos, E. 2003. Maximizing
the spread of influence through a social network. In Pro-
ceedings of the ninth ACM SIGKDD.
Krause, A., and Guestrin, C. 2005. Near-optimal nonmyopic
value of information in graphical models. In UAI.
Krause, A., and Guestrin, C. 2011. Submodularity and
its applications in optimized information gathering. ACM
Transactions on Intelligent Systems and Technology.
Krause, A.; Leskovec, J.; Guestrin, C.; VanBriesen, J.; and
Faloutsos, C. 2008. Efficient sensor placement optimiza-
tion for securing large water distribution networks. Journal
of Water Resources Planning and Management 134(6):516–
526.
Kumar, R.; Moseley, B.; Vassilvitskii, S.; and Vattani, A.
2013. Fast greedy algorithms in mapreduce and streaming.
In SPAA.
Leskovec, J.; Krause, A.; Guestrin, C.; Faloutsos, C.; Van-
Briesen, J.; and Glance, N. 2007. Cost-effective outbreak
detection in networks. In KDD, 420–429. ACM.
Lin, H., and Bilmes, J. 2011. A class of submodular func-
tions for document summarization. In North American chap-
ter of the Assoc. for Comp. Linguistics/Human Lang. Tech.
Minoux, M. 1978a. Accelerated greedy algorithms for max-
imizing submodular set functions. Optimization Techniques,
LNCS 234–243.
Minoux, M. 1978b. Accelerated greedy algorithms for max-
imizing submodular set functions. In Proc. of the 8th IFIP
Conference on Optimization Techniques. Springer.
Mirzasoleiman, B.; Karbasi, A.; Sarkar, R.; and Krause, A.
2013. Distributed submodular maximization: Identifying
representative elements in massive data. In Neural Infor-
mation Processing Systems (NIPS).
Nemhauser, G. L., and Wolsey, L. A. 1978. Best algorithms
for approximating the maximum of a submodular set func-
tion. Math. Oper. Research.
Nemhauser, G. L.; Wolsey, L. A.; and Fisher, M. L. 1978.
An analysis of approximations for maximizing submodular
set functions - I. Mathematical Programming.
Ostfeld, A.; Uber, J. G.; Salomons, E.; Berry, J. W.; Hart,
W. E.; Phillips, C. A.; Watson, J.-P.; Dorini, G.; Jonker-
gouw, P.; Kapelan, Z.; et al. 2008. The battle of the water
sensor networks (bwsn): A design challenge for engineers
and algorithms. Journal of Water Resources Planning and
Management 134(6):556–568.
Rodriguez, M. G.; Leskovec, J.; and Krause, A. 2012. Infer-
ring networks of diffusion and influence. ACM Transactions
on Knowledge Discovery from Data 5(4):21:1–21:37.
Sipos, R.; Swaminathan, A.; Shivaswamy, P.; and Joachims,
T. 2012. Temporal corpus summarization using submodular
word coverage. In CIKM.
Torralba, A.; Fergus, R.; and Freeman, W. T. 2008. 80 mil-
lion tiny images: A large data set for nonparametric object
and scene recognition. IEEE Trans. Pattern Anal. Mach. In-
tell.
Tsanas, A.; Little, M. A.; McSharry, P. E.; and Ramig, L. O.
2010. Enhanced classical dysphonia measures and sparse re-
gression for telemonitoring of parkinson’s disease progres-
sion. In IEEE Int. Conf. Acoust. Speech Signal Process.
Wei, K.; Liu, Y.; Kirchhoff, K.; and Bilmes, J. 2013. Using
document summarization techniques for speech data subset
selection. In HLT-NAACL, 721–726.
Wei, K.; Iyer, R.; and Bilmes, J. 2014. Fast multi-stage
submodular maximization. In ICML.
1818