22.01.2015 Views

High Frequency Traders, News and Volatility - HEC Paris

High Frequency Traders, News and Volatility - HEC Paris

High Frequency Traders, News and Volatility - HEC Paris

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Fast <strong>and</strong> Slow Informed Trading ∗<br />

Ioanid Roşu †<br />

September 12, 2014<br />

Abstract<br />

This paper develops a model in which traders continuously process costly information,<br />

<strong>and</strong> may differ in their trading speed. In equilibrium, (i) the fast informed traders (FTs)<br />

generate most of the volatility <strong>and</strong> trading volume in the market; (ii) they must use their<br />

information quickly, as the value of information decays fast; (iii) when the number of FTs<br />

increases, price volatility, trading volume <strong>and</strong> liquidity increase; (iv) when the FTs’ signal<br />

precision increases, volatility <strong>and</strong> volume increase, while liquidity decreases. FTs can<br />

manage their inventory almost infinitely fast, but only in the presence of slower informed<br />

traders.<br />

Keywords:<br />

Trading volume, inventory, volatility, high frequency trading, price impact,<br />

Kyle model.<br />

∗ Earlier versions of this paper circulated under the title “<strong>High</strong> <strong>Frequency</strong> <strong>Traders</strong>, <strong>News</strong> <strong>and</strong> <strong>Volatility</strong>.”<br />

The author thanks Kerry Back, Laurent Calvet, Thierry Foucault, Johan Hombert, Pete Kyle, Stefano Lovo,<br />

Victor Martinez, Dimitri Vayanos, Jiang Wang; finance seminar participants at <strong>HEC</strong> <strong>Paris</strong>, Durham, Leicester,<br />

<strong>Paris</strong> Dauphine, Copenhagen, Madrid Carlos III, ESSEC; <strong>and</strong> conference participants at the European Finance<br />

Association meetings in Lugano, American Finance Association meetings in San Diego, Society for Advancement<br />

of Economic Theory in Portugal, Central Bank Microstructure Conference in Norway, Market Microstructure<br />

Many Viewpoints Conference in <strong>Paris</strong> for valuable comments.<br />

† <strong>HEC</strong> <strong>Paris</strong>, Email: rosu@hec.fr.<br />

1


1 Introduction<br />

Today’s markets are increasingly characterized by the continuous arrival of vast amounts of<br />

information. A media article about high frequency trading reports on the hedge fund firm<br />

Citadel: “Its market data system, for example, contains roughly 100 times the amount of information<br />

in the Library of Congress. [...] The signals, or alphas, that prove to have predictive<br />

power are then translated into computer algorithms, which are integrated into Citadel’s master<br />

source code <strong>and</strong> electronic trading program.” (“Man vs. Machine,” CNBC.com, September<br />

13 th 2010). The sources of information from which traders obtain these signals usually include<br />

company-specific news <strong>and</strong> reports, economic indicators, stock indexes, prices of other securities,<br />

prices on various other trading platforms, limit order book changes, as well as various<br />

“machine readable news” <strong>and</strong> even “sentiment” indicators. 1<br />

At the same time, financial markets have seen in recent years the spectacular rise of<br />

algorithmic trading, <strong>and</strong> in particular of high frequency trading. 2<br />

This coincidental arrival<br />

raises the question whether or not at least some of the HFTs do process information <strong>and</strong> trade<br />

very quickly in order to take advantage of their speed <strong>and</strong> superior computing power. Recent<br />

empirical evidence suggests that this is indeed the case. 3<br />

But, despite the large role played<br />

by high frequency traders (HFTs) in the current financial l<strong>and</strong>scape, there has been relatively<br />

little progress in explaining their strategies in connection with information processing.<br />

particular, we list a set of questions that a theoretical model about informed HFTs should<br />

address: (i) What are the optimal trading strategies of HFTs who process information (ii)<br />

Why do HFTs account for such a large share of the trading volume (iii) What explains the<br />

race for speed among HFTs What are the effects of HFT on measures of market quality,<br />

such as liquidity <strong>and</strong> price volatility (v) What explains the low inventory of at least some of<br />

the HFTs 4<br />

In this paper, we provide a theoretical model of informed trading which parsimoniously<br />

addresses the questions mentioned above. In our model, informed traders called “speculators”<br />

(i) learn gradually about the value of a risky asset (i.e., obtain “signals”), (ii) pay an<br />

information processing cost per individual signal per trading round, <strong>and</strong> (iii) may differ in<br />

their speed of information processing. Our main results are: (a) there is an exponential decay<br />

1 “Math-loving traders are using powerful computers to speed-read news reports, editorials, company Web<br />

sites, blog posts <strong>and</strong> even Twitter messages—<strong>and</strong> then letting the machines decide what it all means for the<br />

markets.” (“Computers That Trade on the <strong>News</strong>,” New York Times, December 22 nd 2010).<br />

2 Hendershott, Jones, <strong>and</strong> Menkveld (2011) report that from a starting point near zero in the mid-1990s,<br />

high frequency trading rose to as much as 73% of trading volume in the United States in 2009. Chaboud,<br />

Chiquoine, Hjalmarsson, <strong>and</strong> Vega (2014) consider various foreign exchange markets <strong>and</strong> find that starting<br />

from essentially zero in 2003, algorithmic trading rose by the end of 2007 to approximately 60% of the trading<br />

volume for the euro-dollar <strong>and</strong> dollar-yen markets, <strong>and</strong> 80% for the euro-yen market.<br />

3 See Brogaard, Hendershott, <strong>and</strong> Riordan (2014), Baron, Brogaard, <strong>and</strong> Kirilenko (2014), Kirilenko, Kyle,<br />

Samadi, <strong>and</strong> Tuzun (2014), Hirschey (2013), Benos <strong>and</strong> Sagade (2013).<br />

4 In their analysis of the Flash Crash of May 2010, Kirilenko, Kyle, Samadi, <strong>and</strong> Tuzun (2014) define HFTs<br />

by both high trading volume <strong>and</strong> low inventory, both at the end of the day <strong>and</strong> intraday.<br />

In<br />

2


of the benefits that arise from each signal, <strong>and</strong> this decay is faster with more competition<br />

among speculators; thus, information decays quickly, which generates a need for speed; (b) as<br />

a consequence, speculators trade only on their most recent signals; (c) by focusing disproportionately<br />

on their most recent signals, speculators generate a very large trading volume;<br />

(d) such a large trading volume is possible because competition among speculators ultimately<br />

makes the market more efficient <strong>and</strong> liquid; (e) if, in addition, certain traders want to manage<br />

their inventories, they can reduce it to zero almost infinitely fast (in “real time”) <strong>and</strong> still<br />

make a profit, but only if other speculators trade more slowly.<br />

More specifically, we setup our model as in Kyle (1985), but with multiple informed traders<br />

<strong>and</strong> a changing fundamental value, v t , modeled as a diffusion process. The speculators observe<br />

identical signals of the form ∆v t + noise, but some speculators may observe the signal with<br />

a lag.<br />

Along with the signal, we assume that the speculator must also learn the public’s<br />

expectation about the signal, so that he can trade on the difference—the “non-stale” part of<br />

the information. 5 As the public’s expectation of a signal changes after each trading round, we<br />

assume the speculator must pay the cost each time the signal is used. Thus, signals must be<br />

processed individually both in the cross section (different increments) <strong>and</strong> in the time series<br />

(different trading rounds). 6<br />

The individual signal processing cost may be in practice very<br />

small, but as long as it is not zero, with an exponential decay of information the cost poses a<br />

constraint that soon becomes binding.<br />

Thus, any positive signal processing cost is enough to stop the Holden <strong>and</strong> Subrahmanyam<br />

(1992) phenomenon, by which N > 1 informed traders with identical information release the<br />

information so fast that no equilibrium exists in the limit when the number of trading periods<br />

approaches infinity. In our model, the information cost “tames” the equilibrium by making<br />

the speculators optimally forget signals which are older than a threshold.<br />

The key intuition of our model can be grasped from a simple version of the model in which<br />

N equally fast speculators use only their most recent signal in deciding how to trade. Then, in<br />

each trading round t, the N speculators submit their market orders along with noise traders.<br />

The risk neutral competitive dealer sets the price as a linear function of the aggregate order<br />

size. Then, as in the Kyle (1985) model, each speculator faces a Cournot-type problem: he<br />

gets benefits that are proportional to the quantity submitted (which in turn is proportional to<br />

his signal), but faces a quadratic cost (the price impact) which increases with the aggregate<br />

quantity. As in the Cournot model, each speculator behaves in equilibrium like a monopolist<br />

against the residual dem<strong>and</strong> curve—or, more appropriately in our context, like a monopsonist<br />

5 In fact, even observing the signal ∆v for the first time is in practice likely to involve removing its public<br />

expectation, so that only its orthogonal (non-stale) part can be used.<br />

6 This assumption implicitly means that traders cannot costlessly rely on public signals such as the price to<br />

shortcut the learning process. In the Kyle (1985) model, the price is the unique source of information about<br />

the public’s expectations, <strong>and</strong> the speculator knows his information to be superior to that of the public’s along<br />

all dimensions. If, instead, as it is likely in practice, an asset’s value has other components that the market<br />

learns about, a speculator is potentially adversely selected when using the public price to decide his strategy.<br />

3


against the residual supply curve.<br />

As a result, each speculator trades such that the price<br />

impact of his order is on average 1/(N + 1) of his signal. That brings the expected aggregate<br />

price impact to N/(N + 1) of the signal, <strong>and</strong> leaves on average only 1/(N + 1) of the signal<br />

unknown to the dealer.<br />

Figure 1: Profit from Lagged Signals. The figure plots the percentage of a speculator’s<br />

profit from each his lagged signals when there is competition among N ∈ {1, 2, 3, 5, 20, 100} identical<br />

speculators. In these examples, the speculators can trade up to m = 5 lagged signals.<br />

100%<br />

N = 1<br />

100%<br />

N = 2<br />

100%<br />

N = 3<br />

80%<br />

80%<br />

80%<br />

profit<br />

60%<br />

40%<br />

profit<br />

60%<br />

40%<br />

profit<br />

60%<br />

40%<br />

20%<br />

20%<br />

20%<br />

0%<br />

0 1 2 3 4 5<br />

lag<br />

0%<br />

0 1 2 3 4 5<br />

lag<br />

0%<br />

0 1 2 3 4 5<br />

lag<br />

100%<br />

N = 5<br />

100%<br />

N = 20<br />

100%<br />

N = 100<br />

80%<br />

80%<br />

80%<br />

profit<br />

60%<br />

40%<br />

profit<br />

60%<br />

40%<br />

profit<br />

60%<br />

40%<br />

20%<br />

20%<br />

20%<br />

0%<br />

0 1 2 3 4 5<br />

lag<br />

0%<br />

0 1 2 3 4 5<br />

lag<br />

0%<br />

0 1 2 3 4 5<br />

lag<br />

In our dynamic model, the outcome of the trading game is more complicated, but the<br />

basic intuition of the one-period model carries through. Namely, as soon as speculators learn<br />

new information, they must trade very quickly, because even after one trading round only<br />

a small fraction of the signal is left available for subsequent trades. In the next round, the<br />

speculators exploit the new signal in a similar fashion, but in addition they can use old signals<br />

as well. However, if a speculator uses an old signal, the same revelation mechanism only leaves<br />

a small fraction of the small fraction already left. Hence, the benefits from repeatedly trading<br />

on a particular signal decrease at a rate equal to 1/(N + 1) with each trading round. Thus,<br />

with benefits decreasing at such a high rate, even a tiny information processing cost is enough<br />

to ensure that speculators only use their most recent signals. 7<br />

Figure 1 provides a graphic<br />

illustration of this phenomenon when N = 1, 2, 3, 5, 20, 100 identical speculators compete for<br />

information <strong>and</strong> use only their 5 most recent signals. Indeed, one sees that the fraction of<br />

7 Without information processing costs, it turns out that these signals are used ad infinitum, <strong>and</strong> the<br />

equilibrium unravels. This is in essence the Holden <strong>and</strong> Subrahmanyam (1992) effect: as the number of trading<br />

rounds becomes very high, information revelation is almost instantaneous, the market is almost infinitely liquid,<br />

<strong>and</strong> the equilibrium breaks down in the limit.<br />

4


speculator profits that occur from their most recent signal is high even when N is relatively<br />

low.<br />

The same intuition works when comparing the fastest traders (FTs)—i.e., the speculators<br />

that observe new signals immediately,—with the slower traders (STs), that observe signals<br />

with a lag. When the number of FTs is sufficiently high, the profits of STs are very small.<br />

This is because STs can trade only on old signals, <strong>and</strong>, as we discussed before, these signals<br />

lose quickly most of their value. In addition, STs must compete with the FTs, who also trade<br />

on their older signals. This further decreases the profits of STs. 8 Thus, FTs dominate trading<br />

volume, hence our findings are driven mostly by the activities of FTs.<br />

Once the basic intuition that drives our model is clear, our main results follow in a relatively<br />

straightforward manner. Nevertheless, it is worth noting that these results are not obvious<br />

ex ante. In particular, the literature on informed trading has had difficulties attributing the<br />

high trading volume observed in practice to informed traders. Indeed, Kyle (1985), Back <strong>and</strong><br />

Pedersen (1998), or Back, Cao, <strong>and</strong> Willard (2000), as the number of trading rounds is high<br />

(<strong>and</strong> in the continuous time limit) the speculators generate only a tiny fraction of the trading<br />

volume, as speculators trade smoothly on their long-lived information. By contrast, in our<br />

model FTs trade aggressively only on their most recent signals <strong>and</strong> ignore information that is<br />

too old. As a result, informed trading in our model occurs in a noisy fashion, essentially as<br />

noisy as the information received by the FTs. Thus, paradoxically, an information processing<br />

cost dissuades even a single speculator from trading smoothly on long-lived information, <strong>and</strong><br />

instead drives FTs to reveal a large part of their signals almost instantly.<br />

Quantitatively,<br />

the aggregate FT trading volume can be shown to increase almost linearly in N, <strong>and</strong> the FT<br />

participation rate is of the order of N/(N + 1) as a fraction of the overall trading volume.<br />

The reason why FT trading volume can be so high is that in equilibrium the market is very<br />

liquid, <strong>and</strong> FTs can trade without much price impact. Indeed, quick information revelation<br />

makes the market close to strong-form efficient <strong>and</strong> hence very liquid. The dealer knows that<br />

most of the information is eventually released by the speculators, <strong>and</strong> endogenously chooses a<br />

small price impact for trading. This creates an amplification mechanism that provides ample<br />

room for trading.<br />

The effect of FTs on volatility is more muted. In our model, it turns out that the price<br />

volatility is bounded above by the fundamental volatility of the asset.<br />

Nevertheless, price<br />

volatility increases with FT competition <strong>and</strong> signal precision. The intuition is that as FTs<br />

trade more aggressively, they release most of their information, so that in the limit the market<br />

becomes strong-form efficient, <strong>and</strong> price volatility equals fundamental volatility.<br />

One of the most puzzling aspects of high frequency trading is the quick turnaround of HFT<br />

positions, i.e., the maintenance of very low HFT inventories. Kirilenko, Kyle, Samadi, <strong>and</strong><br />

8 Baron, Brogaard, <strong>and</strong> Kirilenko (2014) study the profits of HFTs <strong>and</strong> find out that, in agreement with<br />

our model, HFT profits are concentrated among a small number of incumbents.<br />

5


Tuzun (2014) report that HFTs in their sample liquidate 0.5% of their aggregate inventories<br />

on average each second, implying HFT inventories have an AR(1) half-live of a little over<br />

2 minutes. To study this problem, we allow a subset of the speculators—called HFTs—to<br />

manage their inventory. We find that in our model HFTs can manage their inventories almost<br />

infinitely fast, while still maintaining positive average profits. This, however, can be done only<br />

if there exist “slow” traders. By “slow” traders we mean either (i) STs, who are genuinely slow<br />

<strong>and</strong> can only trade on older information; or (ii) FTs who also trade on their older information<br />

(lagged signals) because it is optimal for them to do so. To underst<strong>and</strong> the intuition for this<br />

result, consider a FT who trades on his information. In the absence of inventory management,<br />

the inventory of the FT represents an accumulation of signals which are positively correlated<br />

with the asset value. By trading to lower his inventory, the FT trades against his information,<br />

<strong>and</strong> if he does this too fast, he may eventually trade at a loss. However, we show that by<br />

selling the asset at the same time that other speculators are buying, the FT may lower his<br />

price impact, <strong>and</strong>, if the inventory management is done properly, the FT can maintain positive<br />

profits. This is true even if the chosen mean-reversion rate is virtually infinite.<br />

Related Literature<br />

Our main intuition regarding the quick information decay when informed traders compete<br />

is also present in Foster <strong>and</strong> Viswanathan (1996). They find that informed traders with<br />

correlated signals engage in a “rat race,” <strong>and</strong> reveal their common information very fast.<br />

Furthermore, in the extreme case when informed traders have identical signals, Holden <strong>and</strong><br />

Subrahmanyam (1992) find that, as the number of trading rounds increases, the signal is<br />

revealed almost instantaneously, such that in the continuous time limit there cannot be any<br />

equilibrium in smooth strategies, as in Kyle (1985). If the signals are not identical, however,<br />

Foster <strong>and</strong> Viswanathan (1996) show that the rat race is followed by a “waiting game,” in which<br />

traders reveal their information very slowly, as they are likely to be on the opposite side of a<br />

trade. Back, Cao, <strong>and</strong> Willard (2000) confirm that the equilibrium in smooth strategies with<br />

the rate race <strong>and</strong> waiting game survives in a continuous time setup. Our paper contributes to<br />

this literature by showing that even a tiny information processing cost slows down the rat race<br />

enough so that the equilibrium survives even when the number of trading rounds becomes very<br />

large (at the “high frequencies”). Nevertheless, in our equilibrium strategies are not smooth as<br />

in Kyle (1985), but they are volatile <strong>and</strong> place a large weight on traders’ most recent signals.<br />

In the papers mentioned above, the signals are static, as they are received only at the<br />

beginning of the trading game. A more recent literature investigates the effect of dynamic<br />

signals, by which traders gradually learn over time. Back <strong>and</strong> Pedersen (1998) extend the<br />

basic setup of Kyle (1985) to allow for dynamic signals, <strong>and</strong> find that in equilibrium the<br />

insider continues to have smooth strategies. In their model, as in Kyle (1985), the insider<br />

6


places the same weight on all the past signals <strong>and</strong> thus trades on the aggregate signal, which<br />

is his forecast of the fundamental value.<br />

In Chau <strong>and</strong> Vayanos (2008), the strategies are<br />

similar, but at high frequencies (i.e., when the time between trading rounds becomes very<br />

small) the insider trades with almost infinite weight <strong>and</strong> the market approaches strong-form<br />

efficiency. 9 This perhaps should not be surprising if one interprets Chau <strong>and</strong> Vayanos (2008)<br />

as a stationary version of Back <strong>and</strong> Pedersen (1998). Indeed, in Back <strong>and</strong> Pedersen (1998),<br />

as in Kyle (1998), the market approaches strong-form efficiency towards the liquidation date,<br />

<strong>and</strong> the insider trades almost with infinite weight. 10<br />

By contrast, in our paper speculators<br />

trade only on their most recent signals, <strong>and</strong> the market is not strong-form efficient (although it<br />

becomes so in the limit when the number of speculators is large). One other paper that breaks<br />

the symmetry among past signals is Foucault, Hombert, <strong>and</strong> Roşu (2013). In their model, a<br />

single speculator receives a signal one instant before the public, <strong>and</strong> as a result, trades with<br />

a large weight on the last signal (the “news”). In their model, the value of information is lost<br />

quickly because of public news. In our model, it decays quickly because of competition among<br />

informed traders.<br />

Our paper is part of a growing theoretical literature on <strong>High</strong> <strong>Frequency</strong> Trading. 11<br />

much of this literature, it is the speed difference that has a large effect in equilibrium. The<br />

usual model setup has certain traders who are faster in taking advantage of an opportunity<br />

that disappears quickly. As a result, traders enter into a winner-takes-all contest, in which<br />

even the smallest difference in speed has a large effect on profits. (See for instance the model<br />

of “sniping” of Budish, Cramton, <strong>and</strong> Shim (2014), or the model of news anticipation of<br />

Foucault, Hombert, <strong>and</strong> Roşu (2013).) By contrast, our main results remain true even if all<br />

informed traders have the same speed. This is because in our model the need for speed arises<br />

endogenously, from competition among informed traders. In our model, being “slow” simply<br />

means trading on older signals. Since in equilibrium speculators also use older signals (the<br />

unanticipated part of the signals, to be precise), in some sense all traders are slow as well.<br />

Nevertheless, it is true that a genuinely slower trader makes less money in our model, since<br />

he can only trade on older information that has already lost much of its value.<br />

The paper is organized as follows. Section 2 describes the model. Section 3 solves the<br />

model with two classes of informed traders: fast <strong>and</strong> slow.<br />

In<br />

Section 4 discusses empirical<br />

implications of the model regarding the effect of fast traders on various measures of market<br />

quality. Section 5 analyzes the inventory management of speculators. Section 6 discusses the<br />

equilibrium in the general case, as well as robust trading strategies, which are strategies that<br />

9 Li (2012) confirms Chau <strong>and</strong> Vayanos (2008) in an extension with multiple traders. He also finds that<br />

because of competition informed trading volume increases by an order of magnitude.<br />

10 Caldentey <strong>and</strong> Stacchetti (2010) find that a r<strong>and</strong>om liquidation deadline has essentially the same effect as<br />

the stationarity in the Chau <strong>and</strong> Vayanos (2008) model.<br />

11 See Budish, Cramton, <strong>and</strong> Shim (2014), Biais, Foucault, <strong>and</strong> Moinas (2014), Foucault, Hombert, <strong>and</strong> Roşu<br />

(2013), Pagnotta <strong>and</strong> Philippon (2013), Aït-Sahalia <strong>and</strong> Saglam (2014), Li (2014), Weller (2013), Hoffmann<br />

(2013), Cartea <strong>and</strong> Penalva (2012), Jovanovic <strong>and</strong> Menkveld (2012).<br />

7


simplicity 12 T = 1. (1)<br />

perform well under alternative specifications of the model. Section 7 concludes. All proofs are<br />

in the Appendix or the Internet Appendix.<br />

2 The Model<br />

2.1 Setup<br />

Trading for a risky asset takes place continuously over the time interval [0, T ], where for<br />

Trading occurs at intervals of length dt apart. Throughout the text, we refer to dt as representing<br />

either one period, or one trading round. The liquidation value of the asset is<br />

v T =<br />

∫ T<br />

0<br />

dv t , with dv t = σ v dB v t , (2)<br />

where Bt<br />

v is a Brownian motion, <strong>and</strong> σ v > 0 is a constant called the fundamental volatility.<br />

We interpret v T as the “long-run” value of the asset; in the high frequency world, this can<br />

be taken to be the asset value at the end of the trading day. The increments dv t are then<br />

the short term changes in value due to the arrival of new information. The risk-free rate is<br />

assumed to be zero.<br />

There are three types of market participants: (a) N ≥ 1 risk neutral speculators, who<br />

observe the flow of information at different speeds, as described below; (b) noise traders; <strong>and</strong><br />

(c) one competitive risk neutral dealer, who sets the price at which trading takes place.<br />

Information. At t = 0, there is no information asymmetry between the speculators <strong>and</strong><br />

the dealer. Subsequently, each speculator receives the following flow of signals:<br />

ds t = dv t + dη t , with dη t = σ η dB η t , (3)<br />

where t ∈ (0, T ] <strong>and</strong> B η t<br />

by<br />

is a Brownian motion independent from all other variables. Denote<br />

w t = E(v T<br />

∣ ∣ {sτ } τ≤t ) (4)<br />

the expected value conditional on the information flow until t. We call w t the value forecast,<br />

or simply forecast. Since there is no information asymmetry at t = 0, we have w 0 = 0. The<br />

increment of w t is<br />

dw t = ads t , with a =<br />

σ 2 v<br />

σv 2 + ση<br />

2 , (5)<br />

12 To eliminate confusion with later notation, we often use T instead of 1. This way, we can denote t − dt<br />

by t − 1 without much confusion.<br />

8


where the parameter a is called signal precision. If we denote by σ w the instantaneous volatility<br />

of w t , we compute<br />

σ w =<br />

( ) Var(dwt ) 1/2<br />

= a<br />

dt<br />

( ) Var(dst ) 1/2<br />

=<br />

dt<br />

σ 2 v<br />

√<br />

σ 2 v + σ 2 η<br />

. (6)<br />

As the forecast volatility σ w <strong>and</strong> the signal precision a are monotonic transformations of each<br />

other, we use these two notions interchangeably when deriving empirical implications.<br />

Speculators obtain their signal with a lag l = 0, 1, . . .. Thus, we define a l-speculator to<br />

be a trader who at t ∈ (0, T ] observes the lagged signal ds t−ldt , from l periods before. For<br />

each lag l = 0, 1, 2, . . ., denote by N l the number of l-speculators. For simplicity of notation,<br />

henceforth we use the following convention:<br />

Notation for trading times: t − l instead of t − ldt. (7)<br />

For instance, we write ds t−l instead of ds t−ldt .<br />

Trading. At each t ∈ (0, T ], denote by dx i t the market order submitted by speculator<br />

i = 1, . . . , N at t, <strong>and</strong> by du t the market order submitted by the noise traders, which is of<br />

the form du t = σ u dBt u , where Bt<br />

u is a Brownian motion independent from all other variables.<br />

Then, the aggregate order flow executed by the dealer at t is<br />

dy t =<br />

N∑<br />

dx i t + du t . (8)<br />

i=1<br />

Because the dealer is risk neutral <strong>and</strong> competitive, she executes the order flow at a price equal<br />

to her expectation of the liquidation value conditional on her information. Let I t = {y τ } τ


which the l-speculator’s strategy is linear in the unpredictable parts of his signals: 13<br />

dx t = γ l,t (dw t−l − z t−l,t ) + γ l+1,t (dw t−l−1 − z t−l−1,t ) + · · · , (11)<br />

where z t−k,t is the dealer’s expectation of dw t−k conditional on her information set at t:<br />

z t−k,t = E ( dw t−k | I t<br />

)<br />

. (12)<br />

We further assume that the speculators take the covariance structure of z t−k,t to be fixed, as<br />

part of the dealer’s pricing rules. 14<br />

A linear equilibrium is such that: (i) at every date t, each speculator’s trading strategy (11)<br />

maximizes his expected trading profit (10) given the dealer’s pricing policy, <strong>and</strong> (ii) the dealer’s<br />

pricing policy given by (9) <strong>and</strong> (12) is consistent with the equilibrium speculators’ trading<br />

strategies.<br />

Information Processing Cost. There is a positive information processing cost c per<br />

individual piece of information <strong>and</strong> per unit of time. Formally, the total information cost from<br />

date τ onwards is:<br />

C τ =<br />

∫ T<br />

τ<br />

∞∑<br />

1 γk,t ≠0 c dt, (13)<br />

where 1 P is the indicator function, which equals 1 if P is true, <strong>and</strong> 0 otherwise.<br />

k=l<br />

We do not model the exact nature of the speculators’ signals <strong>and</strong> the information processing<br />

costs. But, intuitively, an individual information processing cost per each trading round<br />

is plausible for several reasons. First, as explained in the Introduction, in today’s markets<br />

traders use many high frequency sources of information to obtain their signals. In practice,<br />

processing these signals amounts not only to estimating the absolute magnitude of the signal,<br />

but also to measuring it relative to a public expectation. 15<br />

Often, however, these expectations<br />

are not readily available to the traders, <strong>and</strong> must be inferred either from observing the<br />

order flow (by estimating how much information has already been released both via own <strong>and</strong><br />

competitors’ trades), <strong>and</strong> from the analysis of public news (some signals might have already<br />

become public in the meantime). Thus, extracting the unpredictable (non-stale) part for each<br />

piece of information, which we model as the orthogonal increment dw t , is a non-trivial task<br />

which in practice is likely to require fast <strong>and</strong> accurate processing abilities. And once dw t is<br />

computed <strong>and</strong> used to trade, speculators must subsequently use only the non-stale part, which<br />

13 This type of strategies occur in virtually all models of Kyle (1985) type. In the Internet Appendix, we<br />

show in a discrete time version of the model that these type of strategies arise naturally in equilibrium.<br />

14 For more discussion regarding this assumption, see the proof of Theorem 1. Also, in Section 6 we explore<br />

an alternative assumption by which speculators’ trading strategies to affect the covariances of z t−k,t . We find<br />

that the overall effect is very small, so that our main results remain qualitatively the same.<br />

15 For instance, one often measures earnings surprises by comparing the absolute magnitude of the earnings<br />

announcement with the analyst consensus estimate.<br />

10


keeps changing after each trade. 16 Given all these reasons, it is plausible to expect that in<br />

practice an individual information processing per unit of time is necessary.<br />

Note that the assumption of an individual processing cost means implicitly that speculators<br />

cannot simply rely on free public signals, such as the price, to shortcut the learning process.<br />

This is because in reality prices may contain other relevant information about the fundamental<br />

value, along which the speculators are adversely selected. We formalize this intuition in<br />

Section 6.3, <strong>and</strong> introduce an orthogonal dimension of the fundamental value. We show that<br />

trading strategies that rely on prices make an average loss, while trading as in our model<br />

remains profitable.<br />

2.2 The Model of Trading with m Lagged Signals<br />

Solving for the equilibrium in the general setup of Section 2.1 turns out to be challenging.<br />

There are two main reasons for this. First, the model does not specify how many past signals<br />

can be processed by each speculator, <strong>and</strong> this can in principle lead to multiple equilibria.<br />

Second, as the signals are processed individually, the complexity of the problem increases<br />

with the number of signals;<br />

Thus, we propose the following solution technique. We consider the same setup in Section<br />

2.1, but instead of introducing an information processing cost, we set the cost to zero, <strong>and</strong><br />

consider only trading strategies in which signals are used for at most an exogenously chosen<br />

number of trading periods, m.<br />

Assumption 1. The speculators’ trading strategies can involve only their current signal<br />

<strong>and</strong> their most recent past m signals. More precisely, at time t, each speculator’s trading<br />

strategy must be of the form<br />

dx t = γ 0,t (dw t − z t,t ) + γ 1,t (dw t−1 − z t−1,t ) + γ m,t (dw t−m − z t−m,t ), (14)<br />

where z t−k,t is the dealer’s expectation of dw t−k , given her information.<br />

information processing cost, i.e., c = 0.<br />

Also, there is no<br />

To justify this assumption, we argue that a positive signal processing cost c is economically<br />

equivalent to Assumption 1. Indeed, as we show later in Section 6, the benefit of each lagged<br />

signal decreases exponentially with the lag. Then, given a signal processing cost c, the benefit<br />

of each individual signal must decay below c after a large enough number of lags. At that<br />

point, it is not worth using that signal anymore. This determines the number m of past signals<br />

that can be profitably used.<br />

16 Furthermore, we show below that (i) the non-stale part changes by different amounts for the different<br />

pieces of information, (ii) the optimal weights (γ k ) are different even when speculators have the same speed,<br />

(iii) the optimal weights depend on the signal precision, <strong>and</strong> while in the model this precision is constant, in<br />

practice it may different from one signal to the next.<br />

11


Throughout the paper, we refer to the modified setup in this section as the model of trading<br />

with m lagged signals. Because the solution of this model is more involved in the general case,<br />

we postpone its discussion until Section 6. However, as we have mentioned already, one of the<br />

key properties of the general equilibrium is that the benefits of lagged information decay very<br />

quickly. Thus, it turns out that most of the results in the general case can already be obtained<br />

in the model of trading with m = 1 lagged signals. Thus, in the next section we discuss this<br />

case in detail, <strong>and</strong> we see that its solution can be expressed in closed form.<br />

3 Equilibrium with Fast <strong>and</strong> Slow <strong>Traders</strong><br />

In this section, we analyze the model of trading from Section 2.2 in which speculators can<br />

trade with only one lagged signal (m = 1). This means that at time t speculators use only (i)<br />

their most recent signal, dw t , if this is observed; <strong>and</strong> (ii) the signal with one lag, dw t−1 (recall<br />

that t − 1 is notation for t − dt). Thus, there are two types of speculators:<br />

• Fast <strong>Traders</strong>, or FTs, who observe at t both dw t <strong>and</strong> dw t−1 ;<br />

• Slow <strong>Traders</strong>, or STs, who observe at t only dw t−1 .<br />

To proceed with the solution, we need to be more specific about how the dealer sets her<br />

expectation z t−k,t = E(dw t−k | {dy τ } τ


where γ S t = 0 for the slow trader. Also, denote by N F the number of FTs, <strong>and</strong> by N S the<br />

number of STs. The total number of speculators is<br />

N T = N F + N S . (18)<br />

The next result shows that the model has a closed form solution.<br />

Theorem 1. Consider the trading model with m = 1 lagged signals, with N F > 0 fast traders<br />

<strong>and</strong> N S ≥ 0 slow traders. Then, there exists a linear equilibrium of the model with constant<br />

coefficients, of the form (t ∈ (0, T ]):<br />

dx F t = γdw t + µ ˜dw t−1 ,<br />

dx S t = µ ˜dw t−1 ,<br />

˜dw t−1 = dw t−1 − ρdy t−1 ,<br />

(19)<br />

dp t = λdy t ,<br />

where the coefficients γ, µ, ρ, λ are given by:<br />

γ = 1 λ<br />

µ = 1 λ<br />

1<br />

N F + 1 ,<br />

1 1<br />

N T + 1 1 + b ,<br />

√<br />

NF<br />

N F +1 − b<br />

√<br />

NT<br />

N T +1 − ωb<br />

ρ = σ √<br />

w (1 − a)(a − b<br />

σ 2 ) = σ (20)<br />

w 1 1 + b<br />

N<br />

√ T<br />

u σ u NF + 1 ω + b<br />

N T + 1 ,<br />

N F<br />

λ = ρ<br />

N F − b ,<br />

<strong>and</strong> a ∈ (0, 1), b ∈ (0, 1) <strong>and</strong> ω ∈ (1, 2) are defined by<br />

b =<br />

√<br />

ω 2 + 4 N T<br />

N T +1<br />

2<br />

− ω<br />

, with ω = 1 + 1<br />

N F<br />

N T<br />

N T + 1 ,<br />

a = N F − b<br />

N F + 1 . (21)<br />

One consequence of the Theorem is that FTs <strong>and</strong> STs trade with the same intensity on<br />

their lagged signal. This turns out to be true because the new signal dw t is uncorrelated with<br />

the lagged signal ˜dw t−1 . This result, however, does not generalize to the case when there are<br />

more lags (see Section 6). In that case, there is autocorrelation between the signals of different<br />

lags, which reflects a more complicated covariance structure. 18<br />

18 In the language of Section 6, this translates mathematically into the fresh covariance matrix having non<br />

zero entries A i,j when i > j ≥ 1.<br />

13


The next Corollary provides asymptotic results when the number N F of FTs is large. We<br />

say that the asymptotic value of a number X which depends on N F is X ∞ if<br />

X<br />

lim = 1. (22)<br />

N F →∞ X ∞<br />

Corollary 1. In the context of Theorem 1, if N F<br />

the following asymptotic values:<br />

is large, the equilibrium coefficients have<br />

γ ∞<br />

= σ u<br />

σ w<br />

1<br />

√<br />

NF<br />

,<br />

µ ∞ = σ u<br />

σ w<br />

1<br />

√<br />

NF<br />

N F<br />

N T<br />

b ∞ ,<br />

ρ ∞ = λ ∞ = σ w<br />

σ u<br />

1<br />

√<br />

NF<br />

.<br />

(23)<br />

Moreover, the numbers b, a, ω approach<br />

b ∞ =<br />

√<br />

5 − 1<br />

, a ∞ = ω ∞ = 1. (24)<br />

2<br />

We also provide some formulas that follow from Theorem 1, which are used throughout<br />

the text.<br />

Corollary 2. In the context of Theorem 1, the following formulas are satisfied:<br />

Var ( ˜dwt<br />

)<br />

dt<br />

λ ¯γ =<br />

= (1 − a) σ 2 w = 1 + b<br />

N F + 1 σ2 w,<br />

N F<br />

N F + 1 , λ ¯µ = 1<br />

1 + b<br />

Cov ( ) ˜dwt , w t<br />

dt<br />

N T<br />

N T + 1 ,<br />

= 1 − a<br />

1 + b σ2 w =<br />

σw<br />

2 (25)<br />

N F + 1 .<br />

We use the results in Theorem 1 to compute the expected profits of the fast traders <strong>and</strong><br />

the slow traders.<br />

Proposition 1. In the context of Theorem 1, the expected profit of the FTs <strong>and</strong> STs at t = 0<br />

from their equilibrium strategies are given, respectively, by:<br />

π F =<br />

π S =<br />

γ<br />

N F + 1 + 1<br />

1<br />

N F + 1<br />

µ<br />

N T + 1 ,<br />

N F + 1<br />

(26)<br />

µ<br />

N T + 1 .<br />

The ratio of slow to fast profits is therefore<br />

π S<br />

π F = 1<br />

1 + (N T +1) 2 (1+b)<br />

N F +1<br />

π S<br />

=⇒ lim<br />

N F →∞ π F = N F 1<br />

(N F + N S ) 2 . (27)<br />

1 + b ∞<br />

14


Thus, even if there is only one ST (N S = 1), the ST profits are small compared to the<br />

FT profits. The reason is that FTs trade also on their lagged signals, <strong>and</strong> thus compete with<br />

the STs. Indeed, FTs compete for trading on dw t only among themselves, while they also<br />

compete with the STs for trading on the lagged signal ˜dw t−1 .<br />

4 Fast <strong>Traders</strong> <strong>and</strong> Market Quality<br />

We now discuss the effect of fast trading on various measures of market quality, such as<br />

liquidity, trading volume, price volatility, price informativeness, etc. First, we need to define<br />

these measures in the context of our model with fast <strong>and</strong> slow traders. Suppose the market is<br />

in the equilibrium described by Theorem 1, <strong>and</strong> the price changes according to the following<br />

rule:<br />

dp t = λ dy t = λ<br />

(du t + ¯γ dw t + ¯µ ˜dw<br />

)<br />

t−1 , (28)<br />

where ¯γ = N F γ is the aggregate equilibrium weights on dw t , <strong>and</strong> ¯µ = (N F + N S )µ = N T µ is<br />

the aggregate equilibrium weight on ˜dw t−1 .<br />

First, we define λ as the measure of illiquidity, as it is st<strong>and</strong>ard in the literature. Thus,<br />

the market is considered illiquid if the price impact per unit of trade is very large, i.e., if λ is<br />

large.<br />

We define trading volume as the infinitesimal variance of the aggregate order flow dy t :<br />

TV = σy 2 = Var(dy t)<br />

. (29)<br />

dt<br />

We argue that this is a measure of trading volume. Indeed, in each trading round the actual<br />

aggregate order flow is given by dy t . Thus, one can interpret trading volume as the absolute<br />

value of the order flow: |dy t |. From the theory of normal variables, the average trading<br />

volume is given by E ( |dy t | ) √<br />

2<br />

=<br />

π σ y. With our definition TV = σy, 2 we observe that TV<br />

is monotonic in E ( |dy t | ) , <strong>and</strong> thus TV can be used a measure of trading volume. Using the<br />

formula dy t = du t + ¯γ dw t + ¯µ ˜dw t−1 , we compute the trading volume in our model by the<br />

formula<br />

TV = σ 2 u + ¯γ 2 σ 2 w + ¯µ 2 σ 2˜w , with σ2˜w = Var( ˜dwt<br />

)<br />

dt<br />

. (30)<br />

We define price volatility σ p to be the square root of the instantaneous price variance:<br />

σ p =<br />

( ) Var(dpt ) 1/2<br />

. (31)<br />

dt<br />

From (28), it follows that the instantaneous price variance can be computed simply as the<br />

15


product of the illiquidity measure λ <strong>and</strong> the trading volume TV = σ 2 y. Thus,<br />

σ 2 p<br />

= λ 2 TV = λ 2 ( σ 2 u + ¯γ 2 σ 2 w + ¯µ 2 σ 2˜w<br />

)<br />

. (32)<br />

The trading volume measure TV can be decomposed into the noise trading volume <strong>and</strong><br />

the speculator trading volume:<br />

TV = TV n + TV s , with TV n = σu, 2 TV s = ¯γ 2 σw 2 + ¯µ 2 σ<br />

2˜w . (33)<br />

The speculator participation rate is defined as the ratio of speculator trading volume over total<br />

trading volume:<br />

SPR = TV s<br />

TV = ¯γ 2 σw 2 + ¯µ 2 σ 2˜w<br />

σu 2 + ¯γ 2 σw 2 + ¯µ 2 σ 2˜w<br />

. (34)<br />

Note that SPR can also be interpreted as the fraction of price variance due to the speculators.<br />

We define price informativeness as a measure inversely related to the forecast error variance<br />

Σ t = Var ( (w t − p t−1 ) 2) . 19 Thus, if prices are informative, they stay close to the forecast w t ,<br />

i.e., the variance Σ t is small. In Section 6, we show that in the general model of trading with<br />

m lagged signals (Proposition 6) that Σ t evolves according to: Σ ′ t = σ 2 w − σ 2 p, where σ 2 p is the<br />

price variance. Therefore, since Σ ′ t is inversely monotonic in the price variance, we do not use<br />

it as a separate measure of market quality.<br />

The next result gives explicit formulas for our measures of market quality. As before, we<br />

use the following asymptotic notation when N F is large: X ≈ Y st<strong>and</strong>s for<br />

X<br />

lim<br />

N F →∞<br />

Y = 1.<br />

Proposition 2. Consider the setup of Theorem 1 with N F fast traders <strong>and</strong> N S slow traders.<br />

Then, the price impact coefficient, trading volume, price volatility, <strong>and</strong> speculator participation<br />

rate satisfy<br />

λ = σ w<br />

σ u<br />

√<br />

(1 + b)(a − b 2 )<br />

√<br />

NF + 1<br />

N F<br />

N F − b ≈ σ w<br />

σ u<br />

1<br />

√<br />

NF<br />

,<br />

TV = σu(N 2 a<br />

F + 1)<br />

(1 + b)(a − b 2 ) ≈ σ2 u (N F + 1),<br />

(<br />

σp<br />

2 = σw<br />

2 1 − N )<br />

F (1 − b) − b<br />

≈ σ 2<br />

(N F + 1)(N F − b)<br />

w,<br />

(35)<br />

SPR = a + b2 (1 + b)<br />

N F − b<br />

≈ 1.<br />

For comparison, we solve the trading model with m = 0 lags, <strong>and</strong> compute the corresponding<br />

market quality measures.<br />

Proposition 3. Consider the case with N fast traders who can only use their last signal<br />

19 See the definition in equation (62) for the general case with m lagged signals.<br />

16


(m = 0), i.e., their trading strategy is of the form dx t = γ t dw t . Then, the coefficient γ of the<br />

optimal strategy satisfies:<br />

γ = 1 λ<br />

1<br />

N + 1 = σ u<br />

σ w<br />

1<br />

√<br />

N<br />

. (36)<br />

Also, the price impact coefficient, trading volume, price volatility, <strong>and</strong> speculator participation<br />

rate satisfy:<br />

λ = σ w<br />

σ u<br />

TV = σ 2 u(N + 1),<br />

√<br />

N<br />

N + 1 = √ a σ v<br />

σ u<br />

√<br />

N<br />

N + 1 ,<br />

(37)<br />

σp<br />

2 = σw<br />

2 N<br />

N + 1 = N<br />

aσ2 v<br />

N + 1 ,<br />

SPR =<br />

N<br />

N + 1 .<br />

We see that Proposition 3 <strong>and</strong> 2 produce qualitatively similar results, showing that the fast<br />

traders dominate the market quality measures. We thus discuss the empirical implications of<br />

Proposition 3, knowing that they are robust to the existence of slower traders.<br />

We now discuss the effect of an increase in the number N of FTs on the market quality<br />

measures. An important consequence of Proposition 3 is that in our context, the speculator<br />

participation rate can be made arbitrarily close to 1 if the number N of FTs is large. Thus,<br />

noise trading volatility need only be a small part of the total volatility. This st<strong>and</strong>s in sharp<br />

contrast for instance with the models of Kyle (1985) or Back, Cao, <strong>and</strong> Willard (2000), in which<br />

at the very high frequency (in continuous time) virtually all instantaneous price volatility is<br />

generated by noise traders.<br />

Also, note that the market is more efficient when N is large. Indeed, Proposition 6 implies<br />

that the rate of change of the forecast error variance Σ ′ is constant <strong>and</strong> equal to σ 2 w − σ 2 p, <strong>and</strong><br />

therefore Σ 1 = σ 2 w − σ 2 p as well (we have assumed Σ 0 = 0). In our case, Proposition 3 implies<br />

that Σ ′ = σ2 w<br />

N+1 , which implies that Σ t ≤ σ2 w<br />

N+1 for all t. In other words, Σ t is close to zero at all<br />

times, which means that prices stay close to fundamentals at all times. Thus, a larger number<br />

N of FTs, rather than destabilizing the market, make the market in fact more efficient.<br />

The trading volume TV strongly increases with the number N of FTs. 20<br />

This occurs<br />

because of the competition among FTs make them trade more aggressively. At the same time,<br />

by trading more aggressively, they reveal more information which, as we see below, also lowers<br />

the price impact. This has an amplifier effect on the trading aggressiveness, to the point that<br />

the trading volume grows essentially linearly in the number of speculators, as can be seen<br />

from Proposition 3. The speculator participation rate SPR also increases in N, since SPR is<br />

20 This is also true when more lagged signals are used (m > 1), <strong>and</strong> all speculators have equal speed. Indeed,<br />

in this case Proposition 7 implies that TV = σ 2 u(N + 1) as well.<br />

17


the fraction of trading volume caused by the speculators.<br />

Surprisingly, a larger number of FTs make the market more liquid, as more information<br />

is revealed when there are more competing speculators. This appears to be in contradiction<br />

with the fact that more informed trading should increase the amount of adverse selection. To<br />

underst<strong>and</strong> the source of this apparent contradiction, note that illiquidity is measured by the<br />

price impact λ of one unit of volume. But, while the volume of trading TV strongly increases<br />

in N in an unbounded way, at the same time the price impact is bounded by magnitude of the<br />

signal dw t . 21 Thus, the price impact per unit of volume may actually decrease, <strong>and</strong> in fact it<br />

does in our case, as prices are overall more informative. This makes the market more liquid,<br />

hence λ decreases. This is consistent with the empirical studies of Zhang (2010), Hendershott,<br />

Jones, <strong>and</strong> Menkveld (2011), <strong>and</strong> Boehmer, Fong, <strong>and</strong> Wu (2014).<br />

To underst<strong>and</strong> the effect of N on the price volatility σ p , recall the pricing formula dy t =<br />

λdy t which implies σp 2 = λ 2 TV . Thus, there are two effects of N on the price volatility σ P .<br />

First, the trading volume TV strongly increases in N, which has a positive effect on σ P .<br />

Second, price impact λ decreases in N, which has a negative effect on σ P . The first effect is<br />

slightly stronger than the second, thus on net price volatility σ P increases in N. This result<br />

is consistent with the empirical studies of Boehmer, Fong, <strong>and</strong> Wu (2014) <strong>and</strong> Zhang (2010).<br />

A few caveats are in order. First, all these studies analyze the effects of HFT activity,<br />

where activity is proxied either by turnover or by intensity of order-related message traffic,<br />

<strong>and</strong> not by the number of HFTs present in the market. An answer to this concern is that,<br />

as we have seen, trading volume does increase in N. Second, in our paper we do not model<br />

passive HFTs, that is, HFTs that offer liquidity via limit orders. Therefore, it is quite possible<br />

that an increase in the number of passive HFTs decreases price volatility, which would cancel<br />

the opposite effect of the number of active HFTs. For instance, Hasbrouck <strong>and</strong> Saar (2012)<br />

document a negative effect of HFTs on volatility, possibly due to the fact that they also<br />

consider passive HFTs, which by providing liquidity have a stabilizing effect on price volatility.<br />

Moreover, Chaboud, Chiquoine, Hjalmarsson, <strong>and</strong> Vega (2014) find essentially no relation. In<br />

our model, the dependence of price volatility on N is weak, which may explain the different<br />

results in the empirical literature.<br />

We now discuss how the various measures of market quality depend on the FTs’ signal<br />

precision a. An interesting consequence of Proposition 3 is that the signal precision a <strong>and</strong> the<br />

fundamental volatility σ v do not affect the market quality separately, but only via the product<br />

σ w = √ aσ v . In what follows, we discuss only the results regarding the signal precision, but<br />

those results apply equally to the fundamental volatility.<br />

The price volatility σ p increases in the FTs’ signal precision a. Indeed, σ p is the volatility<br />

of dp t , which is the price impact of the aggregate order flow. In particular, the order flow<br />

21 See the discussion surrounding Proposition 9, which makes this intuition more rigorous in the general case.<br />

18


coming from FTs has an aggregate price impact of λ Nγdw t =<br />

N<br />

N+1 dw t. 22 But the volatility<br />

of this is proportional to σ 2 w = aσ 2 v, which is proportional to the signal precision. Hence, σ p<br />

increases in the signal precision a, <strong>and</strong> so does the speculator participation rate SPR.<br />

Signal precision a increases adverse selection between FTs <strong>and</strong> the dealer, hence it increases<br />

the illiquidity λ. Trading volume TV is independent of a. To get some intuition for this result,<br />

note that TV = σ2 p<br />

λ 2 . Since both the numerator <strong>and</strong> denominator increase with signal precision,<br />

the net effect is not clear. Indeed, in the context of Proposition 3, the two effects exactly offset<br />

each other.<br />

We now discuss the autocorrelation of order flow. Brogaard (2011) finds empirically that<br />

the autocorrelation of aggregate HFT order flow is small but positive. We find that in our<br />

theoretical model the aggregate speculator order flow has a positive serial autocorrelation.<br />

To underst<strong>and</strong> why, we consider our baseline model with fast <strong>and</strong> slow traders. Denote the<br />

aggregate speculator order flow by<br />

d¯x t = ¯γdw t + ¯µ ˜dw t−1 , with ¯γ = N F γ, ¯µ = N T µ, (38)<br />

where ¯γ <strong>and</strong> ¯µ are the aggregate equilibrium coefficients on the signals. Since ˜dw t = dw t −ρdy t ,<br />

the order flow serial autocorrelation has two components:<br />

Cov ( dx t , dx t+1<br />

)<br />

Var(dx t )<br />

= ¯µ¯γ Cov( dw t , ˜dw<br />

)<br />

t<br />

Var(dx t )<br />

= ¯µ¯γ Cov( )<br />

dw t , dw t<br />

Var(dx t )<br />

+ ¯µ 2 Cov( ˜dwt−1 , ˜dw t<br />

)<br />

Var(dx t )<br />

+ ¯µ 2 Cov( ˜dwt−1 , −ρ¯µ ˜dw<br />

)<br />

t−1<br />

.<br />

Var(dx t )<br />

Thus, order flow autocorrelation occurs for two reasons: (i) there is a positive correlation<br />

between the new signal of FTs at t (dw t ) <strong>and</strong> the lagged signal of both the FTs <strong>and</strong> STs at<br />

t + 1 ( ˜dw t ); (ii) there is a negative correlation between the lagged signal at t ( ˜dw t−1 ) <strong>and</strong> the<br />

lagged signal at t + 1 ( ˜dw t = dw t − ρdy t ), due to the negative expectation adjustment of dw t<br />

at t + 1.<br />

Proposition 4. In the context of Theorem 1, the speculator order flow serial autocorrelation<br />

is:<br />

Corr ( dx t , dx t+1<br />

)<br />

=<br />

b(b + 1)(a − b 2 )<br />

a 2 + b 2 (1 − a)<br />

1<br />

N F + 1 ≈<br />

(39)<br />

b ∞<br />

N F<br />

. (40)<br />

where b ∞ =<br />

√<br />

5−1<br />

2<br />

= 0.618.<br />

Thus, in the model the speculator order flow autocorrelation is low, yet positive. For<br />

instance, with N F = 20 FTs, the speculator order flow autocorrelation is approximately 3%.<br />

This result is consistent with the empirical literature on HFTs (see Brogaard (2011)). To<br />

get some intuition, note that order flow autocorrelation arises from speculators’s signals this<br />

22 See the equation for γ in Proposition 3.<br />

19


period being correlated with their lagged signals next period. However, when the number of<br />

1<br />

FTs is large, there is only<br />

N F +1<br />

of the signal left to traders for the next period. Hence, one<br />

1<br />

should expect the autocorrelation to decrease by the order of<br />

N F +1<br />

, which is indeed the case.<br />

5 Inventory Management<br />

In this section, we show that the fast traders can manage their inventory extremely quickly,<br />

but only in the presence of slower traders. To get some intuition, consider the model with fast<br />

<strong>and</strong> slow traders from Section 3. For simplicity, we take a partial equilibrium approach <strong>and</strong><br />

assume that a subset of the FTs decides to manage their inventory by adding a mean-reverting<br />

component to their trading strategy, as described below. To distinguish these traders from the<br />

regular FTs, we call them <strong>High</strong> <strong>Frequency</strong> <strong>Traders</strong>, or HFTs in short. This is indeed consistent<br />

with Kirilenko et al. (2014), who include in the definition of HFTs the condition of having a<br />

low inventory.<br />

We find that essentially there are two regimes of inventory management. In the first<br />

regime, called the smooth regime, inventories mean revert to zero after a relatively long time.<br />

In the second regime, called the extreme regime, inventories can mean revert to zero after<br />

an infinitesimal amount of time (of the order of dt). In both cases, we show that inventory<br />

management can be done such that profits remain positive, but this is true only if there exist<br />

slower traders.<br />

Thus, in the smooth regime we assume that HFTs have a trading strategy of the form: 23<br />

dx HFT<br />

Sm,t = −θx t−1 dt + γ ∗ dw t . (41)<br />

With this definition, the HFT’s inventory x t becomes a regular Ornstein–Uhlenbeck process,<br />

which is the generalization of an AR(1) process to continuous time.<br />

In the extreme regime we assume that HFTs have a trading strategy of the form<br />

dx HFT<br />

Ex,t = −Θx t−1 + γ ∗ dw t . (42)<br />

With this definition, the HFT’s inventory x t becomes an actual AR(1) process, only that<br />

everything happens in very rapid succession.<br />

The main difference is that this time there<br />

is no “dt” term that multiplies the inventory x t−1 . In other words, in the extreme regime<br />

the mean reversion coefficient (Θ) is an order of magnitude ( 1 dt<br />

times) larger than the mean<br />

reversion coefficient in the smooth regime (θdt). Interestingly, even under extreme inventory<br />

management, all the relevant variables are well defined, including the expected profit of the<br />

HFTs. Normally, one should expect that the extreme inventory management erodes the HFT<br />

23 To obtain simpler formulas, we have assumed that HFTs do not use any lagged signals. We have also<br />

worked out the case in which HFTs use lagged signals as well, <strong>and</strong> the qualitative results are the same.<br />

20


profits to the point where they become negative. But it turns out that losses can be prevented<br />

if there exist also slower traders.<br />

Note that in the extreme regime we can write the HFT inventory in autoregressive form:<br />

x t = (1 − φ)x t−1 + γ ∗ dw t , with φ = 1 − Θ. (43)<br />

With the usual definition, the half life of the inventory is then defined as ln(1/2)<br />

ln(φ)<br />

. This by<br />

definition is the average number of trading rounds that the process needs to halve the distance<br />

from its mean. Since a period in our model is of length dt, the inventory half life satisfies:<br />

Inventory Half Life Extreme<br />

= ln(1/2)<br />

ln(φ)<br />

dt. (44)<br />

This is the sense in which HFTs in our model can do “real-time” inventory management.<br />

Depending on how dt is interpreted in practice, the inventory half life can be a few seconds,<br />

milliseconds or even microseconds.<br />

By contrast, in the smooth regime, the autoregressive coefficient is 1 − θdt, which gives<br />

the following formula<br />

Inventory Half Life Smooth = ln(1/2)<br />

ln(1 − θdt)<br />

ln(2)<br />

dt = . (45)<br />

θ<br />

Note that, indeed, the inventory half life in the smooth regime is one order of magnitude<br />

higher than in the extreme regime.<br />

The next result analyzes the expected profits of the HFTs in the two inventory management<br />

regimes.<br />

Theorem 2. Consider the model from Section 3, to which we add N I HFTs. These have a<br />

trading strategy engage in inventory management, as described above, using either the smooth<br />

regime or the extreme regime. Suppose ¯γ <strong>and</strong> ¯µ are the aggregate speculator coefficients on<br />

dw t <strong>and</strong> ˜dw t−1 , respectively, <strong>and</strong> ρ <strong>and</strong> λ are the dealer’s pricing coefficients. (These numbers<br />

need not take the equilibrium values.)<br />

In the smooth regime, the expected profit of a HFT (at t = 0) satisfies:<br />

π HFT<br />

Sm,θ<br />

σ 2 w<br />

=<br />

( )<br />

γ ∗ − λγ ∗¯γ e −θt 1 − ρ¯γ<br />

+ λ¯µγ ∗<br />

1 + ρ¯µ<br />

λγ 2 ∫ 1 (<br />

∗<br />

− N I<br />

1 − e<br />

−2θt ) dt.<br />

2(1 + ρ¯µ)<br />

0<br />

∫ 1<br />

0<br />

(<br />

1 − e<br />

−θt ) dt<br />

(46)<br />

This expression has a limit when θ → ∞:<br />

πSm,θ<br />

HFT<br />

lim<br />

θ→∞ σw<br />

2<br />

1 − ρ¯γ<br />

= λ¯µγ ∗<br />

1 + ρ¯µ − N λγ∗<br />

2 I<br />

2(1 + ρ¯µ) . (47)<br />

21


In the extreme regime, the expected profit of a HFT satisfies:<br />

π HFT<br />

Ex,Θ<br />

σ 2 w<br />

1 − ρ¯γ<br />

= λ¯µγ ∗<br />

1 + φρ¯µ − N λγ∗<br />

2 I<br />

, with φ = 1 − Θ. (48)<br />

(1 + φ)(1 + φρ¯µ)<br />

Furthermore, the profits in the two regimes are continuous across regimes, i.e.,<br />

lim<br />

θ→∞ πHFT Sm,θ<br />

= lim<br />

Θ→0<br />

π HFT<br />

Ex,Θ. (49)<br />

Theorem 2 shows that the HFT profits vary continuously under the two inventory management<br />

regimes. With no inventory management (θ = 0), the profits satisfy the usual formula,<br />

π HFT<br />

Sm,θ = ( γ ∗ − λγ ∗¯γ ) σ 2 w. (Recall that for simplicity we assume that the HFT does not trade<br />

on his lagged signal, i.e., µ ∗ = 0.) When θ approaches ∞, the profit becomes equal to the<br />

starting profit for the extreme regime (Θ = 0, or φ = 1). We continue a discussion of these<br />

issues after we give more explicit formulas in Proposition 5.<br />

If in addition 1 − ρ¯γ > 0 (which is true in the equilibrium of Theorem 1), then the HFT<br />

can engage in extreme inventory management if <strong>and</strong> only if ¯µ > 0. 24<br />

This means that there<br />

must be at least some traders that put a positive weight on the lagged signal, ˜dwt−1 .<br />

other words, the HFT needs slower trader to mitigate the extra price impact that comes from<br />

inventory management. To underst<strong>and</strong> the intuition for this fact, consider a speculator who<br />

does not manage inventory. Then, his inventory represents an accumulation of signals which<br />

are positively correlated with the asset value. If instead, the speculator chooses to mean revert<br />

his inventory to zero, this means that he will trade against his information. It is reasonable<br />

then to expect that if he mean reverts too fast, he may eventually trade at a loss. However,<br />

in the presence of slower traders, the HFT lowers his inventory at the same time that other<br />

traders are accumulating inventory.<br />

The aggregate price impact is therefore lower for the<br />

HFTs, <strong>and</strong> they can make positive profits. As we will see shortly, in the presence of slower<br />

traders, the HFTs can maintain positive profits even at the very end of the extreme regime<br />

(Θ = 1).<br />

Theorem 2 also shows that the extreme regime cannot be simply reduced to the case with<br />

no inventory management by taking Θ = 0. Indeed, according to the Theorem, the limit when<br />

Θ → 0 of the extreme regime coincides with the case θ = ∞ in the smooth regime, <strong>and</strong> not<br />

with the case θ = 0, which is the case without inventory management.<br />

The importance of this result is that inventory management in real-time makes a significant<br />

departure from the case of no inventory management. In other words, there is a significant<br />

difference between regular FTs <strong>and</strong> HFTs: for the regular FTs, the inventory follows a r<strong>and</strong>om<br />

walk <strong>and</strong> thus is likely to become large over time; while for the HFTs, the inventory is<br />

24 Formally, this is a linear-quadratic problem in γ ∗ with a positive coefficient on γ ∗ <strong>and</strong> a negative coefficient<br />

on γ 2 ∗.<br />

In<br />

22


stationary <strong>and</strong> it mean reverts extremely quickly to zero. Our result appears consistent with<br />

the behavior of HFTs observed in practice by Kirilenko, Kyle, Samadi, <strong>and</strong> Tuzun (2014).<br />

They report that the HFTs in their sample (several days around the Flash Crash of May 6,<br />

2010) liquidate 0.5% of their aggregate inventories on average each second, implying HFT<br />

inventories have an AR(1) half-live of a little over 2 minutes.<br />

The next result helps to better quantify the tradeoffs that HFTs face when managing<br />

their inventories. To simplify formulas, we consider only one HFT (N I = 1), <strong>and</strong> assume<br />

that the HFT exogenously chooses his weight on dw t equal to the equilibrium weight from<br />

Section 3, i.e., γ ∗ = γ. 25 Finally, to maintain the same aggregate coefficients as in Theorem 1,<br />

we introduce one more slow trader, to compensate for the HFT not trading on his lagged<br />

signal. When N F is large, we also provide asymptotic formulas, where X ≈ Y st<strong>and</strong>s for<br />

X<br />

lim<br />

N F →∞<br />

Y = 1.<br />

Proposition 5. In the context of Theorem 1, assume that N F −1 FTs <strong>and</strong> N S +1 STs choose<br />

their equilibrium strategies described in the Theorem; also, the dealer sets the same pricing<br />

coefficients. In addition, one speculator, called the HFT, chooses a inventory management<br />

strategy as in (41) or (42). Then, the HFT’s expected profit under the different regimes<br />

satisfies:<br />

π HFT<br />

Sm,θ=0<br />

σ 2 w<br />

π HFT<br />

Sm,θ=∞<br />

σ 2 w<br />

π HFT<br />

Ex,Θ<br />

σ 2 w<br />

π HFT<br />

Ex,Θ=1<br />

σ 2 w<br />

= γ(1 − λ¯γ) = γ 1 − a<br />

1 + b ≈ γ<br />

N F<br />

,<br />

= πHFT Ex,Θ=0<br />

σw<br />

2 ≈<br />

γ<br />

=<br />

N F (1 + φb)<br />

≈ 0.<br />

γ b ∞<br />

N F 2 ,<br />

(<br />

λ<br />

N F<br />

b(1 + b)<br />

ρ N F + 1 −<br />

a )<br />

1 + φ<br />

≈<br />

γ 1 φ<br />

N F 1 + φb ∞ 1 + φ ,<br />

(50)<br />

Note that, according to Proposition 1, the expected profit of a regular FT equals<br />

π FT =<br />

γ<br />

N F + 1 + 1<br />

N F + 1<br />

µ<br />

N T + 1 ≈<br />

γ<br />

N F<br />

. (51)<br />

Therefore, without inventory management the profit of the HFT is approximately equal to that<br />

of the FT. We now compute the effect of smooth inventory management, taken to the limit<br />

√<br />

when θ → ∞. According to Proposition 1, this profit is only a fraction b∞ 2<br />

= 5−1<br />

4<br />

= 0.309<br />

of the baseline profit π FT . And, at the other end of the extreme the profit of the HFT is<br />

completely eroded. Thus, smooth inventory management erases approximately 69% of the<br />

25 We could let the HFT maximize also over γ ∗ in the linear-quadratic formula from Theorem 2,<br />

(<br />

but that<br />

would lead to more complicated formulas, <strong>and</strong> to a profit which is approximately equal within O 1<br />

)<br />

NF N F<br />

. The<br />

fact that we set N I = 1 is not really restrictive. Indeed, for N I > 1 HFTs, we can take γ ∗ = γ N I<br />

, <strong>and</strong> then the<br />

formulas below would be the same, except that profits would get divided by N I.<br />

23


aseline profit of the HFT, while extreme inventory management taken to its extreme (Θ = 1)<br />

erases the remaining 31%. This shows that the smooth inventory management is quite costly<br />

as well.<br />

We conclude this section with a brief discussion on inventory management in the extreme<br />

regime. Assuming that HFTs face a quadratic inventory cost, equation (K.38) in the Internet<br />

Appendix implies that:<br />

Denote by<br />

(∫ 1<br />

)<br />

C × E x 2 t dt<br />

0<br />

= C<br />

∫ 1<br />

0<br />

Ω xx<br />

t dt σ 2 w = C γ2 ∗<br />

1 − φ 2 σ2 w = C<br />

γ 2<br />

(1 − φ 2 ) σ2 w. (52)<br />

C ρ = C ρ ≈ C σ u<br />

σ w<br />

√<br />

NF . (53)<br />

Since γ 2 = aγ<br />

ρN F<br />

≈ γ<br />

ρN F<br />

, the HFT expected profit net of inventory costs (normalized by σ 2 w) is:<br />

πC HFT ≈ γ (<br />

φ<br />

N F (1 + φb ∞ )(1 + φ) − C )<br />

ρ<br />

1 − φ 2 , (54)<br />

One verifies that this profit increases near φ = 0 (θ = 1) <strong>and</strong> approaches −∞ near φ = 1<br />

(θ = 0). Thus, as long as C > 0, some inventory management is optimal. The only problem<br />

is that when C is large (numerically, when C ρ > 0.195), the HFT profit becomes negative no<br />

matter what φ is chosen. In that case, the HFT chooses to not participate in trading.<br />

6 Discussion <strong>and</strong> Robustness<br />

6.1 The General Case<br />

We now solve for the equilibrium of the general model of trading with m lagged signals, from<br />

Section 2.2. Then, under an additional assumption stated below, we show that the equilibrium<br />

reduces the solution of a system of equations (see Theorem 3). This system can be solved in<br />

closed form in some particular cases of interest, <strong>and</strong> can in principle be solved numerically.<br />

To proceed with our solution, we need to be more specific about how the dealer sets her<br />

expectation z t−i,t = E(dw t−i | {dy τ } τ


e the unanticipated part of the signal at t:<br />

d t w t−i = dw t−i − z t−i,t . (56)<br />

For all lags i, j = 0, . . . , m, denote by<br />

A i,j,t = Cov( )<br />

d t w t−i , d t w t−j<br />

σw 2 , B j,t = Cov( )<br />

w t , d t w t−j<br />

dt<br />

σw 2 , (57)<br />

dt<br />

Since A measures the instantaneous covariance of fresh signals at the relevant lags, we call A<br />

the fresh covariance matrix. The vector B measures the instantaneous contribution of each<br />

fresh signal to the profit, thus we call B the benefit vector. In Section 2, we have assumed<br />

that the speculator takes A <strong>and</strong> B as fixed, <strong>and</strong> considers them as set by the dealer (just as<br />

ρ j,t <strong>and</strong> λ t ).<br />

For simplicity of notation, we normalize some variables by dividing them with the forecast<br />

variance, σw. 2 We denote this by placing a tilde above the variable. For instance, we define<br />

the normalized instantaneous order flow variance ˜σ y, 2 as well as the normalized instantaneous<br />

noise trader variance ˜σ u 2 as follows:<br />

˜σ y,t 2 = Var(dy t)<br />

σwdt<br />

2 , ˜σ u 2 = Var(du t)<br />

σwdt<br />

2<br />

= σ2 u<br />

σw<br />

2 . (58)<br />

The next result shows that a linear equilibrium exists if a certain system of equations is<br />

satisfied. For simplicity, we write the system of equations using matrix notation. All vectors<br />

are in column format, <strong>and</strong> we denote by X ′ the transpose a vector X. Thus, we collect the<br />

equilibrium coefficients in the following vectors, 26 with l = 0, . . . , m:<br />

γ (l)<br />

= [ γ (l)<br />

l<br />

, . . . , γ (l)<br />

m<br />

] ′, ρ =<br />

[<br />

ρ0 , . . . , ρ m<br />

] ′, (59)<br />

In general, A ≥l<br />

is the matrix with elements A i,j for i, j ≥ l; <strong>and</strong> similarly for the vectors B ≥l<br />

<strong>and</strong> ρ ≥l<br />

. A sum of vectors X ≥l over different l is carried by padding X ≥l<br />

with zeros for the<br />

first l entries.<br />

Theorem 3. Let m ≥ 0 be fixed, <strong>and</strong> consider the model with N l speculators of type l =<br />

0, . . . , m, in which there is no information processing cost (c = 0). Suppose there exists a<br />

26 Note that ρ m, the last entry of the vector ρ, is not part the dealer’s expectations z t−i,t, but we introduce it<br />

in order to simplify notation. In particular, the last row of the equilibrium equation ˜σ 2 yρ = A¯γ can be omitted.<br />

25


linear equilibrium of the model with constant coefficients, of the form:<br />

dx (l)<br />

t<br />

= γ (l)<br />

l<br />

d t w t−l + · · · + γ m (l) d t w t−m , l = 0, . . . , m,<br />

d t w t−i = dw t−i − z t−i,t , i = 0, 1, . . . , m,<br />

z t−i,t = ρ 0 dy t−i + · · · + ρ i−1 dy t−1 , i = 0, 1, . . . , m,<br />

dp t = λdy t .<br />

(60)<br />

Then, the constants λ, ρ i , γ (l)<br />

i<br />

¯γ =<br />

m∑<br />

l=0<br />

satisfy the following system of equations (i = 0, . . . , m):<br />

N l γ (l) , with γ (l) = ( A ≥l<br />

) −1<br />

( 1<br />

λ B ≥l − ˜σ2 yρ ≥l<br />

)<br />

,<br />

˜σ 2 yρ = A¯γ, ˜σ 2 yλ = B ′¯γ, ˜σ 2 y =<br />

A i,j<br />

= 1 i=j − ˜σ 2 y<br />

min(i,j)<br />

∑<br />

k=1<br />

˜σ 2 u<br />

1 − ¯γ ′ ρ ,<br />

ρ i−k ρ j−k , B i = 1 − ˜σ 2 yλ<br />

i∑<br />

ρ i−k ,<br />

k=1<br />

(61)<br />

where 1 P is the indicator function, which equals 1 if P is true, <strong>and</strong> 0 otherwise.<br />

Conversely, suppose the constants λ, ρ i , γ (l)<br />

i<br />

satisfy (61), <strong>and</strong> in addition the following<br />

conditions are satisfied: (i) λ > 0; (ii) for all l = 0, . . . , m, the matrix A ≥l<br />

is invertible; <strong>and</strong><br />

(iii) the numbers β k = ∑ m−k<br />

i=0 ρ i¯γ k+i satisfy 1 > β 1 > · · · > β m > 0. Then, the equations<br />

in (60) provide an equilibrium of the model.<br />

In principle, the system of equations (61) can be solved numerically as follows. To simplify<br />

notation, we make a change of variables <strong>and</strong> denote by r i = ˜σ y ρ i , Λ = ˜σ y λ, g = ¯γ˜σ y<br />

. Then,<br />

suppose we start with some values for r i <strong>and</strong> Λ. Then, A can be expressed only in terms of<br />

r i , <strong>and</strong> the equation ˜σ 2 yρ = A¯γ implies that g = A −1 r can also be expressed only in terms<br />

of r i . Also, B can be expressed only in terms of r i <strong>and</strong> Λ. Then, the first equation in (61)<br />

<strong>and</strong> Λ = B ′ g (which is the rescaled equation, ˜σ 2 yλ = B ′¯γ) become m + 2 equations in the<br />

variables r i <strong>and</strong> Λ.<br />

In practice, however, this procedure does not work well. Numerically, it turns out that<br />

the solution is badly behaved, especially when N or m are large. 27<br />

Moreover, without more<br />

explicit formulas, it is difficult to study properties of the solution. In the next section, we<br />

provide a more explicit solution for the case when all speculators have the same speed.<br />

We now analyze the forecast error variance,<br />

Σ t = Var ( (w t − p t−1 ) 2) . (62)<br />

27 This is because in that case the matrix A is almost singular, <strong>and</strong> thus the equation g = A −1 r produces<br />

unreliable solutions. Indeed, equation (I.59) from the Internet Appendix shows that that the determinant of<br />

(A 0 ) −1 , a matrix close to A −1 , is equal to (N+1)m+1 . This is a large number when m or N are large.<br />

(m+1)N+1<br />

26


Note that Σ t is inversely related to price informativeness. Indeed, when prices are informative,<br />

they stay close to the forecast w t , which implies that the variance Σ t is small. Define the<br />

instantaneous price variance by<br />

σ 2 p = Var(dp t)<br />

dt<br />

= λ2 Var(dy t )<br />

dt<br />

= λ 2 σ 2 y. (63)<br />

The next result shows that growth rate of Σ is constant, <strong>and</strong> it is equal to the difference<br />

between the forecast variance σw 2 <strong>and</strong> the price variance σp.<br />

2<br />

Proposition 6. In the context of Theorem 3, the growth in the forecast error variance is<br />

constant, <strong>and</strong> satisfies the following formula:<br />

Σ ′ t = σ 2 w − σ 2 p. (64)<br />

This result can be explained by the fact that competition among speculators increases<br />

price volatility, with an upper bound given by the forecast volatility σw. 2 But competition<br />

also makes prices more informative, which implies that the forecast error variance grows more<br />

slowly. As we see in the next section, it is a feature of this equilibrium to have the forecast<br />

error variance grow at a positive rate. This result is in contrast to the Kyle (1985) model, in<br />

which the forecast error variance decreases at a constant rate, so that it becomes zero at the<br />

end. The reason for this difference is that in our model traders in equilibrium only use their<br />

most recent signals, <strong>and</strong> thus do not trade on longer-lived information.<br />

6.2 The Equal Speed Case<br />

In this section, we search for an equilibrium in the simpler case when there is no speed<br />

difference among speculators. Theorem 4 provides an efficient numerical procedure to solve<br />

for the equilibrium. Also, when the number of speculators is large, we obtain some asymptotic<br />

formulas for the equilibrium trading strategies <strong>and</strong> pricing functions. Proposition 8 then<br />

shows that the value of information decays exponentially. This result is proved rigorously<br />

only asymptotically, when the number of speculators is large. However, we have verified<br />

numerically that the result remains true for all values of the parameters we have checked.<br />

The first result is a restatement of Theorem 3 to the case when all speculators have the<br />

same speed.<br />

Proposition 7. Let m ≥ 0 be fixed, <strong>and</strong> consider the model with N speculators of type l = 0, in<br />

which there is no information processing cost (c = 0). Suppose there exists a linear equilibrium<br />

27


Figure 2: Optimal Trading Weights. Consider the model in which N ∈ {1, 5, 100}<br />

identical speculators trade on at most m ∈ {1, 2, 5, 20} lagged signals. The figure plots the<br />

1<br />

rescaled aggregate weight g i = Nγ i √ σ w<br />

N+1 σu<br />

(continuous line) against the lag i = 0, . . . , m,<br />

<strong>and</strong> compares it with the value g 0 i =<br />

N<br />

N+1<br />

(m−i+1)N+1<br />

(m+1)N+1<br />

(dashed line).<br />

1<br />

N = 1 , m= 1<br />

1<br />

N = 5 , m= 1<br />

1<br />

N = 100 , m= 1<br />

0.8<br />

0.8<br />

0.8<br />

0.6<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

0.2<br />

0<br />

0 1<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

lag<br />

0<br />

0 1 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

N = 1 , m= 2<br />

lag<br />

0<br />

0 1 2 3 4 5<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

N = 1 , m= 5<br />

lag<br />

N = 1 , m= 20<br />

0<br />

0 5 10 15 20<br />

lag<br />

0<br />

0 1<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

lag<br />

N = 5 , m= 2<br />

0<br />

0 1 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

lag<br />

0<br />

0 1 2 3 4 5<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

N = 5 , m= 5<br />

lag<br />

N = 5 , m= 20<br />

0<br />

0 5 10 15 20<br />

lag<br />

0<br />

0 1<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

lag<br />

N = 100 , m= 2<br />

0<br />

0 1 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

lag<br />

N = 100 , m= 5<br />

0<br />

0 1 2 3 4 5<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

lag<br />

N = 100 , m= 20<br />

0<br />

0 5 10 15 20<br />

lag<br />

28


of the model with constant coefficients, of the form:<br />

dx t = γ 0 d t w t + γ 1 d t w t−1 + · · · + γ m d t w t−m ,<br />

d t w t−i = dw t−i − z t−i,t , i = 0, 1, . . . , m,<br />

z t−i,t = ρ 0 dy t−i + · · · + ρ i−1 dy t−1 , i = 0, 1, . . . , m,<br />

dp t = λdy t .<br />

(65)<br />

Then, the constants λ, ρ i , γ i <strong>and</strong> ¯γ i = Nγ i satisfy the following system of equations (i =<br />

0, . . . , m):<br />

B i =<br />

1<br />

(N + 1) i , ˜σ y = ˜σ u<br />

√<br />

N + 1, ˜σ<br />

2<br />

y ρ i = 1 λ<br />

˜σ 2 yρ = A¯γ, ˜σ 2 yλ = B ′¯γ, A i,j = 1 i=j − 1<br />

˜σ 2 yλ 2<br />

N<br />

(N + 1) i+1 ,<br />

N<br />

N + 2<br />

(N + 1) 2 min(i,j) − 1<br />

(N + 1) i+j .<br />

(66)<br />

Conversely, suppose the constants λ, ρ i , γ i satisfy (66), <strong>and</strong> in addition the following<br />

conditions are satisfied: (i) λ > 0; (ii) the matrix A is invertible; <strong>and</strong> (iii) the numbers<br />

β k = ∑ m−k<br />

i=0 ρ i¯γ k+i satisfy 1 > β 1 > · · · > β m > 0. Then, the equations in (65) provide an<br />

equilibrium of the model.<br />

Note that the system of equations (66) has a simpler form. Following the discussion after<br />

Theorem 3, we make a change of variables <strong>and</strong> denote by r i = ˜σ y ρ i , Λ = ˜σ y λ, g = ¯γ˜σ y<br />

. In<br />

N<br />

(N+1) i+1<br />

this case, we see that r i = 1 Λ<br />

can further be expressed in terms of Λ. This suggests<br />

the following procedure to search for a solution of (66). Suppose we start with some value for<br />

Λ. From (66) we see that all the constants of the model (g, A, B, r) can be expressed as a<br />

function of Λ. Then, the equation Λ = g ′ B becomes the equation that determines Λ. In the<br />

Internet Appendix, we show that the equation in Λ is an infinite polynomial equation, which<br />

in practice can be solved very accurately. Then, the conditions (i) <strong>and</strong> (ii) from Proposition 7<br />

follow from a certain condition on Λ, provided in the Internet Appendix, in the course of the<br />

proof of Theorem 4.<br />

The next result uses the procedure outlined above to find approximations for the equilibrium<br />

coefficients, which use the “big-O” notation. 28<br />

Theorem 4. Consider the case with N speculators of equal speed, <strong>and</strong> let m be the maximum<br />

28 For α ∈ R, we say that the expression x N is of the order of N α , <strong>and</strong> write x N = O N (N α ), if there exists<br />

an integer N ∗ <strong>and</strong> a real number M such that |x N | ≤ M ∣ ∣ N<br />

α for all N ≥ N∗. In other words, x M is of order<br />

N α if x N<br />

N α is bounded when N is sufficiently large.<br />

29


signal lag used by the speculators. Define the following numbers:<br />

γ 0 i<br />

ρ 0 i<br />

λ 0<br />

= σ u 1 m − i + 1<br />

√ , i = 0, 1, . . . , m,<br />

σ w N + 1 m + 1<br />

= σ w<br />

σ u<br />

N<br />

, i = 0, 1, . . . , m − 1,<br />

(N + 1) i+1+1/2<br />

= σ w<br />

σ u<br />

1<br />

√<br />

N + 1<br />

.<br />

(67)<br />

Then if conditions (I.53) <strong>and</strong> (I.54) from the Internet Appendix are satisfied, 29 there exists an<br />

equilibrium. In this equilibrium, the coefficients of the optimal strategy (γ i ) <strong>and</strong> of the pricing<br />

functions (λ, ρ i ) approximate the coefficients in (67) as follows:<br />

γ i<br />

ρ i<br />

= γ 0 i<br />

= ρ 0 i<br />

(<br />

1 + O N<br />

(<br />

1<br />

) ) , i = 0, 1, . . . , m,<br />

(<br />

1 + O N<br />

( 1<br />

N<br />

) ) , i = 0, 1, . . . , m − 1,<br />

λ = λ 0( 1 − O N<br />

( 1<br />

N<br />

) ) .<br />

(68)<br />

Figure 2 plots the optimal weights for various numbers of speculators N <strong>and</strong> various<br />

maximum signal lags m. For all the parameter values we have checked, the weights decrease<br />

with the lag. However, while the approximate weights, γi 0 = σu 1<br />

σ w m+1<br />

, decrease at the<br />

same rate, in the Figure we see that the actual weights decrease less quickly for smaller lags,<br />

√ N+1<br />

m−i+1<br />

<strong>and</strong> then decrease faster for larger lags. When m is large, one can also see that the initial<br />

decrease in the actual weights is very small. 30<br />

An important consequence of the theorem is that the expected profit from each additional<br />

signal decays exponentially.<br />

Proposition 8. Consider the case with N speculators with equal speed, <strong>and</strong> let m be the<br />

maximum signal lag used by the speculators. Let π 0 be the expected profit at t = 0 of a speculator<br />

in equilibrium, <strong>and</strong> let γ i be his optimal trading weight on the signal with lag i = 0, . . . , m.<br />

Then the profit can be decomposed as follows:<br />

π 0<br />

= σ 2 w<br />

m∑<br />

j=0<br />

π 0,j<br />

= σ 2 w<br />

(<br />

γ0<br />

N + 1 + γ 1<br />

(N + 1) 2 + · · · γ m<br />

(N + 1) m+1 )<br />

. (69)<br />

29 Numerically, these conditions are satisfied for all the values of N <strong>and</strong> m we have tried.<br />

30 Thus, in the limit when m approaches infinity, we conjecture that the weights become approximately<br />

equal.<br />

∫<br />

In that case, the informed traders behave as in the Kyle (1985) model, by trading a multiple of the sum<br />

t<br />

dwτ = wt − w0. However, we can see from Theorem 4 that in our model the weights do not become of the<br />

0<br />

order of dt, as in the Kyle model, but rather remain of the same order of magnitude as for the lower m. In our<br />

model therefore prices are very close to strong form efficient when m is large. This equilibrium resembles that<br />

of Caldentey <strong>and</strong> Stacchetti (2010).<br />

30


Moreover, the ratio of two consecutive components of the expected profit is<br />

π 0,j+1<br />

= γ j+1 1<br />

π 0,j γ j N + 1 = O 1<br />

)<br />

N(<br />

N . (70)<br />

A graphic illustration of this result can be found in Figure 1, which plots the profits of<br />

a speculator who can trade on at most m = 5 past signals. The cases studied correspond to<br />

the number of speculators N ∈ {1, 2, 3, 5, 20, 100}. One sees that indeed, when N is large, the<br />

profits coming even from a signal of lag 1 are small.<br />

The next result analyzes in more detail price volatility, <strong>and</strong> shows that volatility has<br />

an upper bound, which makes rigorous the intuition for the general case, discussed after<br />

Proposition 6. Moreover, Proposition 9 provides a more thorough underst<strong>and</strong>ing about how<br />

information is revealed over time by trading. For this purpose, we define signal revelation as<br />

the covariance of a signal dw t with dp t+k , the price change from k trading periods later:<br />

SR k = Cov(dw t, dp t+k )<br />

σ 2 wdt<br />

= Cov(dw t−k, dp t )<br />

σwdt<br />

2 , k = 0, 1, . . . . (71)<br />

Since ∑ ∞<br />

k=0 dw t−k = w t (speculator’s initial forecast is w 0 = 0), the sum of all SR k equals<br />

∞∑<br />

k=0<br />

SR k = Cov(w t, dp t )<br />

σ 2 wdt<br />

= λ Cov(w t, dy t )<br />

σ 2 wdt<br />

= λ2 σ 2 y<br />

σ 2 w<br />

= σ2 p<br />

σw<br />

2 , (72)<br />

where we have used the formula Cov(w t , dy t ) = λ Var(dy t ) = λσy 2 from the dealer’s pricing<br />

equation for λ, proved in (I.23) in the Internet Appendix.<br />

Proposition 9. In the context of Theorem 4, price volatility is always smaller than the forecast<br />

volatility. Their difference is small when the number of speculators is large:<br />

σ 2 w − σ 2 p = O N<br />

( 1<br />

N<br />

)<br />

. (73)<br />

The signal revelation measure satisfies<br />

SR k =<br />

N<br />

m<br />

(N + 1) k+1 , k = 0, 1, . . . , m, =⇒ ∑<br />

SR k<br />

k=0<br />

= 1 −<br />

1<br />

. (74)<br />

(N + 1) m+1<br />

Therefore, the difference σw 2 − σp 2 is also small when m is large.<br />

Thus, an interesting implication of the Proposition is that, when the number of lags m is<br />

large, each signal dw t gets revealed by trading almost entirely. From Proposition 6, this case<br />

coincides with the one in which the growth rate of Σ, the forecast error variance, is very small.<br />

31


6.3 Robust Trading Strategies<br />

In this section, we discuss the assumption of an individual information processing cost, 31 <strong>and</strong><br />

we show that it is related to the notion of robust trading strategy. Intuitively, we think of<br />

a successful trading strategy as one which produces positive expected profits under many<br />

different scenarios <strong>and</strong> specifications, or conversely a strategy that does not depend too much<br />

on the the model assumptions being exactly true.<br />

A particularly strong assumption in many models of informed trading is that the fundamental<br />

value has only one component that traders learn about. For instance, in the influential<br />

Kyle (1985) model, the unique informed trader (the “insider”) uses private information to<br />

generate profits smoothly over time. The insider’s strategy is of the form 32<br />

dx t = β t (w t − p t )dt. (75)<br />

Thus, the insider compares his forecast with the price, <strong>and</strong> then buys slowly if his forecast is<br />

above the price, <strong>and</strong> sells otherwise. The implicit assumption in Kyle (1985) is that the price<br />

only contains information about his signal, <strong>and</strong> thus the insider has no inference problem: he<br />

knows his information to be superior to that of the public’s.<br />

We now relax this assumption, <strong>and</strong> allow the fundamental value to have more than one<br />

component, such that there exist informed traders that learn about each component. We then<br />

show that, under this alternative assumption, the smooth strategy described above starts<br />

losing money if the other component is large enough. In other words, smooth strategies<br />

are not robust. By contrast, we show that our fast strategies are robust to this alternative<br />

specification. For instance, consider the strategy that relies on the most recent signal:<br />

dx t = γ t dw t . (76)<br />

Then, Proposition 11 below shows that the expected profit from this strategy is positive, <strong>and</strong><br />

stays constant under all these specifications (taking the price impact coefficient λ as given).<br />

Intuitively, when the fundamental value has multiple components, a speculator who specializes<br />

in only one component is potentially adversely selected when using the price to decide<br />

his strategy. For instance, suppose the value of IBM has both a domestic <strong>and</strong> an international<br />

component. Then, suppose that a hedge fund that specializes only in the IBM’s domestic<br />

component uses a smooth strategy as in (75).<br />

Then, by buying <strong>and</strong> selling at the public<br />

price, the hedge fund essentially behaves as a noise trader with respect to the international<br />

31 We have seen that because of the individual cost, speculators much choose which signals to process, <strong>and</strong><br />

since the value of each signal decays quickly, after a short time the speculators must drop it. Thus, speculators<br />

optimally trade only on their most recent signals, which in turn drives the other results.<br />

32 In Kyle (1985) w t is in fact constant. However, Back <strong>and</strong> Pedersen (1998) show that the same type of<br />

strategies are optimal even if the fundamental value changes over time.<br />

32


component, <strong>and</strong> can therefore make losses on average. If instead, the hedge fund uses a fast<br />

strategy <strong>and</strong> buys if its signal about the domestic component is positive, its average profit is<br />

not affected by what happens in the international component.<br />

To formalize this intuition, we consider an extension of our model in which the fundamental<br />

value v t has two orthogonal components. To focus on the economic intuition, we choose the<br />

simple case with no lagged signals (m = 0), as in Proposition 3. Recall that in our baseline<br />

model, the speculators learn gradually about the terminal value v T by receiving signals of the<br />

form ds t = dv t + dη t . The speculators form at each t the forecast w t , which has increment<br />

dw t = ads t , with a =<br />

σ2 v<br />

. Then, v<br />

σv+σ 2 η<br />

2 T can be decomposed as a sum of orthogonal components,<br />

v T = w T + e T . Also, the increment dv t has the following orthogonal decomposition:<br />

dv t = dw t + de t , with σe 2 = σv 2 − σw 2 = σ2 vση<br />

2<br />

σv 2 + ση<br />

2 . (77)<br />

Proposition 10. Consider N w + N e speculators, of which N w speculators learn about w t (the<br />

w-speculators), <strong>and</strong> N e speculators learn about e t (the e-speculators).<br />

Denote the position<br />

in the risky asset of an w- or e-speculator, respectively, by x w,t <strong>and</strong> x e,t . Assume that the<br />

speculators can only trade on their most recent signal, dw t or de t , respectively. Then, there<br />

exists a unique linear equilibrium, in which the speculators’ trading strategies, <strong>and</strong> the dealer’s<br />

pricing functions are of the form:<br />

dx w,t = γ w dw t , dx e,t = γ e de t , dp t = λdy t , (78)<br />

with equilibrium coefficients <strong>and</strong> profits given by<br />

γ w<br />

= 1 λ<br />

1<br />

N w + 1 , γ e = 1 λ<br />

π w =<br />

(<br />

1<br />

N e + 1 , λ = N w σw<br />

2<br />

(N w + 1) 2 +<br />

σ 2 w<br />

λ(N w + 1) 2 , π e =<br />

σ 2 u<br />

N e σ 2 ) 1/2<br />

e<br />

(N e + 1) 2 σu<br />

2 ,<br />

(79)<br />

σw<br />

2<br />

λ(N e + 1) 2 .<br />

Proposition 10 shows that, taking λ as given, the w-speculators have indeed the same<br />

strategy, <strong>and</strong> the same expected profits, regardless of what the structure of the e-component.<br />

That is, the strategy <strong>and</strong> profits of the w speculators, do not depend on N e or σ e . However,<br />

the magnitude of the price impact coefficient λ does change with the specification, because of<br />

the increase in adverse selection in the other component.<br />

Proposition 11. In the context of Proposition 10, suppose now that one of the w-speculators,<br />

now called the β-speculator, adds a smooth component to his equilibrium trading strategy:<br />

dx β t = β t (w t−1 − p t−1 )dt + γ w dw t , (80)<br />

while the other traders <strong>and</strong> the dealer maintain their equilibrium strategies. Then, the expected<br />

33


profit of the β-speculator at t = 0 equals:<br />

π β = π β=0 + 1<br />

2λ<br />

∫ 1<br />

0<br />

( ) ( )<br />

1<br />

N e<br />

1 − εt (1 + ε t )<br />

N w + 1 σ2 w − (1 − ε t )<br />

N e + 1 σ2 e dt, (81)<br />

where<br />

π β=0 = π w =<br />

σw<br />

2 ( ∫<br />

λ(N w + 1) 2 , ε t = e −λ 1<br />

)<br />

t βτ dτ . (82)<br />

An important implication of Proposition 11 is that the profit π β depends on the specification<br />

of the model. To see this, consider first the case when σ e = 0, i.e., the fundamental<br />

value has only one component.<br />

σ 2 w<br />

2λ(N w+1)<br />

Then, the β-speculator increases his profit by β-trading:<br />

∫ 1<br />

0 (1 − ε2 t )dt > 0. Moreover, if as in Kyle (1985) the β-speculator sets β t = β 0<br />

1−t > 0,<br />

then ε t = 0, <strong>and</strong> the profit is maximized.<br />

If however, there is more than one component of the fundamental value (σ e > 0), then<br />

we show that β t > 0 produces a loss for certain values of σ e (<strong>and</strong> N e ). The condition β t > 0<br />

translates into ε t not being identically equal to 1, or equivalently ∫ 1<br />

0 (1 − ε t) 2 dt > 0. Choose<br />

a value σ e such that<br />

σ 2 e > N e + 1<br />

2<br />

(N +<br />

w+1) ∫ 2<br />

N 1<br />

e<br />

∫ 1<br />

0 (1−ε2 t )dt<br />

N w+1<br />

0 (1 − ε t) 2 dt<br />

σ 2 w. (83)<br />

Then, one verifies that π β < 0. In other words, the β-strategy is not robust to the fundamental<br />

value having more than one component.<br />

7 Conclusion<br />

We have presented a theoretical model in which traders continuously receive signals over time<br />

about the value of an asset, but can do so only at a cost per signal <strong>and</strong> per unit of time. We<br />

have found that competition among these speculators generates strategies that weigh heavily<br />

on the traders’ most recent signals. As a result, information decays fast, <strong>and</strong> a trader who<br />

is just one instant slower than the other traders loses the majority of the profits by being<br />

slow. Another consequence is that, because competition among speculators reveals most of<br />

their information to the public, the market is very efficient <strong>and</strong> liquid. As a feedback effect,<br />

because of the small price impact (high market liquidity), the informed traders are capable<br />

of trading even more aggressively. Thus, in equilibrium, the informed trading volume is very<br />

large <strong>and</strong> dominates the overall trading volume. Finally, a speculator who wants to have a<br />

small inventory at all times can do that almost infinitely fast <strong>and</strong> still make a positive profit.<br />

This, however, is possible only if there exist traders who have a “slower” component of trading,<br />

i.e., trade on their lagged signals. This is because the existence of slower traders mitigates the<br />

price impact of the speculator who wants to manage his inventory.<br />

34


A<br />

Proofs of Results<br />

A.1 Notation Preliminaries<br />

In general, a tilde above a symbol denotes normalization by σ w . For instance, if σ u is the<br />

instantaneous volatility of the noise trader order flow, <strong>and</strong> σ y the instantaneous volatility of<br />

the total order flow, we denote by<br />

˜σ u<br />

= σ u<br />

σ w<br />

, ˜σ y = σ y<br />

, with σu 2 = Var(du t)<br />

σ w dt<br />

, σy 2 = Var(dy t)<br />

. (A.1)<br />

dt<br />

Also, to be consistent with the general case of Theorem 3, we define the following numbers:<br />

A 1,1,t = Var( ˜dwt<br />

)<br />

σ 2 wdt<br />

= Ṽar( ˜dwt<br />

)<br />

, B1,t = Cov( w t , ˜dw t<br />

)<br />

σ 2 wdt<br />

= ˜Cov ( w t , ˜dw t<br />

)<br />

, (A.2)<br />

where here we use tilde notation to represent normalization by σ 2 w, but also by dt.<br />

Consider a trading strategy dx t for t ∈ (0, T ], where T = 1. 33 Denote by ˜π the normalized<br />

expected profit at t = 0 from the strategy dx t :<br />

A.2 Proof of Theorem 1<br />

˜π = 1<br />

σ 2 w<br />

∫ 1<br />

0<br />

(w t − p t )dx t . (A.3)<br />

We are interested in the existence of a linear equilibrium of the trading model with m =<br />

1 lagged signals. Thus, we look for an equilibrium with the following properties: (i) the<br />

equilibrium is symmetric, in the sense that the FTs have identical trading strategies, <strong>and</strong> the<br />

same for the STs; (ii) the equilibrium coefficients are constant with respect to time.<br />

To solve for the equilibrium, in the first step we take the dealer’s pricing functions as given,<br />

<strong>and</strong> solve for the optimal trading strategies for the FTs <strong>and</strong> STs. In the second step, we take<br />

the speculators’ trading strategies as given, <strong>and</strong> we compute the dealer’s pricing functions.<br />

In Section 2.1, we have assumed that the speculators take the signal covariance structure as<br />

given, which in the current context means that the speculators take the covariances A 1,1 <strong>and</strong><br />

B 1 from equation (A.2) as given (<strong>and</strong> constant). 34<br />

33 To alleviate potential confusion regarding the notation t − 1 for t − dt, we sometimes use T instead of 1.<br />

34 We can think of this assumption as saying that the dealer also sets A 1,1 <strong>and</strong> B 1, in addition to setting λ<br />

<strong>and</strong> ρ. We provide more discussion in the part of the proof that computes the dealer’s pricing rules.<br />

35


Speculator’s Optimal Strategy (γ, µ)<br />

Consider a FT, indexed by i = 1, . . . , N F . He chooses dx i t = γ i tdw t + µ i t ˜dw t−1 , <strong>and</strong> takes as<br />

given the dealer’s pricing rule. Thus, FT i assumes that at each t ∈ (0, T ], the price satisfies:<br />

dp t = λ dy t , with dy t = ( γ i t + γ −i<br />

t<br />

)<br />

dwt + ( µ i t + µ −i ) ˜dwt−1 + du t , (A.4)<br />

where the superscript “−i” indicates the aggregate quantity from the other speculators. Since<br />

dw t <strong>and</strong> ˜dw t−1 are both orthogonal on the public information set I t , <strong>and</strong> p t−1 ∈ I t , it follows<br />

that dx i t is orthogonal to p t−1 as well. Moreover, since we look for equilibria with constant coefficients,<br />

the FT assumes that Ṽar( ˜dwt−1<br />

)<br />

= A1,1 <strong>and</strong> ˜Cov ( w t , ˜dw t−1<br />

)<br />

= ˜Cov<br />

(<br />

wt−1 , ˜dw t−1<br />

)<br />

=<br />

B 1 are constant. Then, the normalized expected profit of FT i at t = 0 satisfies:<br />

˜π F = 1<br />

σw<br />

2 E<br />

∫ 1<br />

0<br />

= γ i t − λγ i t<br />

(<br />

(<br />

w t − p t−1 − λ (γt i + γt<br />

−i )dw t + (µ i t + µ −i<br />

t ) ˜dw<br />

) )<br />

t−1 + du t dx i t<br />

(<br />

γ<br />

i<br />

t + γ −i<br />

t<br />

)<br />

+ µ<br />

i<br />

t B 1 − λµ i t(<br />

µ<br />

i<br />

t + µ −i )<br />

A1,1 .<br />

t<br />

t<br />

(A.5)<br />

This is a pointwise optimization problem, hence it is enough to consider the profit at t = 0,<br />

<strong>and</strong> maximize the expression over γt i <strong>and</strong> µ i t. The solution of this problem is λγt i = 1−λγ−i t<br />

2<br />

,<br />

<strong>and</strong> λµ i t = B 1/A 1,1 −λµ −i<br />

t<br />

2<br />

. The ST solves the same problem, only that his coefficient on dw t is<br />

= 0. Thus, all γ’s are equal for the FTs, <strong>and</strong> all µ’s are equal for the FTs <strong>and</strong> STs. We<br />

γ j t<br />

also find that they are constant:<br />

γ = 1 λ<br />

Dealer’s Pricing Rules (λ, ρ, A 1,1 , B 1 )<br />

1<br />

, µ = B 1/A 1,1<br />

1 + N F λ<br />

1<br />

1 + N T<br />

(A.6)<br />

The dealer takes the speculators’ strategies as given, <strong>and</strong> assumes the aggregate order flow is<br />

of the form<br />

dy t = du t + ¯γ dw t + ¯µ ˜dw t−1 , with ¯γ = N F γ, ¯µ = N T µ, (A.7)<br />

where the speculators set, for k = 0, . . . , m,<br />

˜dw t−1 = dw t−1 − ρ ∗ dy t−1 , (A.8)<br />

36


with ρ ∗ constant. (Of course, in equilibrium the dealer will eventually set ρ t = ρ ∗ .) The order<br />

flow is orthogonal to I t , hence the dealer sets λ t , ρ t , A 1,1,t , B 1,t as follows:<br />

dp t = λ t dy t , λ t = ˜Cov(w t , dy t )<br />

Ṽar(dy t )<br />

˜dw t = dw t − ρ t dy t , ρ t = ˜Cov(dw t , dy t )<br />

Ṽar(dy t )<br />

σ 2 y,t = Ṽar(dy2 t ) = ˜σ 2 u + ¯γ 2 + ¯µ 2 A 1,1,t−1 ,<br />

= ¯γ + ¯µ B 1,t−1<br />

σy,t<br />

2 ,<br />

= ¯γ<br />

σy,t<br />

2 ,<br />

B 1,t = ˜Cov ( w t , dw t − ρ ∗ dy t<br />

)<br />

= (1 − ρ∗¯γ) − ρ ∗¯µB 1,t−1 ,<br />

(A.9)<br />

A 1,1,t<br />

= Ṽar( dw t − ρ ∗ dy t<br />

)<br />

= 1 − 2ρ∗¯γ + (ρ ∗ ) 2 σ 2 y,t<br />

= 1 − 2ρ ∗¯γ + (ρ ∗ ) 2 (˜σ 2 u + ¯γ 2 ) + (ρ ∗ ) 2¯µ 2 A 1,1,t−1 .<br />

Next, Lemma A.1 implies that A 1,1 does not depend on t, as long as |ρ ∗¯µ| < 1. But this<br />

follows from the equilibrium condition ρ¯µ = b ∈ (0, 1), to be proved below. Similarly, the<br />

same method shows that B 1 does not depend on t. For both A 1,1 <strong>and</strong> B 1 , Lemma A.1 implies<br />

that<br />

A 1,1<br />

= (1 − ρ∗¯γ) 2 + (ρ ∗ ) 2˜σ u<br />

2<br />

1 − (ρ ∗¯µ) 2 , B 1 = 1 − ρ∗¯γ . (A.10)<br />

1 + ρ∗¯µ Then, equation (A.9) shows that λ, ρ, ˜σ y are independent on t as well.<br />

We also provide justification for the assumption that the speculators take A 1,1 <strong>and</strong> B 1 as<br />

fixed when deciding their optimal strategies. Consider A 1,1 = Ṽar( ) ( ˜dwt = ˜Cov dwt , dw t −<br />

) ( )<br />

ρdy t = 1 − ρ˜Cov dwt , dy t . Then, one might think that by increasing the speculator’s coefficient<br />

γt i on dw t , the covariance ˜Cov ( )<br />

dw t , dy t would increase, which would then decrease A1,1 .<br />

But the covariance ˜Cov ( )<br />

dw t , dy t is part of the formula for ρ (see (A.9)), <strong>and</strong> ρ is considered<br />

fixed by the speculators, as the dealer’s coefficient for the expectation z t−1,t = E t (dw t−1 ). 35<br />

Equilibrium Conditions<br />

We now use the equations derived above to solve for the equilibrium coefficients γ, µ, λ, ρ = ρ ∗ ,<br />

˜σ y . Denote by<br />

a = ρ ¯γ, b = ρ ¯µ, R = λ ρ . (A.11)<br />

From (A.10) we have A 1,1 = (1−a)2 +ρ 2˜σ u<br />

2 . Then, substitute A<br />

1−b 2 1,1 in ˜σ y 2 = ˜σ u 2 + ¯γ 2 + ¯µ 2 A 1,1<br />

from (A.9), to obtain ρ 2˜σ y 2 = ρ2˜σ u+(a 2 2 +b 2 −2ab 2 )<br />

. To summarize,<br />

1−b 2<br />

B 1<br />

= 1 − a<br />

1 + b , A 1,1 = (1 − a)2 + ρ 2˜σ 2 u<br />

1 − b 2 , ρ 2˜σ 2 y = ρ2˜σ 2 u + (a 2 + b 2 − 2ab 2 )<br />

1 − b 2 . (A.12)<br />

35 In Section 6 we discuss an alternative setup in discrete time, for which speculators’ trading strategies do<br />

affect A 1,1 <strong>and</strong> B 1.<br />

37


Using (A.9), we get R = λ ρ = ¯γ+¯µB 1−a<br />

a+b<br />

1<br />

1+b<br />

¯γ<br />

=<br />

a<br />

= a+b<br />

a(1+b)<br />

. Also, the equation for ρ implies<br />

ρ = ¯γ˜σ = ρa<br />

y<br />

2 ρ 2˜σ . Using the formula for ρ 2˜σ 2 y<br />

2 y in (A.12), we compute ρ 2˜σ u 2 = (1 − a)(a − b 2 ).<br />

Using this formula, we obtain ρ 2˜σ y 2 = a <strong>and</strong> A 1,1 = 1 − a. To summarize,<br />

R = λ ρ =<br />

a + b<br />

a(1 + b) , ρ2˜σ 2 u = (1 − a)(a − b 2 ), ρ 2˜σ 2 y = a, A 1,1 = 1 − a. (A.13)<br />

A 1,1<br />

N T<br />

N T +1 = a+b<br />

1+b<br />

N F<br />

From (A.6), we have<br />

N F +1 = λ¯γ = λ ρ a = a+b<br />

1+b . From this, a = N F −b<br />

N F +1 , <strong>and</strong> B 1 = 1−a<br />

1<br />

N F +1 . Also, B 1 N T<br />

N T +1 = λ¯µ = λ ρ b = b(a+b)<br />

a(1+b) . Since B 1<br />

A 1,1<br />

= 1<br />

1+b , we have N T<br />

a<br />

b(1+b)<br />

. The two formulas for<br />

a+b<br />

1+b imply b(1 + b) N F<br />

N F +1 = a N T<br />

N T +1<br />

1+b<br />

1+b = N F +1<br />

1+b<br />

=<br />

N T +1 = b(a+b)<br />

a<br />

, or<br />

. To summarize,<br />

a = N F − b<br />

N F + 1 , B 1 =<br />

1<br />

N F + 1 , b(1 + b) N F<br />

N F + 1 = N F − b<br />

N F + 1<br />

N T<br />

N T + 1 .<br />

(A.14)<br />

Thus, we have the quadratic equation b 2 + bω =<br />

( )<br />

has a solution, b = −ω+ ω+4 N 1/2<br />

T<br />

N T +1<br />

N T<br />

N T +1 , with ω = 1 + 1<br />

N F<br />

N T<br />

N T +1<br />

. This equation<br />

( )<br />

b = ω+ ω+4 N 1/2<br />

T<br />

N T +1<br />

2<br />

> 1, which leads to a negative ˜σ y 2 (see (A.12)). We use again the formula<br />

λ<br />

N F<br />

2<br />

∈ (0, 1), as stated in the Theorem. The other solution is<br />

N F<br />

N F −b<br />

ρ a = N F +1 , as well as the formula for a in (A.14), to obtain λ ρ = , as desired. Some<br />

more algebraic manipulation proves the second part of the formula for ρ in (20), <strong>and</strong> the proof<br />

is now complete.<br />

The next result shows that under certain conditions, the solution of a recursive equation<br />

converges to a fixed number, regardless of the starting point.<br />

Lemma A.1. Let x 1 ∈ R, <strong>and</strong> consider a sequence x t ∈ R which satisfies the following<br />

recursive equation:<br />

Then the sequence x t converges to ¯x =<br />

β ∈ (−1, 1).<br />

x t − βx t−1 = α, t ≥ 2. (A.15)<br />

α<br />

1−β , regardless of the initial value of x 1, if <strong>and</strong> only if<br />

Proof. First, note that ¯x is well defined as long as β ≠ 1. If we denote by y t = x t − ¯x, the new<br />

sequence y t satisfies the recursive equation y t −βy t−1 = 0. We now show that y t converges to 0<br />

(<strong>and</strong> ¯x is well defined) if <strong>and</strong> only if β ∈ (−1, 1). Then, the difference equation y t − βy t−1 = 0<br />

has the following general solution:<br />

y t = Cβ t , t ≥ 1, with C ∈ R. (A.16)<br />

Then, y t is convergent for any values of C if <strong>and</strong> only if all β ∈ (−1, 1]. But in the latter case,<br />

1 − β = 0, which makes ¯x nondefined.<br />

38


A.3 Proof of Corollary 1<br />

When N F is large, a =<br />

N F<br />

N F −b ≈ 1, ω = 1 + 1<br />

N F<br />

follows from the formulas in Theorem 1.<br />

N T<br />

N T +1 ≈ 1, hence b ≈ b ∞ =<br />

√<br />

5−1<br />

2<br />

. The rest<br />

A.4 Proof of Corollary 2<br />

In the proof of Theorem 1, equation (A.6) implies λ¯γ =<br />

N F<br />

N F +1 , λ¯µ = B 1<br />

A 1,1 N T +1 . But<br />

from (A.12) <strong>and</strong> (A.13), we have B 1<br />

A 1,1<br />

= 1<br />

1+b<br />

, which proves the first row in (25). The second<br />

row in (25) just rewrites the formulas for A 1,1 <strong>and</strong> B 1 from equations (A.12) <strong>and</strong> (A.13).<br />

N T<br />

A.5 Proof of Proposition 1<br />

In the proof of Theorem 1, equation (A.6) implies λ¯γ =<br />

the equilibrium expected profit of the FT is<br />

N F<br />

N F +1 , λ¯µ = B 1<br />

A 1,1<br />

N T<br />

N T +1<br />

. From (A.5),<br />

(<br />

˜π F = γ − λγ¯γ + µB 1 − λµ¯µA 1,1 = γ 1 − N ) (<br />

F<br />

+ B 1 µ 1 − N )<br />

T<br />

N F + 1<br />

N T + 1<br />

(A.17)<br />

From (A.14), B 1 = 1<br />

N F +1 , which proves the desired formula for πF . The profit of the ST is<br />

the same as for the FT, but with γ = 0. The last statement follows by using the asymptotic<br />

results in Corollary 1.<br />

A.6 Proof of Proposition 2<br />

This follows from simple algebraic manipulation of the formulas for the proof of Theorem 1.<br />

A.7 Proof of Proposition 3<br />

As in Theorem 1, we start with the speculators’ choice of optimal trading strategy.<br />

Each<br />

speculator i = 1, . . . , N observes dw t , <strong>and</strong> chooses dx i t = γ i tdw t to maximize the expected<br />

profit:<br />

π 0<br />

(∫ 1 (<br />

)<br />

= E w t − p t−1 − λ t (dx i t + dx −i<br />

t + du t )<br />

=<br />

∫ 1<br />

0<br />

0<br />

γ i tσ 2 wdt − λ t γ i t<br />

(<br />

γ<br />

i<br />

t + γt<br />

−i )<br />

σ<br />

2<br />

w dt,<br />

dx i t<br />

)<br />

(A.18)<br />

where the superscript “−i” indicates the aggregate quantity from the other N − 1 speculators.<br />

This is a pointwise quadratic optimization problem, with solution λ t γt i = 1−λtγ−i t<br />

2<br />

. Since this<br />

is true for all i = 1, . . . , N, the equilibrium is symmetric <strong>and</strong> we compute γ t = 1 1<br />

λ t 1+N .<br />

The dealer takes the speculators’ strategies as given, thus assumes that the aggregate order<br />

flow is of the form dy t = du t + Nγ t dw t . To set λ t , the dealer sets p t such that the market is<br />

39


efficient, which implies<br />

dp t = λ t dy t , with λ t = Cov(w t, dy t )<br />

Var(dy t )<br />

=<br />

Nγ t σw<br />

2<br />

σu 2 + N 2 γt 2 . (A.19)<br />

σ2 w<br />

This implies λ 2 t σu 2 + (Nγ t λ t ) 2 σw 2 = Nγ t λ t σw. 2 But Nλ t γ t =<br />

N<br />

N+1 . Hence, λ2 t σu 2 + ( N 2σ 2<br />

N+1)<br />

w =<br />

N<br />

N+1 σ2 w, or λ 2 t σu 2 N<br />

= σ 2 √<br />

(N+1) w, which implies the formula λ = σw N<br />

2 σ u N+1<br />

. We then compute<br />

γ t = 1 N<br />

λ t 1+N = √ ˜σu<br />

N<br />

.<br />

√<br />

We have also seen that TV = σy 2 = σu 2 + N 2 γ 2 σw. 2 But Nγ = ˜σ u N, hence TV =<br />

σ 2 u(1 + N). Next, σ 2 p = λ 2 TV = σ2 w<br />

σ 2 u (N+1)−σ2 u<br />

σ 2 u (N+1) = N<br />

N+1 .<br />

A.8 Proof of Proposition 4<br />

σ 2 u<br />

N<br />

σ<br />

(N+1) u(N 2 + 1) = σ 2 2 w<br />

N<br />

N+1 . Finally, SPR = T V −σ2 u<br />

TV<br />

=<br />

Using the notations <strong>and</strong> formulas from the Proof of Theorem 1, recall that A 1,1 = 1 − a. Since<br />

˜dw t is orthogonal on dy t , we have ˜Cov ( ˜dwt , dw t<br />

)<br />

= ˜Cov<br />

( ˜dwt , ˜dw t<br />

)<br />

= A1,1 . Denote by d¯x t the<br />

aggregate speculator order flow, which is of the form d¯x t = ¯γdw t +¯µ ˜dw t−1 . Also, the total order<br />

flow is dy t = d¯x t + du t . Compute ˜Cov ( ˜dwt , ˜dw t−1<br />

)<br />

= ˜Cov<br />

(<br />

dwt − ρ¯γdw t − ρ¯µ ˜dw t−1 , ˜dw t−1<br />

)<br />

=<br />

−ρ¯µA 1,1 . Therefore,<br />

˜Cov ( d¯x t+1 , d¯x t<br />

)<br />

= ˜Cov<br />

(¯γdwt+1 + ¯µ ˜dw t , ¯γdw t + ¯µ ˜dw t−1<br />

)<br />

= ¯µ¯γA1,1 + ¯µ 2( −bA 1,1<br />

)<br />

Ṽar ( d¯x t<br />

)<br />

= Ṽar (¯γdw t + ¯µ ˜dw t−1<br />

)<br />

= ¯γ 2 + ¯µ 2 A 1,1 .<br />

(A.20)<br />

From (A.14) we have a = N F −b<br />

1+b<br />

N F +1<br />

, which implies 1 − a =<br />

N F +1 . Consider the ratio of the<br />

formulas in (A.20), <strong>and</strong> divide <strong>and</strong> multiply by ρ 2 . Then, Corr ( )<br />

d¯x t+1 , d¯x t =<br />

ab−b 3<br />

a 2 +b 2 (1−a) (1 −<br />

a) = (a−b2 )b<br />

b ≈ b ∞ =<br />

a 2 +b 2 (1−a)<br />

References<br />

1+b<br />

N F +1 , which proves the desired formula. When N F is large, a ≈ 1, <strong>and</strong><br />

√<br />

5−1<br />

2<br />

, which solves b ∞ (1 + b ∞ ) = 1. Hence, Corr ( d¯x t+1 , d¯x t<br />

)<br />

≈<br />

1−b 2 ∞<br />

N F<br />

= b∞<br />

N F<br />

.<br />

[1] Aït-Sahalia, Yacine, <strong>and</strong> Mehment Saglam (2014): “<strong>High</strong> <strong>Frequency</strong> <strong>Traders</strong>: Taking Advantage of<br />

Speed,” Working Paper.<br />

[2] Back, Kerry, Henry Cao, <strong>and</strong> Gregory Willard (2000): “Imperfect Competition among Informed<br />

<strong>Traders</strong>,” Journal of Finance, 55, 2117–2155.<br />

[3] Back, Kerry, <strong>and</strong> Hal Pedersen (1998): “Long-Lived Information <strong>and</strong> Intraday Patterns,” Journal of<br />

Financial Markets, 1, 385–402.<br />

[4] Boehmer, Ekkehart, Kingsley Fong, <strong>and</strong> Julie Wu (2014): “International Evidence on Algorithmic<br />

Trading,” Working Paper.<br />

[5] Brogaard, Jonathan (2011): “<strong>High</strong> <strong>Frequency</strong> Trading <strong>and</strong> Its Impact on Market Quality,” Working<br />

Paper.<br />

[6] Baron, Matthew, Jonathan Brogaard, <strong>and</strong> Andrei Kirilenko (2014): “Risk <strong>and</strong> Return in <strong>High</strong><br />

<strong>Frequency</strong> <strong>Traders</strong>,” Working Paper.<br />

[7] Benos, Evangelos, <strong>and</strong> Satchit Sagade (2013): “<strong>High</strong>-<strong>Frequency</strong> Trading Behaviour <strong>and</strong> Its Impact<br />

on Market Quality: Evidence from the UK Equity Market,” Working Paper.<br />

40


[8] Biais, Bruno, Thierry Foucault, <strong>and</strong> Sophie Moinas (2014): “Equilibrium Fast Trading,” Working<br />

Paper.<br />

[9] Brogaard, Jonathan, Terrence Hendershott, <strong>and</strong> Ryan Riordan (2014): “<strong>High</strong>-<strong>Frequency</strong> Trading<br />

<strong>and</strong> Price Discovery,” Review of Financial Studies, 27, 2267–2306.<br />

[10] Budish, Eric, Peter Cramton, <strong>and</strong> John Shim (2014): “The <strong>High</strong>-<strong>Frequency</strong> Trading Arms Race:<br />

Frequent Batch Auctions as a Market Design Response,” Working Paper.<br />

[11] Caldentey, René, <strong>and</strong> Ennio Stacchetti (2010): “Insider Trading with a R<strong>and</strong>om Deadline,” Econometrica,<br />

78, 245–283.<br />

[12] Cartea, Álvaro, <strong>and</strong> José Penalva (2012): “Where is the Value in <strong>High</strong> <strong>Frequency</strong> Trading,” Quarterly<br />

Journal of Finance, 2, 1–46.<br />

[13] Chaboud, Alain, Benjamin Chiquoine, Erik Hjalmarsson, <strong>and</strong> Clara Vega (2014): “Rise of the<br />

Machines: Algorithmic Trading in the Foreign Exchange Market,” Journal of Finance, forthcoming.<br />

[14] Chau, Minh, <strong>and</strong> Dimitri Vayanos (2008): “Strong-Form Efficiency with Monopolistic Insiders,” Review<br />

of Financial Studies, 18, 2275–2306.<br />

[15] Foster, Douglas, <strong>and</strong> S. Viswanathan (1996): “Strategic Trading When Agents Forecast the Forecast<br />

of Others,” Journal of Finance, 51, 1437–1478.<br />

[16] Foucault, Thierry, Johan Hombert, <strong>and</strong> Ioanid Roşu (2013): “<strong>News</strong> Trading <strong>and</strong> Speed,” Working<br />

Paper.<br />

[17] Hasbrouck, Joel, <strong>and</strong> Gideon Saar (2013): “Low-Latency Trading,” Journal of Financial Markets,<br />

16, 646–679.<br />

[18] Hendershott, Terrence, Charles Jones, <strong>and</strong> Albert Menkveld (2011): “Does Algorithmic Trading<br />

Improve Liquidity,” Journal of Finance, 66, 1–33.<br />

[19] Hirschey, Nicholas (2013): “Do <strong>High</strong>-<strong>Frequency</strong> <strong>Traders</strong> Anticipate Buying <strong>and</strong> Selling Pressure,”<br />

Working Paper.<br />

[20] Hoffmann, Peter (2013): “A Dynamic Limit Order Market with Fast <strong>and</strong> Slow <strong>Traders</strong>,” Working<br />

Paper.<br />

[21] Holden, Craig, <strong>and</strong> Avanidhar Subrahmanyam (1992): “Long-Lived Private Information <strong>and</strong> Imperfect<br />

Competition,” Journal of Finance, 47, 247–270.<br />

[22] Jovanovic, Boyan, <strong>and</strong> Albert Menkveld (2012): “Middlemen in Limit-Order Markets,” Working<br />

Paper.<br />

[23] Kirilenko, Andrei, Albert Kyle, Mehrdad Samadi, <strong>and</strong> Tugkan Tuzun (2014): “The Flash Crash:<br />

The Impact of <strong>High</strong> <strong>Frequency</strong> Trading on an Electronic Market,” Working Paper.<br />

[24] Kyle, Albert (1985): “Continuous Auctions <strong>and</strong> Insider Trading,” Econometrica, 53, 1315–1335.<br />

[25] Li, Su (2012): “Speculative Dynamics I: Imperfect Competition, <strong>and</strong> the Implications for <strong>High</strong> <strong>Frequency</strong><br />

Trading,” Working Paper.<br />

[26] Li, Wei (2014): “<strong>High</strong> <strong>Frequency</strong> Trading with Speed Hierarchies,” Working Paper.<br />

[27] Pagnotta, Emiliano, <strong>and</strong> Thomas Philippon (2013): “Competing on Speed,” Working Paper.<br />

[28] Weller, Brian (2013): “Fast <strong>and</strong> Slow Liquidity,” Working Paper.<br />

[29] Zhang, X. Frank (2010): “<strong>High</strong> <strong>Frequency</strong> Trading, Stock <strong>Volatility</strong>, <strong>and</strong> Price Discovery,” Working<br />

Paper.<br />

41


z t−i,t of dw t−i : 36 ρ = [ ρ 0 , . . . , ρ m<br />

] ′, (I.2)<br />

Internet Appendix for “Fast <strong>and</strong> Slow Informed Trading” – Not<br />

for publication<br />

I<br />

Proofs for the General Case<br />

I.1 Notation Preliminaries<br />

To simplify formulas, in some of the proofs we use matrix notation. We denote by M a,b the<br />

set of matrices of real numbers with a rows <strong>and</strong> b columns, by M a = M a,a the set of square<br />

matrices, <strong>and</strong> by V a = M a,1 the set of column vectors. Let X ′ denotes the transpose of X.<br />

In our most general setup, speculators observe signals at different speeds. Thus, we search<br />

for an equilibrium which is symmetric in the sense that the l-speculators choose the same<br />

weights for the same l ∈ {0, 1, . . . , m}. Denote by γ (l) ∈ V m+1 as follows: if i ≥ l, γ (l)<br />

i<br />

is the<br />

l-speculator’s equilibrium weight on dw t−i − z t−i,t ; otherwise, γ (l)<br />

i<br />

= 0. Define the aggregate<br />

speculator weights, ¯γ ∈ V m+1 :<br />

¯γ =<br />

m∑<br />

l=0<br />

N l γ (l) , with γ (l) = [ γ (l)<br />

0 , . . . , γ m<br />

(l) ] ′. (I.1)<br />

Let ρ ∈ V m+1 be the vector which collects the coefficients involved in the dealer’s expectation<br />

In general, a tilde above a symbol denotes normalization by σ w . Thus, denote by<br />

˜σ u<br />

= σ u<br />

, ˜σ y = σ y<br />

, with σy 2 = Var(dy t)<br />

. (I.3)<br />

σ w σ w dt<br />

For i = 0, . . . , m, let d t w t−i be the unpredictable part of the signal dw t−i (computed before<br />

trading at t):<br />

d t w t−i = dw t−i − z t−i,t . (I.4)<br />

Define the matrices A ∈ M m+1 <strong>and</strong> B ∈ V m+1 . For i, j = 0, . . . , m, let<br />

A i,j = ˜Cov(d t w t−i , d t w t−j ) = 1<br />

σ 2 wdt Cov(d tw t−i , d t w t−j ),<br />

B j = ˜Cov(w t , d t w t−j ) = 1<br />

σ 2 wdt Cov(w t, d t w t−j ).<br />

(I.5)<br />

36 The last entry, ρ m, is not part of any of dealer’s expectations z t−i,t, but we introduce it in order to simplify<br />

notation.<br />

1


We now rescale ¯γ, ρ, λ, by defining r, g ∈ V m+1 <strong>and</strong> Λ ∈ R as follows:<br />

g = ¯γ˜σ y<br />

, r = ˜σ y ρ, Λ = ˜σ y λ. (I.6)<br />

I.2 Proof of Theorem 3<br />

Using the notation from Section I.1, we need to prove that a linear equilibrium exists if there<br />

exists a solution (g, r, Λ, ˜σ y , A, B) to the following system of equations:<br />

g =<br />

m∑<br />

( )<br />

( ) −1 1<br />

N l A≥l<br />

Λ B − r ,<br />

≥l ≥l<br />

l=0<br />

A i,j<br />

r = Ag, Λ = g ′ B, ˜σ 2 y =<br />

˜σ 2 u<br />

1 − g ′ r ,<br />

min(i,j)<br />

∑<br />

= 1 i=j − r i−k r j−k , B i = 1 − Λ<br />

k=1<br />

i∑<br />

r i−k .<br />

k=1<br />

(I.7)<br />

Recall that 1 P is the indicator function, which equals 1 if P is true, <strong>and</strong> 0 otherwise. Also,<br />

A ≥l<br />

is the matrix with elements A i,j for i, j ≥ l; <strong>and</strong> similarly for the vectors B ≥l<br />

<strong>and</strong> r ≥l<br />

.<br />

The sum of vectors X ≥l over different l is carried by padding X ≥l<br />

with zeros for the first l<br />

entries.<br />

Speculator’s Optimal Strategy (γ)<br />

We begin by analyzing the optimal strategy of a l-speculator, where l ∈ {0, . . . , m}. This<br />

speculator takes as given (i) the dealer’s pricing rules: dp t = λdy t <strong>and</strong> z t−i,t = ρ 0 dy t−i + · · · +<br />

ρ i−1 dy t−1 for i = 0, 1, . . . , m; <strong>and</strong> (ii) the other speculators’ trading strategies. For instance, if<br />

another speculator is of type k, he is assumed to trade according to dx (k)<br />

t = ∑ m<br />

j=k γ(k) j<br />

d t w t−j .<br />

Also, as we have discussed, the l-speculator chooses among trading strategies of the form:<br />

dx t = γ l,t d t w t−l + · · · + γ m,t d t w t−m . Therefore, the l-speculator assumes that the total order<br />

flow at t satisfies<br />

where<br />

∑l−1<br />

dy t = du t + ¯γ j d t w t−j +<br />

j=0<br />

m∑ (<br />

γj,t + γj<br />

− )<br />

dt w t−j . (I.8)<br />

j=l<br />

¯γ j =<br />

j∑<br />

k=0<br />

N k γ (k)<br />

j<br />

, j = 0, . . . , m, γ − j<br />

= (N l − 1)γ (l)<br />

j<br />

+<br />

j∑<br />

k=0<br />

k≠l<br />

N k γ (k)<br />

j<br />

, j = l, . . . , m. (I.9)<br />

At t = 0, equation (10) implies that his normalized expected profit is<br />

˜π 0 = π 0<br />

σ 2 w<br />

= 1<br />

σ 2 w<br />

(∫ 1<br />

E<br />

0<br />

(<br />

wt − p t−1 − λdy t<br />

)<br />

dxt<br />

)<br />

. (I.10)<br />

2


By construction, the terms d t w t−j are orthogonal to I t , hence also to p t−1 . Hence, dx t is<br />

also orthogonal to p t−1 . We now use (I.8) <strong>and</strong> the definitions: A i,j = ˜Cov(d t w t−i , d t w t−j ),<br />

B j = ˜Cov(w t , d t w j ) to compute:<br />

˜π 0<br />

⎛<br />

= E ⎝<br />

=<br />

j=l<br />

∫ 1<br />

⎞<br />

( ∑l−1<br />

m∑ (<br />

w t − λ ¯γ i d t w t−i − λ γi,t + γi<br />

− ) ) ∑<br />

m<br />

dt w t−i γ j,t d t w t−j<br />

⎠<br />

0<br />

i=0<br />

i=0 j=l<br />

i=l<br />

m∑ ∑l−1<br />

m∑<br />

B j γ j,t − λ ¯γ i A i,j γ j,t − λ<br />

i,j=l<br />

j=l<br />

m∑ (<br />

γi,t + γi<br />

− )<br />

Ai,j γ j,t .<br />

(I.11)<br />

Thus, we have reduced the problem to a linear–quadratic optimization. The first order condition<br />

with respect to γ k,t , for k = l, . . . , m, is:<br />

∑l−1<br />

m∑ (<br />

B k − λ ¯γ i A i,k − λ 2γi,t + γi<br />

−<br />

i=0<br />

i=l<br />

)<br />

Ai,k = 0. (I.12)<br />

Denote by γ t the (m − l + 1)-column vector of trading weights at t. We divide the matrix A<br />

into four blocks, by restricting indices to be either < l or ≥ l. With matrix notation, the first<br />

order condition (I.13) becomes<br />

B ≥l<br />

− λ A ≥l,


0, . . . , m, padding with zeroes where necessary. We get<br />

¯γ =<br />

m∑<br />

N l γ (l) =<br />

l=0<br />

m∑<br />

l=0<br />

N l<br />

(<br />

A≥l<br />

) −1<br />

( 1<br />

λ B ≥l − ˜σ2 y ρ ≥l<br />

)<br />

. (I.17)<br />

After dividing this equation by ˜σ y , use g = ¯γ˜σ y<br />

, r = ˜σ y ρ, <strong>and</strong> Λ = ˜σ y λ to obtain the<br />

corresponding equation in (I.7).<br />

So far, we have shown that equation (I.16) is a necessary condition for equilibrium. We<br />

now prove that it is also sufficient condition for the speculator’s problem, if one imposes two<br />

additional conditions: (i) λ > 0; <strong>and</strong> (ii) for all l = 0, . . . , m, the matrix A ≥l<br />

is invertible.<br />

Indeed, the second order condition in the maximization problem above for the l-speculator is:<br />

λ det(A ≥l<br />

) > 0.<br />

(I.18)<br />

Normally, we expect that det(A ≥l<br />

) > 0, since economically A is the covariance matrix of the<br />

fresh signals, <strong>and</strong> the signals dw t−i are independent. But if A is just the solution of a system<br />

of equations, this condition needs to be checked. If det(A ≥l<br />

) > 0, then the second order<br />

condition from (I.18) becomes λ > 0, which is just condition (i).<br />

Dealer’s Pricing Rules (λ, ρ, A, B)<br />

The dealer takes as given that the aggregate order flow is of the form:<br />

dy t = du t + ¯γ 0 dw t + ¯γ 1 d t w t−1 + · · · + ¯γ m d t w t−m , (I.19)<br />

where the speculators set, for k = 0, . . . , m,<br />

d t w t−k = dw t−k − ( ρ ∗ 0dy t−k + · · · + ρ ∗ k−1 dy t−1)<br />

, (I.20)<br />

with ρ ∗ i constant. (Of course, in equilibrium the dealer will eventually set ρ i = ρ ∗ i .) We<br />

combine the two equations:<br />

dy t = du t +<br />

= du t +<br />

m∑<br />

¯γ k dw t−k −<br />

k=0<br />

m∑<br />

¯γ k dw t−k −<br />

k=0<br />

m∑<br />

k=0<br />

m∑<br />

∑<br />

ρ ∗ i dy t−k+i<br />

¯γ k<br />

k−1<br />

m−k<br />

∑<br />

k=1 i=0<br />

i=0<br />

ρ ∗ i ¯γ k+i dy t−k .<br />

(I.21)<br />

Because each speculator only trades on the unpredictable part of his signal, dy t are or-<br />

4


thogonal to each other. Thus, the dealer computes<br />

= E ( ) ∑k−1<br />

dw t−k | dy t−k , . . . , dy t−1 = ρ i,t−k dy t−k+i , k = 0, . . . , m,<br />

z t−k,t<br />

i=0<br />

dp t = λ t dy t ,<br />

(I.22)<br />

where the coefficients ρ i,t−k <strong>and</strong> λ t are given by: 37<br />

ρ i,t−k = Cov(dw t−k, dy t−k+i )<br />

, k = 0, . . . , m, i = 0, . . . , k − 1,<br />

Var(dy t−k+i )<br />

λ t = Cov(v 1, dy t )<br />

Var(dy t )<br />

= Cov(w t, dy t )<br />

.<br />

Var(dy t )<br />

(I.23)<br />

At the end of this proof, we show that the following numbers do not depend on t: λ t , ρ i,t ,<br />

A i,j,t = ˜Cov(d t w t−i , d t w t−j ), B j,t = ˜Cov(w t , d t w j ).<br />

Taking these numbers as constant, we now prove the rest of the equations in (I.7). First,<br />

note that in equilibrium ρ ∗ i = ρ i . We rewrite the equation for ρ in (I.23) by taking k = i.<br />

Thus, ρ i = ˜Cov(dw t−i ,dy t)<br />

, <strong>and</strong> note that dy t is orthogonal on all other dy t−k for k > 0. Hence,<br />

Ṽar(dy t)<br />

from (I.19) <strong>and</strong> (I.20), we get<br />

ρ i = ˜Cov(d t w t−i , dy t )<br />

Ṽar(dy t )<br />

=<br />

˜Cov ( d t w t−i , ∑ m<br />

j=0 ¯γ jd t w t−j<br />

)<br />

Ṽar(dy t )<br />

=<br />

∑ m<br />

j=0 A i,j¯γ j<br />

. (I.24)<br />

Since we have denoted by g = ¯γ˜σ y<br />

<strong>and</strong> r = ˜σ y ρ, we get r i = (Ag) i , or in matrix notation<br />

r = Ag. This proves the corresponding equation in (I.7). Also, from the equation from λ<br />

in (I.23), we get<br />

λ = ˜Cov(w t , dy t )<br />

Ṽar(dy t )<br />

=<br />

˜Cov ( w t , ∑ m<br />

j=0 ¯γ jd t w t−j<br />

)<br />

Ṽar(dy t )<br />

=<br />

˜σ 2 y<br />

∑ m<br />

j=0 ¯γ jB j<br />

. (I.25)<br />

Since Λ = λ ˜σ y , we obtain Λ = ∑ m<br />

j=0 g jB j , or in matrix notation Λ = g ′ B. This proves the<br />

corresponding equation in (I.7).<br />

By computing ˜Cov(dy t , dy t ) <strong>and</strong> using (I.19), it follows that ˜σ 2 y = Ṽar(dy t) satisfies ˜σ 2 y =<br />

˜σ 2 u + ∑ m<br />

i,j=0 A i,j¯γ i¯γ j , or in matrix notation ˜σ 2 y = ˜σ 2 u + ¯γ ′ A¯γ. Since ¯γ = g ˜σ y , we compute<br />

˜σ 2 y = ¯γ ′ A¯γ + ˜σ 2 u = g ′ Ag ˜σ 2 y + ˜σ 2 u. (I.26)<br />

˜σ 2 y<br />

But we have already shown that Ag = r, hence ˜σ 2 y = g ′ r ˜σ 2 y + ˜σ 2 u, which implies ˜σ 2 y =<br />

This proves the corresponding equation in (I.7).<br />

˜σ2 u<br />

1−g ′ r .<br />

Now, consider the equation A k,l = ˜Cov(d t w t−k , d t w t−l ). From (I.20), d t w t−k = dw t−k −<br />

37 Note that in principle ρ i,t−k might also depend on t, the time at which the expectation is computed.<br />

However, the formula shows that ρ only depends on i <strong>and</strong> t − k, <strong>and</strong> not on t separately.<br />

5


(<br />

ρ0 dy t−k + · · · + ρ k−1 dy t−1<br />

)<br />

. But dt w t−l is orthogonal to the previous order flow, hence<br />

A k,l = ˜Cov(dw t−k , d t w t−l ). Because A is a symmetric matrix, without loss of generality<br />

assume k ≥ l, which implies l = min(k, l). Since ˜Cov(dw t−i , dy t ) = ρ i˜σ 2 y 1 i≥0 , we get<br />

A k,l<br />

= ˜Cov<br />

(dw t−k , dw t−l − ( ) )<br />

ρ 0 dy t−l + · · · + ρ l−1 dy t−1<br />

= 1 k=l −<br />

∑l−1<br />

ρ j ρ k−l+j ˜σ y 2 = 1 k=l −<br />

j=0<br />

l∑<br />

ρ k−i ρ l−i ˜σ y.<br />

2<br />

i=1<br />

(I.27)<br />

Since r = ρ ˜σ y , we get A k,l = 1 k=l − ∑ l<br />

j=1 r k−ir l−i , which proves the corresponding equation<br />

in (I.7). We also compute B l = ˜Cov(w t , d t w l ). From (I.20),<br />

B l<br />

= ˜Cov<br />

(w t , dw t−l − ( ) )<br />

ρ 0 dy t−l + · · · + ρ l−1 dy t−1<br />

= 1 − λ ( ) (I.28)<br />

ρ 0 + · · · + ρ l−1<br />

Since Λ = λ ˜σ y <strong>and</strong> r = ρ ˜σ y , we obtain B l = 1 − Λ ∑ l−1<br />

j=0 r j = 1 − Λ ∑ l<br />

i=1 r i−k. This proves<br />

the corresponding equation in (I.7).<br />

We now prove that the various pricing coefficients do not depend on t. For this, we show<br />

that the following numbers are independent of t: Cov(dw t , dy t+k ) for all k; Cov(w t , dy t+k ) for<br />

all k; Var(dy t ); Cov(w t , d t w j ) for j = 0, . . . , m; <strong>and</strong> Cov(d t w i , d t w j ) for i, j = 0, . . . , m.<br />

First, we prove by induction that ˜Cov(dw t , dy t+k ) does not depend on t for k ≥ 0. (This<br />

is trivially true for k < 0.) The statement is true for k = 0, since equation (I.19) implies<br />

˜Cov(dw t , dy t ) = ¯γ 0 . Assume that the statement is is true for all i < k. We now prove that<br />

˜Cov(dw t , dy t+k ) does not depend on t. Equations (I.21) implies that dy t+k only involves three<br />

types of terms: (i) du t+k , (ii) dw t+k−i for i = 0, . . . , m, <strong>and</strong> (iii) dy t+k−1−i for i = 0, . . . , m−1.<br />

Also, the coefficients ρ ∗ i do not depend on time. Therefore, by the induction hypothesis all<br />

these terms have covariances with dw t that do not depend on t.<br />

Next, we prove that a t = ˜Cov(w t , dy t ) does not depend on t. Equation (I.21) implies the<br />

following recursive formula for all t:<br />

a t =<br />

m∑<br />

¯γ k −<br />

k=0<br />

m∑<br />

m−k<br />

∑<br />

k=1 i=0<br />

ρ ∗ i ¯γ k+i a t−k .<br />

(I.29)<br />

But Lemma I.2 implies that a t does not depend on t, provided that<br />

1 > β 1 > · · · > β m > 0, with β k =<br />

m−k<br />

∑<br />

i=0<br />

ρ ∗ i ¯γ k+i .<br />

(I.30)<br />

Therefore, ˜Cov(w t , dy t ) does not depend on t. This result also implies that ˜Cov(w t , dy t+k )<br />

does not depend on t for any integer k. To see this, note first that the case k > 0 reduces to<br />

6


the case k = 0. Indeed, ˜Cov(w t , dy t+k ) = ˜Cov(w t+k , dy t+k ) − ∑ k<br />

i=1 ˜Cov(dw t+i , dy t+k ), <strong>and</strong> we<br />

already proved that all these terms are independent of t. Also, the case k < 0 reduces to the<br />

case k = 0, since ˜Cov(w t , dy t−i ) = ˜Cov(w t−i , dy t−i ) if i ≥ 0.<br />

We now prove that Ṽar(dy t) = ∑ m<br />

k=0 ¯γ k Cov ( d t w t−k , dy t<br />

)<br />

does not depend on t. Since dyt<br />

is orthogonal to previous order flow, Ṽar(dy t) = ∑ m<br />

k=0 ¯γ k Cov ( dw t−k , dy t<br />

)<br />

. But these terms<br />

have already been proved to be independent of t.<br />

Finally, one uses the results proved above to show that B j,t = ˜Cov(w t , d t w j ) <strong>and</strong> A i,j,t =<br />

˜Cov(d t w i , d t w j ) do not depend on t. Indeed, we have already shown that ˜Cov(dw t , dy t−k ) <strong>and</strong><br />

˜Cov(w t , dy t−k ) are independent of t, <strong>and</strong> all is left to do is to use the fact that dy t−k are<br />

orthogonal to each other.<br />

So far, we have provided necessary equations for the equilibrium. We now prove that the<br />

conditions in (I.7) except for the first one are also sufficient to justify the dealer’s pricing<br />

equations, if one imposes an additional condition: (iii) the numbers β k = ∑ m−k<br />

i=0 ρ i¯γ k+i satisfy<br />

1 > β 1 > · · · > β m > 0. But we have already seen that condition (iii) ensures that the<br />

equilibrium pricing coefficients are well defined <strong>and</strong> constant. More generally, one can use<br />

Lemma I.2 to replace condition (iii) with the condition that the (complex) roots of the polynomial<br />

Q(z) = z m +β 1 z m−1 +· · ·+β m−1 z +β m lie in the open unit disk D = {z ∈ C | |z| < 1}.<br />

Lemma I.2. Let x 1 , . . . , x m ∈ R, <strong>and</strong> consider a sequence x t ∈ R which satisfies the following<br />

recursive equation:<br />

x t + β 1 x t−1 + · · · + β m x t−m = α, t ≥ m + 1. (I.31)<br />

Then the sequence x t converges to ¯x =<br />

α<br />

1+(β 1 +···+β m) , regardless of the initial values x 1, . . . , x m ,<br />

if <strong>and</strong> only if all the (complex) roots of the polynomial Q(z) = z m +β 1 z m−1 +· · ·+β m−1 z +β m<br />

lie in the open unit disk D = {z ∈ C | |z| < 1}. For this, a sufficient condition is that the<br />

coefficients β i satisfy<br />

1 > β 1 > · · · > β m > 0. (I.32)<br />

Furthermore, if α = α t is not constant, then under the same conditions on β i , the difference<br />

x t −<br />

α t<br />

1+(β 1 +···+β m)<br />

converges to zero, regardless of the initial values for x.<br />

Proof. First, note that ¯x is well defined as long as 1+β 1 +· · ·+β m ≠ 0. Indeed, if 1+β 1 +· · ·+<br />

β m = 0 then Q(z) would have z = 1 as a root, which does not lie in D. Now, if we denote by<br />

y t = x t − ¯x, the new sequence y t satisfies the recursive equation y t +β 1 y t−1 +· · ·+β m y t−m = 0.<br />

We now show that y t converges to 0 (<strong>and</strong> ¯x is well defined) if <strong>and</strong> only if all the roots of Q lie<br />

in D. Denote these roots by q 1 , . . . , q m . Then, the difference equation in complex numbers:<br />

y t + β 1 y t−1 + · · · + β m y t−m = 0 has the following general solution:<br />

y t = C 1 q t 1 + · · · C m q t m, t ≥ 1, (I.33)<br />

7


where C 1 , . . . , C m are arbitrary complex constants. 38 But then, y t is convergent for any values<br />

of C i if <strong>and</strong> only if all q i lie in D, or if q i = 1. But in the latter case, 1 + β 1 + · · · + β m = 0,<br />

which makes ¯x nondefined.<br />

The statement that (I.32) implies that all roots of Q lie in D is known as the Eneström–<br />

Kakeya theorem. For completeness, we include the proof here. First, we prove that all roots<br />

of Q must lie in ¯D = {z ∈ C | |z| ≤ 1}.<br />

By contradiction, suppose that there exists z ∗<br />

a root of Q with |z ∗ | > 1. Then, we also have (1 − z ∗ )Q(z ∗ ) = 0 which implies z∗ m+1 =<br />

β m + ∑ m−1<br />

i=0 (β i − β i+1 )z∗ m−i , where β 0 = 1. After taking absolute values, we obtain |z∗<br />

m+1 | ≤<br />

β m + ∑ m−1<br />

i=0 (β i−β i+1 )|z∗<br />

m−i | < β m |z∗ m |+ ∑ m−1<br />

i=0 (β i−β i+1 )|z∗ m | = ( β m + ∑ m−1<br />

i=0 (β i−β i+1 ) ) |z∗ m | =<br />

|z∗ m |. Thus, |z∗<br />

m+1 | < |z∗ m |, which is a contradiction. We have just proved that all the roots of Q<br />

must lie in ¯D. Finally, we show that the roots of any Q(z) = z m + β 1 z m−1 + · · · + β m−1 z + β m<br />

satisfying (I.32) must actually lie in D.<br />

have r m > β 1 r m−1 > . . . > β m−1 r > β m > 0.<br />

Let r < 1 be sufficiently close to 1 so that we<br />

Then, we have seen that the polynomial<br />

Q r (z) = Q(rz) must have all roots in ¯D. Let z ∗ be a root of Q. Then, Q r<br />

( z∗r<br />

)<br />

= Q(z∗ ) = 0,<br />

which implies that z∗ r<br />

complete.<br />

∈ ¯D, or equivalently z ∗ ∈ r ¯D. But r ¯D ⊂ D, <strong>and</strong> the proof is now<br />

I.3 Proof of Proposition 6<br />

Since the forecast error variance equals Σ t = Var(w t − p t−1 ) = E ( (w t − p t−1 ) 2) , we compute<br />

the derivative of Σ t as follows:<br />

Σ ′ t = 1 dt E (2(dw t+1 − dp t )(w t − p t−1 ) + (dw t+1 − dp t ) 2) ,<br />

= −2 Cov(w t, dp t )<br />

dt<br />

+ σw 2 + Var(dp t)<br />

.<br />

dt<br />

(I.34)<br />

The pricing equation (I.23) from the proof of Theorem 3 shows that Cov(w t , dy t ) = λ Var(dy t ) =<br />

λσ 2 y, therefore using dp t = λdy t we get<br />

Var(dp t )<br />

dt<br />

= Cov(w t, dp t )<br />

dt<br />

= λ 2 σ 2 y = σ 2 p. (I.35)<br />

The equation (I.34) now implies that Σ ′ t = σ 2 w − σ 2 p, which finishes the proof.<br />

I.4 Proof of Proposition 7<br />

Compared to the setup of Theorem 3, here there are only speculators with zero lag (l = 0).<br />

Therefore, to finish the proof of this Proposition, all we need to do is to show that the system<br />

in (61) reduces to the system in (66). With the usual notation, the system in (61) translates<br />

38 To obtain real values of y t, one needs to impose the following conditions: (i) if q i is real, then so is C i;<br />

<strong>and</strong> (ii) if q i <strong>and</strong> q j are complex conjugate, then so are C i <strong>and</strong> C j.<br />

8


into (I.7). In the particular case when all speculators are of type l = 0, <strong>and</strong> there are N 0 = N<br />

of them, equation (I.7) becomes: 39<br />

( ) 1<br />

A<br />

N g + g<br />

A i,j<br />

= 1 Λ B, r = Ag, Λ = g′ B, ˜σ 2 y =<br />

min(i,j)<br />

∑<br />

= 1 i=j − r i−k r j−k , B i = 1 − Λ<br />

k=1<br />

˜σ 2 u<br />

1 − g ′ r ,<br />

i∑<br />

r i−k .<br />

We now proceed by observing that the first two equations imply Ag = r =<br />

i = 0, this equation implies r 0 = 1 Λ<br />

By induction, we obtain (i = 0, . . . , m):<br />

N<br />

N+1 . When i = 1, we get r 1 = 1 Λ<br />

r i = 1 Λ<br />

This proves the equation for ρ i in (66). We also get that<br />

k=1<br />

N<br />

N+1<br />

1<br />

Λ<br />

N<br />

N+1 (1−Λ r 0) = 1 Λ<br />

(I.36)<br />

B. When<br />

N<br />

(N+1) 2 .<br />

N<br />

. (I.37)<br />

(N + 1) i+1<br />

B i =<br />

1<br />

(N + 1) i , (I.38)<br />

which proves the equation for B i in (66). Moreover, we compute g ′ r = g ′ B N<br />

N+1<br />

g ′ B = Λ, hence<br />

This implies<br />

˜σ 2 y =<br />

g ′ r =<br />

N<br />

N + 1 .<br />

1<br />

Λ .<br />

But<br />

(I.39)<br />

˜σ 2 u<br />

1 − g ′ r = (N + 1)˜σ2 u, or ˜σ y = √ N + 1 ˜σ u , (I.40)<br />

which proves the equation for ˜σ y in (66). Using the formula (I.37) for r, we use (I.36) to<br />

compute (i, j = 0, . . . , m):<br />

A i,j = 1 i=j − 1 Λ 2 N<br />

N + 2<br />

(N + 1) 2 min(i,j) − 1<br />

(N + 1) i+j , (I.41)<br />

which proves the equation for A i,j in (66). Finally, the two equations, r = Ag <strong>and</strong> Λ = g ′ B,<br />

after rescaling are equivalent to ˜σ 2 yρ = A¯γ <strong>and</strong> ˜σ 2 yλ = B ′¯γ, which finishes the proof of this<br />

Proposition.<br />

I.5 Proof of Theorem 4<br />

We use the notations from the proofs of Theorem 3 <strong>and</strong> Proposition 7. The idea is to study<br />

the behavior of (A, Λ, r, g) around Λ = 1, but without yet imposing the condition Λ = g ′ B.<br />

39 Note that except for the first equation, the other equations in (I.36) are the same as in (I.7). We also<br />

provide a direct derivation of the first equation in the proof of Proposition 8.<br />

9


To do that, define the following numbers:<br />

A 0 i,j = 1 i=j − N<br />

N + 2<br />

r 0 i =<br />

One verifies the formula: 40<br />

(N + 1) 2 min(i,j) − 1<br />

(N + 1) i+j , Λ 0 = 1,<br />

N<br />

(N + 1) i+1 , g0 i = N<br />

N + 1<br />

(m − i + 1)N + 1<br />

(m + 1)N + 1 . (I.42)<br />

A 0 g 0 = r 0 , or, equivalently, g 0 = ( A 0) −1 r 0 . (I.43)<br />

For ε < 1, define the following variables:<br />

A ε i,j = 1 i=j − 1<br />

1 − ε<br />

ri ε 1<br />

= √ 1 − ε<br />

N<br />

N + 2<br />

N<br />

(N + 1) i+1 , gε = (A ε ) −1 r ε ,<br />

(N + 1) 2 min(i,j) − 1<br />

(N + 1) i+j , Λ ε = √ 1 − ε<br />

(I.44)<br />

whenever A ε is invertible. Using (I.43), one sees that the variables defined in (I.42) are the<br />

same as the variables in (I.44) in the particular case when ε = 0. In other words, given a<br />

solution (A, B, Λ, r, g, ˜σ y ) to the system (I.7), if we let<br />

ε = 1 − Λ 2 ,<br />

(I.45)<br />

it must be that (Λ, A, r, g) satisfy<br />

Λ = Λ ε , A = A ε , r = r ε , g = g ε . (I.46)<br />

We now multiply the equation for A = A ε from (I.44) by Λ 2 = 1 − ε, to obtain<br />

Λ 2 A = A 0 − εI,<br />

(I.47)<br />

where I is the identity matrix (I i,j = 1 i=j ). This implies 1 Λ 2 A −1 = ( A 0 − εI ) −1 . Multiplying<br />

this equation to the right with r = 1 Λ r0 , we obtain<br />

g<br />

Λ = ( A 0 − εI ) −1 r 0 .<br />

(I.48)<br />

If we multiply this equation to the left with B ′ , <strong>and</strong> use B ′ g = Λ, we obtain<br />

1 = B ′( A 0 − εI ) −1 r 0 . (I.49)<br />

40 This can be done either directly, or by using the method described below which involves recursive computation<br />

of the inverse matrix, (A 0 ) −1 . Then, one verifies by induction that (A 0 ) −1 r 0 = g 0 .<br />

10


This equation determines ε, or equivalently Λ = √ 1 − ε. We make this equation more explicit<br />

by observing that the inverse matrix ( A 0 − εI ) −1 has the following series expansion:<br />

(<br />

A 0 − εI ) −1 = (A 0 ) −1 + ε (A 0 ) −2 + ε 2 (A 0 ) −3 + · · · . (I.50)<br />

By multiplying this equation to the left by B ′ <strong>and</strong> to the right by r 0 , we get<br />

1 = B ′( A 0 − εI ) −1 r 0 = B ′ (A 0 ) −1 r 0 + ε B ′ (A 0 ) −2 r 0 + ε 2 B ′ (A 0 ) −3 r 0 + · · · . (I.51)<br />

We compute 1 − B ′ g 0 =<br />

1<br />

(m+1)N+1 . Since ( A 0) −1 r 0 = g 0 , we obtain<br />

1<br />

(m + 1)N + 1 = ε B′( A 0) −1 g 0 + ε 2 B ′( A 0) −2 g 0 + ε 3 B ′( A 0) −3 g 0 + · · · . (I.52)<br />

We now determine sufficient conditions for the existence of an equilibrium.<br />

From the<br />

previous discussion, we need the following conditions: (i) ε < 1 (Λ is well defined <strong>and</strong> Λ ><br />

0); (ii) (A 0 − εI) is invertible or equivalently A = A ε is invertible (g is well defined); <strong>and</strong><br />

(iii) the numbers β k from the proof of Theorem 3 satisfy (I.30) for k = 1, . . . , m (which implies<br />

Cov(w t , dy t ) is independent of t). With the current notation, condition (iii) requires that 41<br />

1 > β 1 > · · · > β m > 0, with β k =<br />

m−k<br />

∑<br />

i=0<br />

r i g k+i .<br />

(I.53)<br />

We also introduce the following condition that implies (i) <strong>and</strong> (ii):<br />

Equation (I.52) has a solution ε ∈<br />

(<br />

0 ,<br />

1<br />

N(m+2) 2<br />

8<br />

+ (m+2)2<br />

m+1<br />

)<br />

. (I.54)<br />

Since N(m+2)2<br />

8<br />

+ (m+2)2<br />

m+1<br />

> 1, clearly (I.54) implies ε < 1, which proves (i). The difficult part<br />

is to show that (I.54) also implies that (A 0 − εI) is invertible, which proves (ii). For this,<br />

we need a better underst<strong>and</strong>ing of the inverse matrix ( A 0) −1 . Denote by A<br />

0<br />

(m)<br />

∈ M m+1<br />

the matrix A 0 from (I.42) by making explicit the dependence on m. Since A 0 satisfies A 0 i,j =<br />

1 i=j − ∑ min(i,j)<br />

k=1<br />

r i−k r j−k , it follows that the block ( A 0 11<br />

(m))<br />

which is obtained A<br />

0<br />

(m)<br />

by removing<br />

the last row <strong>and</strong> the last column is the same as A 0 (m−1). We then have<br />

[ ]<br />

A<br />

A 0 0<br />

(m) = (m−1)<br />

a (m)<br />

, (I.55)<br />

a ′ (m)<br />

α (m)<br />

for some m-column vector a (m) , <strong>and</strong> scalar α (m) . Write the inverse matrix H (m) = (A 0 (m) )−1<br />

41 One can check that condition (iii) is implied by the condition g 1 > g 2 > · · · > g m > 0.<br />

11


also in block format:<br />

H (m) =<br />

[ ]<br />

H<br />

11<br />

(m)<br />

h (m)<br />

. (I.56)<br />

h ′ (m)<br />

η (m)<br />

From the theory of block matrices, we have the following formulas:<br />

η (m) =<br />

1<br />

( )<br />

α (m) − a ′ (m) A<br />

0 −1a(m)<br />

,<br />

(m−1)<br />

h (m) = −η (m)<br />

(<br />

A<br />

0<br />

(m−1)<br />

) −1a(m)<br />

,<br />

(I.57)<br />

H 11<br />

(m)<br />

= ( A 0 ) −1<br />

h (m) h ′ (m)<br />

(m−1) + .<br />

η (m)<br />

By induction, one verifies that<br />

η (m) =<br />

h (m) =<br />

(mN + 1)(N + 1)<br />

(m + 1)N + 1 ,<br />

N 2<br />

(m + 1)N + 1<br />

[<br />

] (I.58)<br />

′.<br />

0 , 1 , · · · , m − 1<br />

Using the equations above, one can now prove various useful formulas. As a first result, we<br />

prove by induction that<br />

det ( A 0 (m)<br />

) (m + 1)N + 1<br />

= . (I.59)<br />

(N + 1) m+1<br />

For m = 0, the equality is true, since in this case A 0 (0)<br />

= 1. Suppose it is true for m − 1. From<br />

the theory of block matrices, det ( A 0 ) ( ) (<br />

(m) = det A<br />

0<br />

(m−1)<br />

det α (m) − a ′ ( ) )<br />

(m) A<br />

0 −1a(m)<br />

(m−1)<br />

=<br />

( )<br />

det A 0 (m−1)<br />

η (m)<br />

, which together with the formula for η (m) from (I.58) proves the induction step.<br />

Another useful result is:<br />

H i,j ≥ 0, i, j = 0, . . . , m. (I.60)<br />

Indeed, one uses the recursive formula H(m) 11 = H (m−1) + h (m)h ′ (m)<br />

η<br />

<strong>and</strong> the explicit formulas<br />

(m)<br />

in (I.58) to verify by induction that all entries of H = H (m) are positive. 42<br />

In order to prove that the matrix (A 0 − εI) is invertible, we rewrite equation (I.61):<br />

(<br />

A 0 − εI ) (<br />

)<br />

−1 = H 1 + εH + ε 2 H 2 + ε 3 H 3 + · · · . (I.61)<br />

Thus, if we can show that the right h<strong>and</strong> side is a convergent series (in the space of matrices),<br />

then its limit is a matrix which coincides with the matrix inverse ( A 0 − εI ) −1 .<br />

To prove<br />

convergence, we use the infinity norm, ‖H‖ ∞ , which is the maximum absolute row sum of the<br />

∑<br />

matrix, i.e., H =<br />

m<br />

j=0 |H i,j|, . Thus, if we can show that ‖εH‖ ∞ < 1, this proves<br />

condition (ii).<br />

max<br />

i=0,...,m<br />

42 The inequality is strict except that H i,0 = H 0,i = 0 for i > 0.<br />

12


We now search for an upper bound for ‖H‖ ∞ . For instance, we show that ∥ ∥ H ∥∞ ≤<br />

N(m+2)<br />

( 2 1<br />

4 1 +<br />

1 ON<br />

N )) . For this, define ¯h (m) the (m + 1)-column vector given by (¯h(m) )i = (N +<br />

1) ( (m + 1)N + 1 ) ∑ m<br />

( )<br />

j=0 H(m) . This is proved by induction to be a polynomial in N of<br />

i,j<br />

degree 3. Denote by C (m) the vector of coefficients of N 3 in ¯h (m) . Note that ¯h (m) =<br />

max<br />

i=0,...,m<br />

N 2 (<br />

1<br />

(m+1)‖H‖ ∞ 1+ON<br />

N )) . At the same time, we have max ¯h (m) = N 3 max C (m), which<br />

i=0,...,m<br />

i=0,...,m<br />

implies ∥ ∥ H ∥∞ 1<br />

= max<br />

N(m+2) 2 (m+1)(m+2) C ( 2<br />

(m) 1 +<br />

1 ON<br />

i=0,...,m<br />

N )) . Now one computes C (0) = 0,<br />

<strong>and</strong> for m > 1 one uses the recursive formulas above for H to get a recursive formula for<br />

( ) ( )<br />

( )<br />

C(m)<br />

C. More precisely,<br />

m+1<br />

= C(m−1)<br />

i<br />

m<br />

+ i i<br />

2 for i = 0, . . . , m − 1, <strong>and</strong> C(m)<br />

m+1<br />

= m m<br />

2 . By<br />

(<br />

induction then, one shows that max C(m)<br />

i<br />

)i ≤ (m+1)2 (m+2)<br />

4<br />

, which implies the upper bound<br />

stated above for ‖H‖ ∞ . By similar methods, one verifies a sharper estimate:<br />

Note that condition (I.54) implies ε N(m+2) 2 <<br />

1<br />

∥ H ∥∥∥∞<br />

∥N(m + 2) 2 ≤ 1 8 + 1<br />

(m + 1)N . (I.62)<br />

1<br />

8 + 1<br />

(m+1)N<br />

, which together with (I.62) implies<br />

ε ‖H‖ ∞ < 1. (I.63)<br />

This proves that the series 1 + εH + ε 2 H 2 + ε 3 H 3 + · · · is convergent, <strong>and</strong> that the limit<br />

coincides with ( A 0 − εI ) −1 .<br />

Next, we analyze how well (Λ, r, g) approximate (Λ 0 , r 0 , g 0 ). Recall that<br />

γ = ¯γ N = g˜σ y<br />

N , ρ = r˜σ y<br />

, λ = Λ˜σ y<br />

, (I.64)<br />

where from (I.40),<br />

˜σ y = √ N + 1 ˜σ u = √ N + 1 σ u<br />

σ w<br />

.<br />

Note that in the statement of the Theorem we have defined (i = 0, . . . , m):<br />

(I.65)<br />

γ 0 i = ˜σ y<br />

N + 1<br />

m − i + 1<br />

m + 1 , ρ0 i = 1˜σ y<br />

N<br />

(N + 1) i+1 , λ0 = 1˜σ y<br />

. (I.66)<br />

(<br />

1<br />

Condition (I.54) implies ε < , which shows that ε = O 1<br />

)<br />

N(m+2) 2 + (m+2)2<br />

N N . Also, since<br />

8 m+1<br />

Λ = √ (<br />

1 − ε, it follows that Λ = 1 − O 1<br />

)<br />

N N . Thus, we obtain<br />

ε = O N<br />

( 1<br />

N<br />

)<br />

, Λ =<br />

√<br />

1 − ε = 1 − ON<br />

( 1<br />

N<br />

)<br />

. (I.67)<br />

Now, from (I.64) <strong>and</strong> (I.66), we obtain λ λ 0 = Λ = 1 − O N<br />

( 1<br />

N<br />

)<br />

. This proves the approximate<br />

equation for λ in (68). From (I.64) <strong>and</strong> (I.66), we also compute ρ i<br />

= r ρ 0 i<br />

, since r 0<br />

i ri<br />

0 i = N<br />

.<br />

(N+1) i+1<br />

But r = 1 Λ r0 , which implies ρ i<br />

= 1 ρ 0 Λ = 1 + O 1<br />

)<br />

N(<br />

N . This proves the approximate equation<br />

i<br />

13


for ρ i in (68). Finally, from (I.64) <strong>and</strong> (I.66), we get γ i<br />

γi<br />

0<br />

We now show that g i<br />

gi<br />

0<br />

γ 0 i<br />

=<br />

N<br />

N+1<br />

g i<br />

m−i+1<br />

m+1<br />

= g i<br />

g 0 i<br />

(<br />

1 + ON ( 1 N )) .<br />

= O N (1), which proves the approximate equation for γ i in (68). Since<br />

= O N (1), it is enough to show that g i − g 0 i = O N (1), or from (I.67) it is enough to show<br />

g i<br />

Λ − g0 i = O N(1). If we combine equations (I.48) <strong>and</strong> (I.61), <strong>and</strong> use ( A 0) −1 r 0 = g 0 , we get<br />

g<br />

(<br />

)<br />

Λ = 1 + εH + ε 2 H 2 + ε 3 H 3 + · · · g 0 .<br />

(I.68)<br />

Therefore, we get g Λ −g0 = ( εH +ε 2 H 2 +ε 3 H 3 +· · ·) g 0 . But (I.63) implies that the convergent<br />

series of matrices is of the order O N (1), hence it remains of order O N (1) when multiplied with<br />

g 0 = O N (1).<br />

I.6 Proof of Proposition 8<br />

Following the proof of Theorem 3, consider a speculator who must choose the weights γ i,t on<br />

d t w t−i . He assumes that all the other speculators use γi ∗ , hence with an aggregate weight of<br />

(N − 1)γi ∗ on d t w t−i . Then, equation (I.10) for the speculator’s normalized expected profit at<br />

t = 0 becomes<br />

˜π 0 = π 0<br />

σw<br />

2 ⎛<br />

∫ 1<br />

= E ⎝<br />

=<br />

= 1<br />

σ 2 w<br />

0<br />

(<br />

w t − λ<br />

m∑<br />

B j γ j,t − λ<br />

j=0<br />

(∫ 1<br />

)<br />

E (w t − p t−1 − λdy t )dx t<br />

0<br />

m∑ (<br />

γi,t + (N − 1)γi ∗ ) ) ∑<br />

m<br />

dt w t−i<br />

i=0<br />

m∑<br />

i,j=0<br />

(<br />

γi,t + (N − 1)γi<br />

∗ )<br />

Ai,j γ j,t .<br />

The first order condition with respect to γ k,t , for k = 0, . . . , m, is:<br />

B k − λ<br />

j=0<br />

γ j,t d t w t−j<br />

⎞<br />

⎠<br />

(I.69)<br />

m∑ (<br />

2γi,t + (N − 1)γi<br />

∗ )<br />

Ai,k = 0. (I.70)<br />

i=0<br />

Since this equation is true for all speculators, we obtain that all γ i,t are equal <strong>and</strong> independent<br />

on t, i.e., γ i = γi ∗ = ¯γ N<br />

. Using matrix notation, we have B = λ(N + 1)Aγ, hence B =<br />

λ N+1<br />

N A¯γ.43 N<br />

B<br />

Thus,<br />

N+1B = λA¯γ, which implies B − λA¯γ =<br />

N+1<br />

. Thus, in equilibrium the<br />

normalized expected profit is equal to<br />

˜π 0 =<br />

(<br />

)<br />

m∑<br />

m∑<br />

B j − λ A j,i¯γ i γ j = ∑ B j<br />

N + 1 γ j.<br />

j=0<br />

i=0<br />

j=0<br />

(I.71)<br />

43 Since Λ = λ ˜σ y <strong>and</strong> g = ¯γ˜σ y<br />

, we obtain 1 B = N+1 Ag, which provides a direct proof of the first equation<br />

Λ N<br />

in (I.36).<br />

14


From equation (I.38), B j = 1<br />

(N+1) j . We compute<br />

˜π 0 =<br />

m∑<br />

˜π 0,j =<br />

j=0<br />

m∑<br />

j=0<br />

γ j<br />

, (I.72)<br />

(N + 1) j+1<br />

which proves (69). The ratio of two consecutive components is<br />

˜π 0,j+1<br />

˜π 0,j<br />

= γ j+1<br />

γ j<br />

1<br />

N + 1 = g j+1<br />

g j<br />

1<br />

N + 1 .<br />

(I.73)<br />

But in the proof of Theorem 4 we have shown that g j = O N (1). Thus, ˜π 0,j+1<br />

˜π 0,j<br />

= O N<br />

( 1<br />

N<br />

)<br />

, which<br />

finishes the proof of the Proposition.<br />

I.7 Proof of Propositions 9<br />

With the notations from Section I.1, σ 2 w − σ 2 p = σ 2 w(1 − λ 2˜σ 2 y) = σ 2 w(1 − Λ 2 ). But in the proof<br />

of Theorem 4, we have seen that 1 − Λ 2 = ε, <strong>and</strong> this is strictly positive according to the<br />

condition (I.54). Moreover, from (I.67), we have ε = O N<br />

( 1<br />

N<br />

)<br />

, which finishes the first part of<br />

the Proposition.<br />

For the second part, we need only to prove the equation SR k =<br />

From the definition of SR k , we obtain<br />

N<br />

(N+1) k+1 when k = 0, . . . , m.<br />

SR k = λ Cov(dw t−k, dy t )<br />

σ 2 wdt<br />

= λρ k Var(dy t )<br />

σ 2 wdt<br />

= λρ k˜σ 2 y =<br />

N<br />

, (I.74)<br />

(N + 1) k+1<br />

where the second equality comes from (I.23), <strong>and</strong> the last equation comes from (66).<br />

15


J<br />

Discrete Time<br />

In this section, we develop a discrete time version of our baseline model from Section 2. The<br />

goal is twofold. First, we want to show, in a very general framework, that discrete-time<br />

optimal linear trading strategies must be indeed orthogonal to the public information, as we<br />

have assumed in (11). This is a plausible assumption in continuous time, but we prove it as a<br />

result in discrete time. Second, we want to underst<strong>and</strong> the mechanism by which the discrete<br />

time model approaches the continuous model in the limit, <strong>and</strong> for instance we want to justify<br />

why speculators take the covariance structure of signals as given (constant) when choosing<br />

their optimal strategies.<br />

J.1 Model<br />

Trading occurs at intervals of length ∆t apart, at times t 1 = ∆t, t 2 = 2∆t, . . ., t T = T ∆t. To<br />

simplify notation, we refer to these times as 1, 2, . . . , T . The liquidation value of the asset is<br />

v T =<br />

T∑<br />

∆v t , with ∆v t = v t − v t−1 = σ v ∆Bt v , (J.1)<br />

t=1<br />

where Bt<br />

v is a Brownian motion. The risk-free rate is assumed to be zero.<br />

There are three types of market participants: (a) N ≥ 1 risk neutral speculators, who<br />

observe the flow of information at different speeds, as described below; (b) noise traders; <strong>and</strong><br />

(c) one competitive risk neutral dealer, who sets the price at which trading takes place.<br />

Information. At t = 0, there is no information asymmetry between the speculators <strong>and</strong><br />

the dealer. Subsequently, each speculator receives the following flow of signals:<br />

∆s t = ∆v t + ∆η t , with ∆η t = σ η dB η t , (J.2)<br />

where t = 1, . . . , T <strong>and</strong> B η t is a Brownian motion independent from all other variables. Define<br />

the speculators’ forecast by<br />

Its increment is ∆w t = a∆s t , where a =<br />

w t = E(v T<br />

∣ ∣ {s τ } τ≤t ). (J.3)<br />

σ2 v<br />

σ 2 v +σ2 η<br />

Speculators obtain their signal with a lag l = 0, 1, . . ..<br />

is the signal precision parameter. Also, w 0 = 0.<br />

A l-speculator is a trader who<br />

at t = l + 1, . . . , T observes the lagged signal ∆s t−l , from l periods before.<br />

l = 0, 1, 2, . . ., denote by N l the number of l-speculators.<br />

For each lag<br />

Trading. At each t ∈ (0, T ], denote by ∆x i t the market order submitted by speculator<br />

i = 1, . . . , N at t, <strong>and</strong> by ∆u t the market order submitted by the noise traders, which is of the<br />

form ∆u t = σ u ∆B u t , where B u t<br />

is a Brownian motion independent from all other variables.<br />

16


Then, the aggregate order flow executed by the dealer at t is<br />

∆y t =<br />

N∑<br />

∆x i t + ∆u t .<br />

i=1<br />

(J.4)<br />

Because the dealer is risk neutral <strong>and</strong> competitive, she executes the order flow at a price equal<br />

to her expectation of the liquidation value conditional on her information. Let I t = {y τ } τ


J.2 General Case<br />

We first prove a lemma that helps compute the expected profit of a l-speculator. Define the<br />

following numbers (t = 1, . . . , T , i, j = 1, . . . , t):<br />

A i,j,t = Cov t<br />

(<br />

∆wi , ∆w j<br />

)<br />

= Cov<br />

(<br />

∆t w i , ∆ t w j<br />

)<br />

. (J.11)<br />

This is called the fresh covariance matrix. For a given t, it is a t × t-matrix. We write it in<br />

block format, with blocks above <strong>and</strong> below t − l:<br />

[ ]<br />

A<br />

l<br />

A = 11,t A l 21,t<br />

. (J.12)<br />

A l 12,t A l 22,t<br />

Thus, note that A l 11,t is (t − l) × (t − l) <strong>and</strong> Al 12,t<br />

is l × (t − l).<br />

Lemma J.3. Consider a l-speculator, <strong>and</strong> let k = t − l + 1, . . . , t. Then,<br />

E l ( ) ∑t−l<br />

t ∆t w k = c l k,j,t ∆ tw j ,<br />

j=1<br />

(J.13)<br />

where the constants c l k,j,t ∈ I t form a l × (t − l)-matrix c l t which satisfies<br />

c l t = A l 12,t(<br />

A<br />

l<br />

11,t<br />

) −1, (J.14)<br />

<strong>and</strong> A is the fresh covariance matrix.<br />

Proof. Since J l<br />

t = I t ∪ ∆w 1 ∪ · · · ∪ ∆w t−l , we write<br />

E l ( ) ∑t−l<br />

t ∆t w k = c l k,j,t ∆ tw j + h, with h, c l k,j,t ∈ I t. (J.15)<br />

j=1<br />

Since E t<br />

(<br />

∆t w j<br />

)<br />

= 0, if we take Et (·) on both sides, by the law of iterated expectations we<br />

obtain 0 = h.<br />

We now write:<br />

∆ t w k =<br />

∑t−l<br />

c l k,i,t ∆ tw i + ε k ,<br />

i=1<br />

(J.16)<br />

with ε k orthogonal to Jt l . Taking covariance with ∆ t w j (j = 1, . . . , t − l) on both sides of the<br />

equation above, we get<br />

∑t−l<br />

A k,j,t = c l k,i,t A i,j,t.<br />

(J.17)<br />

In block format this equation becomes c l t = A l 12,t(<br />

A<br />

l<br />

11,t<br />

) −1, which completes the proof.<br />

i=1<br />

18


The next result reduces the equilibrium of the discrete time model to a set of equations<br />

that the optimal weights <strong>and</strong> the pricing coefficients must satisfy. For simplicity, we only<br />

consider the case when there is no information processing cost, <strong>and</strong> therefore speculators can<br />

trade on all their signals.<br />

Theorem J.1. Consider the discrete time model described above with T trading rounds. Suppose<br />

speculators can trade on all their signals. Then, there exists a linear equilibrium of the<br />

form<br />

∆x l t = γ1,t∆ l t w 1 + · · · + γt−l,t l ∆ tw t−l ,<br />

∆ t w j = ∆w j − z j,t ,<br />

z j,t+1 = z j,t + ρ j,t ∆y t ,<br />

∆p t = λ t ∆y t .<br />

(J.18)<br />

if there exist numbers λ t , ρ i,t , γ l k,t , A i,j,t, with t = 1, . . . , T , i, j = 1, . . . , t − 1, k = 1, . . . , t − l,<br />

such that the variables satisfy a system of equations described below (in the proof).<br />

Proof. We work backwards, starting from t = T .<br />

Speculators’ Trading Strategies<br />

Consider the l-speculator’s decision in the last trading round, at t = T . He takes as given the<br />

dealer’s linear pricing rule:<br />

∆p T = λ T ∆y T , (J.19)<br />

<strong>and</strong> the other speculators’ trading strategies:<br />

∆x k t = γ k 1,t∆ t w 1 + · · · + γ k t−k,t ∆ tw t−k , k = 1, . . . , T − 1. (J.20)<br />

where ∆ T w j = ∆w j − z j,T , <strong>and</strong> z j,T = E T (∆w j ). Thus, if he submits ∆x at T , he assumes<br />

the aggregate order flow at T to be of the form:<br />

∆y T = ∆x + ∆u T +<br />

g l j,T =<br />

T −j<br />

∑<br />

k=0<br />

T∑<br />

gj,T l ∆ T w j ,<br />

j=1<br />

with<br />

N −l<br />

k γk j,T , <strong>and</strong> N −l<br />

k<br />

= N k − 1 k=l .<br />

(J.21)<br />

Then, he computes the expected profit as follows:<br />

π T<br />

= E l T<br />

( (<br />

w T − p T −1 − λ T<br />

(<br />

∆x + ∆u T +<br />

T∑ ) ) )<br />

gk,T l ∆ T w k ∆x . (J.22)<br />

k=1<br />

19


Since p T −1 = E T (w T ), z k,T = E T (∆w k ), <strong>and</strong> w T = ∑ T<br />

k=1 ∆w k (recall w 0 = 0), we have<br />

w T − p T −1 =<br />

T∑<br />

∆ T w k .<br />

k=1<br />

(J.23)<br />

From Lemma J.3,<br />

E l T<br />

T<br />

( ) ∑−l<br />

wT − p T −1 = ∆ T w k +<br />

=<br />

k=1<br />

T∑<br />

−l<br />

k=1<br />

∆ T w k<br />

⎛<br />

⎝1 +<br />

T∑<br />

k=T −l+1 j=1<br />

T∑<br />

j=T −l+1<br />

T∑<br />

−l<br />

c l k,j,T ∆ T w j<br />

c l j,k,T<br />

⎞<br />

⎠ .<br />

(J.24)<br />

We compute also<br />

E l T<br />

(<br />

∑ T<br />

)<br />

gk,T l ∆ T w k<br />

k=1<br />

=<br />

=<br />

T∑<br />

−l<br />

gk,T l ∆ T w k +<br />

k=1<br />

T∑<br />

−l<br />

k=1<br />

∆ T w k<br />

⎛<br />

⎝g l k,T +<br />

T∑<br />

gk,T<br />

l<br />

k=T −l+1 j=1<br />

T ∑<br />

j=T −l+1<br />

T∑<br />

−l<br />

c l k,j,T ∆ T w j<br />

⎞<br />

gj,T l c l ⎠<br />

j,k,T .<br />

(J.25)<br />

Putting together the formulas above, we get<br />

E l T<br />

∑<br />

(w T ) T<br />

T −p T −1 − λ T gk,T l ∆ ∑−l<br />

T w k = Ck,T l ∆ T w k ,<br />

k=1<br />

C l k,T = 1 − λ T g l k,T + T ∑<br />

j=T −l+1<br />

k=1<br />

(J.26)<br />

c l (<br />

j,k,T 1 − λT gj,T<br />

l )<br />

.<br />

From (J.22) <strong>and</strong> (J.26), we compute<br />

π T =<br />

( T∑ −l<br />

)<br />

Ck,T l ∆ T w k − λ T ∆x ∆x.<br />

k=1<br />

(J.27)<br />

The first order condition implies<br />

∆x l T = 1<br />

2λ T<br />

T −l<br />

∑<br />

Ck,T l ∆ T w k ,<br />

k=1<br />

(J.28)<br />

20


from which we obtain the equation (k = 1, . . . , T − l):<br />

γ l k,T<br />

= Cl k,T<br />

2λ T<br />

= 1 − λ T gk,T l + ∑ T<br />

(<br />

j=T −l+1 cl j,k,T 1 − λT gj,T<br />

l )<br />

,<br />

2λ T<br />

with g l j,T =<br />

T −j<br />

∑<br />

k=0<br />

N −l<br />

k γk j,T .<br />

(J.29)<br />

The second order condition for a maximum is simply<br />

λ T > 0. (J.30)<br />

We now compute the value function at T of the l-speculator. This is the maximum expected<br />

profit, which corresponds to ∆x = ∆x l T :<br />

∑<br />

k=1<br />

(<br />

VT l = 1 T −l<br />

) 2<br />

Ck,T l 4λ ∆ T w k =<br />

T<br />

T∑<br />

−l<br />

i,j=1<br />

with α l i,j,T −1 = 1<br />

4λ T<br />

C l i,T C l j,T .<br />

α l i,j,T −1 ∆ T w i ∆ T w j ,<br />

(J.31)<br />

Now, we indicate how to find the equations that arise from the l-speculator’s optimization<br />

at t < T . We guess a quadratic value function at t + 1 of the form<br />

t+1−l<br />

∑<br />

Vt+1 l = α 0,t + αi,j,t l ∆ t+1 w i ∆ t+1 w j .<br />

i,j=1<br />

(J.32)<br />

The l-speculator assumes that ∆y t is of the form<br />

t∑<br />

∆y t = ∆x + ∆u t + gj,t∆ l t w j , with gj,t l =<br />

j=1<br />

t−j<br />

∑<br />

k=0<br />

N −l<br />

k γk j,t. (J.33)<br />

Using the same method as above, we get the following formula:<br />

E l t( (∆t<br />

w i − ρ i,t ∆y t<br />

) (<br />

∆t w j − ρ j,t ∆y t<br />

) ) = ρ i,t ρ j,t ∆x 2 + L i,j,t ∆x + Q i,j,t , (J.34)<br />

where as functions of the fresh signals ∆ t w k (k = 1, . . . , t − l), L i,j,t is linear <strong>and</strong> Q i,j,t is<br />

quadratic. Then, the value function at t satisfies the Bellman equation:<br />

V l<br />

t<br />

( (wt )<br />

= max<br />

∆x<br />

El t − p t−1 − λ t ∆y t ∆x + V<br />

l<br />

)<br />

= max<br />

∆x<br />

( ∑t−l<br />

Ck,t l ∆ tw k − λ t ∆x<br />

k=1<br />

t+1<br />

)<br />

t+1−l<br />

∑<br />

∆x + α 0,t +<br />

i,j=1<br />

α l i,j,t<br />

)<br />

(ρ i,t ρ j,t ∆x 2 + L i,j,t ∆x + Q i,j,t .<br />

(J.35)<br />

21


The first order condition for ∆x implies<br />

∆x l t =<br />

∑ T −l<br />

k=1 Cl k,T ∆ T w k + ∑ t+1−l<br />

i,j=1<br />

αl i,j,t L i,j,t<br />

2 ( λ t − ∑ t+1−l<br />

i,j=1 αl i,j,t ρ ) , (J.36)<br />

i,tρ j,t<br />

<strong>and</strong> the second order condition for a maximum is<br />

t+1−l<br />

∑<br />

λ t − αi,j,tρ l i,t ρ j,t > 0. (J.37)<br />

i,j=1<br />

Now we identify the coefficients of ∆x l t to obtain the equations for the optimal weights γ l k,t .<br />

By substituting ∆x = ∆x l t in the Bellman equation, we get a quadratic formula in the fresh<br />

signals, from which we obtain equations for αi,j,t l . Furthermore, using similar formulas, one<br />

computes the recursive formula for the fresh covariance matrix A. We give more explicit<br />

formulas in Section J.2.<br />

Dealer’s Pricing Rules<br />

The dealer assumes that the aggregate order flow is of the form:<br />

t∑<br />

∆y t = ∆u t + ¯γ j,t ∆ t w j , with ¯γ j,t =<br />

j=1<br />

t−j<br />

∑<br />

N k γj,t.<br />

k<br />

k=0<br />

(J.38)<br />

Because each speculator only trades on the unpredictable part of his signal, ∆y t are orthogonal<br />

to each other. Thus, the dealer computes<br />

z k,t<br />

= E ( ) ∑t−1<br />

∆w k | ∆y 1 , . . . , ∆y t−1 = ρ k,j ∆y j , k = 0, . . . , m,<br />

∆p t = λ t ∆y t ,<br />

j=k<br />

(J.39)<br />

where the coefficients ρ k,t <strong>and</strong> λ t are given by:<br />

ρ k,t = Cov(∆w k, ∆y t )<br />

,<br />

Var(∆y t )<br />

λ t = Cov t(v 1 , ∆y t )<br />

Var t (∆y t )<br />

= Cov( )<br />

w t − p t−1 , ∆y t<br />

.<br />

Var(∆y t )<br />

(J.40)<br />

Since the order flow is orthogonal, we can write<br />

ρ k,t = Cov(∆ tw k , ∆y t )<br />

=<br />

Var(∆y t )<br />

t∑ Cov ( )<br />

∆ t w k , ∆y t<br />

λ t =<br />

Var(∆y t )<br />

k=1<br />

∑ t<br />

j=1 A k,j,t¯γ j,t<br />

∑ t<br />

i,j=1 A i,j,t¯γ i,t¯γ j,t<br />

,<br />

=<br />

∑ t<br />

k,j=1 A k,j,t¯γ j,t<br />

∑ t<br />

k,j=1 A k,j,t¯γ k,t¯γ j,t<br />

.<br />

(J.41)<br />

22


These equations complete the system of equations that must be satisfied in equilibrium. Note<br />

that here λ t = ∑ t<br />

k=1 ρ k,t. This is not true if instead of taking m = T we take m < T .<br />

J.3 Fast <strong>and</strong> Slow <strong>Traders</strong> in Discrete Time (m = 1)<br />

Consider a model with fast <strong>and</strong> slow traders, as in Section 3, but in discrete time. As in the<br />

continuous time case, we simplify notation, <strong>and</strong> normalize covariances <strong>and</strong> variances using the<br />

tilde notation. For instance,<br />

˜Cov(∆w t , ∆w t ) = Cov(∆w t, ∆w t )<br />

σ 2 w∆t<br />

= 1. (J.42)<br />

Speculators’ Optimal Strategies<br />

In order to find the optimal strategy of the speculator, we first prove a lemma that helps<br />

compute the speculator’s expected profit.<br />

Lemma J.4. The FT computes<br />

∣ ∣∣<br />

E<br />

(w t − p t−1 It , ∆w t , ˜∆w<br />

)<br />

t−1<br />

= ∆w t + C t˜∆wt−1 ; (J.43)<br />

the coefficient C t is given by<br />

C t<br />

= B t<br />

A t<br />

,<br />

where B t , D t , A t satisfy the following recursive formulas:<br />

(J.44)<br />

B t+1 = 1 − N F λ t γ ∗ t − N F ρ t γ ∗ t − ρ t (N F µ ∗ t + N S ν ∗ t )B t + λ t ρ t D t ,<br />

D t+1 = (N F γ ∗ t+1) 2 + (N F µ ∗ t+1 + N S ν ∗ t+1) 2 A t+1 + ˜σ 2 u,<br />

(J.45)<br />

A t+1 = 1 − 2N F ρ t γ ∗ t + ρ 2 t D t ,<br />

<strong>and</strong> γ ∗ , µ ∗ , ν ∗ are the equilibrium values of the corresponding coefficients.<br />

The ST computes<br />

with C t as above.<br />

∣ ∣∣<br />

E<br />

(w t − p t−1 It , ˜∆w<br />

)<br />

t−1<br />

= C t˜∆wt−1 , (J.46)<br />

Proof. Since all the variables involved are jointly multivariate normal, the conditional expectation<br />

in (J.43) is of the form<br />

∣<br />

E<br />

(w t − p t−1 I t , ∆w t , ˜∆w<br />

)<br />

t−1<br />

= c 1,t ∆w t + c 2,t˜∆wt−1 + c 0,t , with c i,t ∈ I t . (J.47)<br />

23


Because w t − p t−1 , ∆w t <strong>and</strong> ˜∆w t−1 are orthogonal to I t , we have<br />

c 0,t = 0, c 1,t = Cov( )<br />

w t − p t−1 , ∆w t<br />

Var ( ) , c 2,t = Cov( w t − p t−1 , ˜∆w<br />

)<br />

t−1<br />

∆w t Var (˜∆wt−1 ) . (J.48)<br />

Since p t−1 ∈ I t , we have c 1,t = 1. Denote by<br />

B t = ˜Cov ( w t − p t−1 , ˜∆w t−1<br />

)<br />

, Dt = Ṽar( ∆y t<br />

)<br />

, At = Ṽar(˜∆wt−1<br />

)<br />

. (J.49)<br />

We now give a recursive formula for<br />

C t = c 2,t = B t<br />

A t<br />

.<br />

(J.50)<br />

The aggregate order flow has the form:<br />

∆y t = N F γ ∗ t ∆w t + (N F µ ∗ t + N S ν ∗ t )˜∆w t−1 + ∆u t . (J.51)<br />

Therefore, we compute (˜σ 2 u = σ2 u<br />

σ 2 w<br />

):<br />

A t+1<br />

= Ṽar( ∆w t − ρ t ∆y t<br />

)<br />

= 1 − 2NF ρ t γ t + ρ 2 t D t<br />

D t+1 = (N F γt+1) ∗ 2 + (N F µ ∗ t+1 + N S νt+1) ∗ 2 A t+1 + ˜σ u,<br />

2<br />

B t+1 = ˜Cov ( )<br />

w t − p t−1 − λ t ∆y t , ∆w t − ρ t ∆y t<br />

= 1 − ρ t N F γ ∗ t − ρ t (N F µ ∗ t + N S ν ∗ t )B t − λ t N F γ ∗ t + λ t ρ t D t .<br />

(J.52)<br />

These proves the desired formulas.<br />

For the ST, we have the same computation as for the FT.<br />

This lemma raises a subtle issue. Even though the speculator computes the weight C t<br />

that he associates to the signal, it is as if the dealer performed the computation.<br />

Indeed,<br />

the coefficient C t is computed using only the public information I t , with no extra knowledge<br />

about the actual value of the signals. Then, we can imagine that the speculator announces<br />

the C t at the beginning of the trading game, as part of his strategy. This is similar to the way<br />

in which the dealer chooses the pricing coefficient ρ t . 44<br />

We now proceed with computing the FT’s value function. Denote by E F t (X) = E(X|I t , ∆w t , ˜∆w t−1 ),<br />

the expectation from the point of view of the FT at t. Then, the FT’s value function at t<br />

44 In fact, ρ t does not affect any price set by the dealer. Instead, it is just a coefficient that is used in<br />

computing the dealer’s expectation of ∆w t, given the order flow at t. And we have already assumed that ρ t<br />

is announced at the beginning of the game, <strong>and</strong> the speculators take that into account when deciding on their<br />

strategies.<br />

24


satisfies the Bellman equation:<br />

V F<br />

t<br />

(<br />

= max E F (<br />

t (vT − p t )∆x ) )<br />

+ Vt+1<br />

F . (J.53)<br />

∆x<br />

As in the general case, we conjecture a value function for the FT that is quadratic in the<br />

current signals:<br />

V F<br />

t = α 0 t−1 + α t−1<br />

(˜∆wt−1<br />

) 2 + 2α<br />

′<br />

t−1<br />

(˜∆wt−1<br />

)(<br />

∆wt<br />

)<br />

+ α<br />

′′<br />

t−1(<br />

∆wt<br />

) 2. (J.54)<br />

Then, the Bellman equation becomes<br />

V F<br />

t<br />

= max<br />

∆x<br />

EF t<br />

( (wt<br />

− p t−1 − λ t ∆y t<br />

)<br />

∆x<br />

+ α 0 t + α t<br />

(<br />

∆wt − ρ t ∆y t<br />

) 2 + 2α<br />

′<br />

t<br />

(<br />

∆wt − ρ t ∆y t<br />

)(<br />

∆wt+1<br />

)<br />

+ α<br />

′′<br />

t<br />

(<br />

∆wt+1<br />

) 2<br />

)<br />

,<br />

(J.55)<br />

where ∆y t is assumed by the FT to be of the form<br />

∆y t = ∆x + (N F − 1)γ ∗ t ∆w t + ( (N F − 1)µ ∗ t + N S ν ∗ t<br />

)˜∆wt−1 + ∆u t . (J.56)<br />

From equation (J.43), we compute E F ( )<br />

t wt −p t−1 = ∆wt +C t˜∆wt−1 , with C t satisfying certain<br />

equations described in Lemma J.4. Therefore,<br />

V F<br />

t<br />

= max<br />

∆x<br />

EF t<br />

( (∆wt<br />

+ C t˜∆wt−1 − λ t ∆y t<br />

)<br />

∆x+<br />

α 0 t + α t<br />

(<br />

∆wt − ρ t ∆y t<br />

) 2 + α<br />

′′<br />

t σ 2 w∆t<br />

)<br />

,<br />

(J.57)<br />

The terms can be rearranged to write<br />

V F<br />

t<br />

(<br />

= max W1 − λ t ∆x ) (<br />

+ α t W2 − ρ t ∆x) 2 + Z, (J.58)<br />

∆x<br />

where<br />

W 1 = W 11 ∆w t + W 12˜∆wt−1 , with<br />

W 11 = 1 − (N F − 1)λ t γ ∗ t , W 12 = C t − λ t<br />

(<br />

(NF − 1)µ ∗ t + N S ν ∗ t<br />

)<br />

,<br />

W 2 = W 21 ∆w t + W 22˜∆wt−1 , with<br />

W 21 = 1 − (N F − 1)ρ t γ ∗ t , W 22 = −ρ t<br />

(<br />

(NF − 1)µ ∗ t + N S ν ∗ t<br />

)<br />

,<br />

Z = α 0 t + α ′′<br />

t σ 2 w∆t + α t σ 2 u∆t.<br />

(J.59)<br />

The first order condition with respect to ∆x is<br />

W 1 − 2λ t ∆x − 2α t ρ t (W 2 − ρ t ∆x) = 0.<br />

(J.60)<br />

25


If we denote by<br />

˜λ t = λ t − α t ρ 2 t , (J.61)<br />

the first order condition implies<br />

∆x = W 1 − 2α t ρ t W 2<br />

2˜λ t<br />

= W 11 − 2α t ρ t W 21<br />

2˜λ t<br />

∆w t + W 21 − 2α t ρ t W 22<br />

2˜λ t<br />

˜∆wt−1 . (J.62)<br />

The second order condition for a maximum is<br />

˜λ t > 0. (J.63)<br />

By identifying the coefficients of Vt<br />

F , we obtain<br />

αt−1 0 = Z,<br />

( ) 2<br />

W12 − 2α t ρ t W 22<br />

α t−1 =<br />

α ′ t−1 =<br />

α ′′<br />

t−1 =<br />

+ α t W22,<br />

2<br />

4˜λ<br />

( t<br />

)( )<br />

W11 − 2α t ρ t W 21 W12 − 2α t ρ t W 22<br />

+ α t W 11 W 21 ,<br />

4˜λ t<br />

( ) 2<br />

W11 − 2α t ρ t W 21<br />

4˜λ t<br />

+ α t W 2 11.<br />

(J.64)<br />

Note that only α t is involved in a recursive equation, while all the other coefficients can be<br />

computed using α t (<strong>and</strong> the equilibrium coefficients).<br />

explicitly:<br />

α t−1 =<br />

(<br />

C t − ( λ t − 2α t ρ 2 )(<br />

t (NF − 1)µ ∗ t + N S νt<br />

∗ ) ) 2<br />

4˜λ t<br />

We write the equation for α t more<br />

+ α t ρ 2 (<br />

t (NF − 1)µ ∗ t + N S νt<br />

∗ ) 2. (J.65)<br />

From (J.62), we obtain the equations for the coefficients γ t <strong>and</strong> µ t for the FT:<br />

γ t = 1 − 2α tρ t − ( λ t − 2α t ρ 2 )<br />

t (NF − 1)γt<br />

∗ ,<br />

2˜λ t<br />

µ t = C t − ( λ t − 2α t ρ 2 )(<br />

t (NF − 1)µ ∗ t + N S νt ∗ )<br />

.<br />

2˜λ t<br />

(J.66)<br />

We now proceed in a similar way to compute the ST’s value function. Denote by E S t (X) =<br />

E(X|I t , ˜∆w t−1 ), the expectation from the point of view of the ST at t. Then, the ST’s value<br />

function at t satisfies the Bellman equation:<br />

V S<br />

t<br />

(<br />

= max E S (<br />

t (vT − p t )∆x ) )<br />

+ Vt+1<br />

S . (J.67)<br />

∆x<br />

26


We conjecture a value function for the ST that is quadratic in the current signal:<br />

V F<br />

t = β 0 t−1 + β t−1<br />

(˜∆wt−1<br />

) 2. (J.68)<br />

With a similar computation as for the FT, β t satisfies the recursive equation<br />

β t−1 =<br />

(<br />

C t − ( λ t − 2β t ρ 2 )(<br />

t NF µ ∗ t + (N S − 1)νt<br />

∗ ) ) 2<br />

4λ ′ t<br />

where λ ′ t = λ t − β t ρ 2 t . We also obtain<br />

ν t<br />

+ β t ρ 2 t<br />

(<br />

NF µ ∗ t + (N S − 1)ν ∗ t<br />

) 2, (J.69)<br />

= C t − ( λ t − 2β t ρ 2 )(<br />

t NF µ ∗ t + (N S − 1)νt ∗ )<br />

2λ ′ . (J.70)<br />

t<br />

Note that if µ t = ν t , α t <strong>and</strong> β t satisfy the same equation. Thus, we can take<br />

µ t = ν t , α t = β t . (J.71)<br />

Thus, we look for a symmetric equilibrium in which γ t = γ ∗ t , µ t = ν t = µ ∗ t = ν ∗ t , <strong>and</strong> α t = β t .<br />

We get the following equations<br />

γ t =<br />

1 − 2α t ρ t<br />

(N F + 1)λ t − 2N F α t ρ 2 ,<br />

t<br />

α t−1<br />

C t<br />

µ t =<br />

(N T + 1)λ t − 2N T α t ρ 2 ,<br />

t<br />

(<br />

)<br />

= µ 2 t λ t + (NT 2 − 2N T )α t ρ 2 t .<br />

(J.72)<br />

Dealer’s Pricing Rules<br />

The dealer takes the speculator’s strategies as given, which means that she assumes the aggreate<br />

order flow to be of the form<br />

∆y t = N F γ t ∆w t + N T µ t˜∆wt−1 + du t . (J.73)<br />

Therefore, she sets λ t <strong>and</strong> ρ t according to the usual formulas:<br />

λ t = Cov t(v T , ∆y t )<br />

Var t (∆y t )<br />

ρ t = Cov t(∆v t , ∆y t )<br />

Var t (∆y t )<br />

=<br />

=<br />

˜Cov ( w t − p t−1 , ∆y t<br />

)<br />

Ṽar(∆y t )<br />

˜Cov ( )<br />

∆w t , ∆y t<br />

Ṽar(∆y t )<br />

= N F γ t + N T µ t C t<br />

D t<br />

,<br />

= N F γ t<br />

D t<br />

.<br />

(J.74)<br />

27


We now rewrite the equations from Lemma J.4, using the equation we derived above: ρ t D t =<br />

N F γ t :<br />

B t+1 = 1 − N F ρ t γ t − N T ρ t µ t B t ,<br />

A t+1 = 1 − N F ρ t γ t ,<br />

D t+1 = (N F γ t+1 ) 2 + (N T µ t+1 ) 2 (1 − N F ρ t γ t ) + ˜σ 2 u,<br />

(J.75)<br />

C t<br />

= B t<br />

A t<br />

=<br />

B t<br />

1 − N F ρ t−1 γ t−1<br />

.<br />

Equilibrium Conditions<br />

Figure J.1: Comparison of Equilibrium Weights. The figure compares the equilibrium<br />

modified weights a <strong>and</strong> b, which are a solution of the system of equations (J.80), with the modified<br />

weights a 0 <strong>and</strong> b 0 from the baseline model, in which the choice of weights does not affect the covariance<br />

structure (see Theorem 1). In each model, there are N = 1, 2, . . . , 10 identical speculators.<br />

1<br />

a 0 vs. a<br />

1<br />

b 0 vs. b<br />

0.9<br />

0.9<br />

0.8<br />

0.8<br />

0.7<br />

a 0<br />

0.7<br />

0.6<br />

a<br />

0.6<br />

a<br />

0.5<br />

0.4<br />

b<br />

0.5<br />

0.4<br />

b<br />

b 0<br />

0.3<br />

0.3<br />

0.2<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

1 2 3 4 5 6 7 8 9 10<br />

m<br />

0<br />

1 2 3 4 5 6 7 8 9 10<br />

m<br />

Following the continuous time version of the model, we define the following variables:<br />

a t = N F ρ t γ t , b t = N T ρ t µ t , R t = λ t<br />

ρ t<br />

.<br />

(J.76)<br />

From ρD t = N F γ t , we obtain N F ρ t γ t = ρ 2 t D t = (N F ρ t γ t ) 2 +(N T ρ t µ t ) 2 (1−N F ρ t−1 γ t−1 )+ρ 2˜σ 2 u.<br />

With the new notation, this equation becomes<br />

a t = a 2 t + b 2 t (1 − a t−1 ) + ρ 2 t ˜σ 2 u. (J.77)<br />

28


Also, we compute<br />

R t<br />

= λ t<br />

ρ t<br />

= a t + b t B t<br />

ρ 2 D t<br />

= a t + b t B t<br />

a t<br />

= 1 + b tB t<br />

a t<br />

. (J.78)<br />

We put together the equations that determine the equilibrium:<br />

a t =<br />

1 − 2α t ρ t<br />

,<br />

N F +1<br />

N F<br />

R t − 2α t ρ t<br />

C t<br />

b t =<br />

,<br />

N T +1<br />

N T<br />

R t − 2α t ρ t<br />

B t<br />

C t = ,<br />

1 − a t−1<br />

a t = a 2 t + b 2 t (1 − a t−1 ) + ρ 2 t ˜σ 2 u,<br />

(J.79)<br />

R t<br />

= 1 + b tB t<br />

a t<br />

,<br />

B t+1 = 1 − a t − b t B t = 1 − R t a t ,<br />

(<br />

α t−1 = b 2 Rt<br />

(<br />

t<br />

NT 2ρ + 1 − 2 ) )<br />

α t .<br />

t N T<br />

In the spirit of Lemma I.2, we treat all these numbers as constant, by removing their dependence<br />

on t. 45 For instance, from the recursive equation for B, we have B = 1−a<br />

1+b , which<br />

coincides with the value of B 1 from the continuous time version. We obtain the following<br />

equations:<br />

B = 1 − a<br />

1 + b , C = 1<br />

1 + b , R = a + b<br />

a(1 + b) , ρ2˜σ 2 u = (a − b) 2 (1 − a),<br />

α =<br />

b2 R<br />

(<br />

NT 2ρ + b2 1 − 2 )<br />

α, a =<br />

N T<br />

Solving the equation for α, <strong>and</strong> multiplying by 2ρ, we obtain<br />

(J.80)<br />

1 − 2αρ<br />

N F +1<br />

N F<br />

R − 2αρ , b = C<br />

N T +1<br />

N T<br />

R − 2αρ .<br />

2αρ =<br />

2b 2 R<br />

1 − b 2( 1 − 2 ) 1<br />

N 2 . (J.81)<br />

N T T<br />

Note that the first 4 equations in (J.81) coincide with the corresponding ones in the continuous<br />

time case. However, the last 2 equations, for γ <strong>and</strong> µ differ from the continuous time value<br />

by the term 2αρ. From (J.81), <strong>and</strong> using the fact that all the other terms are of order one,<br />

1<br />

we see that the term 2αρ is of the order of :<br />

NT<br />

2<br />

(<br />

2αρ = O 1<br />

)<br />

NT N . (J.82)<br />

T<br />

2<br />

45 The system of equations can be solved numerically, <strong>and</strong> for all the parameter values we have checked, the<br />

solutions are numerically very close to a constant, except when t is either close to 0, or close to T .<br />

29


We now describe a numerical procedure which computes with high accuracy a solution<br />

(a, b) of the system above. Denote by a 0 <strong>and</strong> b 0 the corresponding equilibrium values from<br />

Theorem 1, in which the choice of weights does not affect the covariance structure. Then,<br />

starting with (a 0 , b 0 ), we compute R 0 = a0 +b 0<br />

a 0 (1+b 0 ) , <strong>and</strong> then 2α0 ρ 0 =<br />

2(b 0 ) 2 R 0<br />

)<br />

1−(b 0 )<br />

(1− 1<br />

2 2 NT<br />

2 N T<br />

using<br />

(J.81). Using (J.80), we recompute the values of (a, b), <strong>and</strong> denote them (a 1 , b 1 ). Iterate<br />

the procedure until it converges. Then, define (a, b) = lim n→∞ (a n , b n ). Then, (a, b) satisfy<br />

the system of equations in (J.80). In Figure J.1, we plot the solution for the case when there<br />

are only FTs, <strong>and</strong> their number is N ∈ {1, . . . , 10}.<br />

(The introduction of STs makes the<br />

approximation even better, since N T = N F + N S increases.) From the figure, we see that the<br />

approximation is good even for low N.<br />

K<br />

Inventory Management<br />

In this section, we prove Theorem 2 <strong>and</strong> Proposition 5 from Section 5.<br />

K.1 Proof of Theorem 2<br />

The Smooth Regime<br />

Denote by<br />

Ω xx<br />

t<br />

= E( x 2 )<br />

t<br />

σw<br />

2 , Ω xe<br />

t = E( x t (w t − p t ) )<br />

σw<br />

2 , E t = E( (w t − p t ) ˜dw<br />

)<br />

t<br />

σw 2 . (K.1)<br />

dt<br />

Define the following aggregate coefficients for the N I speculators who manage inventories:<br />

θ I = N I θ, γ I = N I γ ∗ , µ I = N I µ ∗ . (K.2)<br />

Define also the following variables:<br />

Ω xw<br />

t = E( x t w t ) )<br />

σ 2 w<br />

, Ω xp<br />

t = E( )<br />

x t p t<br />

. (K.3)<br />

σ 2 w<br />

Note first that<br />

From the definition of Ω xw , we get<br />

Ω xp<br />

t<br />

= Ω xw<br />

t<br />

− Ω xe<br />

t . (K.4)<br />

dΩ xw<br />

t<br />

dt<br />

= 1<br />

σ 2 wdt E( dx t w t−1 + x t−1 dw t + dx t dw t<br />

)<br />

= −θΩ xw<br />

t−1 + γ ∗ .<br />

(K.5)<br />

30


As there is no initial inventory, Ω xw<br />

0 = 0. Thus, the solution for the differential equation (K.5)<br />

is:<br />

In order to compute Ω xx<br />

t<br />

X t = E( x t ˜dwt<br />

)<br />

σ 2 w dt<br />

Ω xw<br />

t<br />

1 − e −θt<br />

= γ ∗ . (K.6)<br />

θ<br />

<strong>and</strong> Ω xe<br />

t , we need to define additional covariances. 46 Denote by<br />

, W t = E( w t ˜dwt<br />

)<br />

σ 2 w dt<br />

, P t = E( p t ˜dwt<br />

)<br />

σ 2 w dt<br />

, E t = E( (w t − p t ) ˜dw t<br />

)<br />

σ 2 w dt<br />

. (K.7)<br />

To compute these covariances, we derive recursive formulas for them. Note that the aggregate<br />

order flow at t is of the form:<br />

dy t = −θ I x t−1 dt + ¯γdw t + ¯µ ˜dw t−1 + du t , (K.8)<br />

where ¯γ <strong>and</strong> ¯µ are the aggregate equilibrium coefficients. To simplify notation, denote by<br />

ā = ρ¯γ, ¯b = ρ¯µ, Ā1,1 = 1 − ā, ¯B1 = 1 − ā . (K.9)<br />

1 + ¯b<br />

Also, ˜dw t = dw t − ρdy t can be written as follows:<br />

˜dw t = ρθ I x t−1 dt + dw t (1 − ā) − ¯b ˜dw t−1 − ρdu t . (K.10)<br />

The recursive formula for X t is:<br />

X t = 1<br />

σ 2 wdt E ( (xt−1<br />

+ dx t<br />

)(<br />

ρθ I x t−1 dt + dw t (1 − ā) − ¯b ˜dw t−1 − ρdu t<br />

) )<br />

= ρθ I E( (x t−1 ) 2)<br />

σ 2 wdt<br />

− ¯b E( x t−1 ˜dwt−1<br />

)<br />

σ 2 wdt<br />

= ρθ I Ω xx<br />

t−1 − ¯bX t−1 + (1 − ā)γ ∗ .<br />

+ (1 − ā) E( dx t dw t<br />

)<br />

σ 2 wdt<br />

− ¯b E( dx t ˜dwt−1<br />

)<br />

σ 2 wdt<br />

(K.11)<br />

Thus, X t + bX t−1 = ρθ I Ω xx<br />

t−1 + (1 − ā)γ ∗. Since ¯b ∈ (0, 1), we use Lemma I.2 to obtain the<br />

following formula: 47 X t = ρθI Ω xx<br />

t<br />

+ (1 − ā)γ ∗<br />

1 + ¯b<br />

= ρθI Ω xx<br />

t<br />

1 + ¯b + ¯B 1 γ ∗ . (K.12)<br />

46 The inventory management term is of the ( order of dt, <strong>and</strong> thus it does not affect any instantaneous<br />

covariances with infinitesimal terms, such as E ( ˜dw t ) 2)<br />

= A<br />

σw 1,1. However, covariances with aggregate measures<br />

2 dt<br />

such as x t, w t, p t are affected by the(<br />

slow ) accumulation of the dt term. For instance, as we see from the<br />

formula (K.14) for W t, the equation E w t ˜dwt<br />

= B<br />

σw 2 dt 1 is no longer true here.<br />

47 The difference between Ω xx<br />

t <strong>and</strong> Ω xx<br />

t−1 is infinitesimal, hence it can be ignored. In other words, one can<br />

use the Lemma either for α t or for α t−1, <strong>and</strong> obtain the same result.<br />

31


Similarly, the recursive formula for W t is:<br />

W t = 1 ( (wt−1<br />

σwdt 2 E )(<br />

+ dw t ρθ I x t−1 dt + dw t (1 − ā) − ¯b ˜dw<br />

) )<br />

t−1 − ρdu t<br />

= ρθ I E( )<br />

w t−1 x t−1<br />

σwdt<br />

2 − ¯b E( )<br />

w t−1 ˜dwt−1<br />

σwdt<br />

2 + (1 − ā) E( )<br />

dw t dw t<br />

σwdt<br />

2<br />

= ρθ I Ω xw<br />

t−1 − ¯bW t−1 + (1 − ā).<br />

(K.13)<br />

Thus, W t + ¯bW t−1 = ρθ I Ω xw<br />

t−1 + (1 − ā). Since ¯b ∈ (0, 1), we use Lemma I.2 to obtain the<br />

following formula:<br />

W t<br />

The recursive formula for P t is:<br />

= ρθI Ω xw<br />

t + (1 − ā)<br />

1 + ¯b<br />

= ρθI Ω xw<br />

t<br />

1 + ¯b<br />

P t = 1 ( (pt−1<br />

σwdt 2 E )(<br />

+ λdy t ρθ I x t−1 dt + dw t (1 − ā) − ¯b ˜dw<br />

) )<br />

t−1 − ρdu t<br />

= ρθ I E( )<br />

p t−1 x t−1<br />

+ (1 − ā)λ E( )<br />

dy t dw t<br />

σ 2 wdt<br />

− ¯b E( p t−1 ˜dwt−1<br />

)<br />

σ 2 wdt<br />

= ρθ I Ω xp<br />

t−1 − ¯bP t−1 + (1 − ā)λ¯γ − ¯bλ¯µĀ1,1 − λρ˜σ 2 u.<br />

σ 2 wdt<br />

+ ¯B 1 . (K.14)<br />

− ¯bλ E( dy t ˜dwt−1<br />

)<br />

σ 2 wdt<br />

− λρ˜σ 2 u<br />

(K.15)<br />

Define<br />

¯M = (1 − ā)λ¯γ − bλ¯µĀ1,1 − λρ˜σ 2 u.<br />

(K.16)<br />

One can check that ¯M = 0 when ā = a <strong>and</strong> ¯b = b are the equilibrium values of Section 3, which<br />

simply reflects the fact that ˜dw t is orthogonal to p t in the absence of inventory management,<br />

i.e., P t = 0 when θ = 0. 48 As in the case of X t , we use Lemma I.2 to obtain<br />

P t<br />

= ρθI Ω xp<br />

t + ¯M<br />

1 + ¯b<br />

= ρθI (Ω xw<br />

t − Ω xe<br />

t ) + ¯M<br />

1 + ¯b . (K.17)<br />

From (K.14) <strong>and</strong> (K.17) we also get<br />

E t = W t − P t = ρθI Ω xe<br />

t − ¯M<br />

1 + ¯b<br />

+ ¯B 1 . (K.18)<br />

We now compute Ω xx<br />

t . From its definition, we have<br />

dΩ xx<br />

t<br />

dt<br />

= 1<br />

σ 2 wdt E( 2x t−1 dx t + (dx t ) 2)<br />

= 1<br />

σ 2 wdt E (<br />

2x t−1<br />

(<br />

−θxt−1 dt + γdw t + µ ˜dw t−1<br />

)<br />

+ (dxt ) 2) .<br />

= −2θΩ xx<br />

t−1 + γ 2 ∗.<br />

(K.19)<br />

48 Indeed, using ρ 2˜σ 2 u = (1 − a)(a − b 2 ), we compute λ ρ<br />

(<br />

(1 − a)a − (1 − a)b 2 − (1 − a)(a − b 2 ) ) = 0.<br />

32


This is a first order ODE with solution:<br />

Ω xx<br />

t<br />

= γ 2 ∗<br />

1 − e −2θt<br />

. (K.20)<br />

2θ<br />

Finally, we compute Ω xe<br />

t . Since dw t − dp t = λθ I x t−1 dt + (1 − λ¯γ)dw t − λ¯µ ˜dw t−1 − λdu t , from<br />

the definition of Ω xe<br />

t , we have<br />

dΩ xe<br />

t<br />

dt<br />

= 1<br />

σ 2 wdt E( (w t−1 − p t−1 )dx t + x t−1 (dw t − dp t ) + (dw t − dp t )dx t<br />

)<br />

= −θΩ xe<br />

t−1 + λθ I Ω xx<br />

t−1 − λ¯µX t−1 + (1 − λ¯γ)γ ∗ .<br />

(K.21)<br />

From (K.12), we have X t−1 = ρθI Ω xx<br />

t−1<br />

+ ¯B 1+¯b 1 γ ∗ . We obtain<br />

dΩ xe<br />

t<br />

dt<br />

= −θ<br />

(1 − ρµI )<br />

1 + ¯b Ω xe<br />

t−1 + λθ I( 1 −<br />

ρ¯µ )<br />

1 + ¯b Ω xx<br />

t−1 − λ¯µ ¯B 1 γ ∗ + (1 − λ¯γ)γ ∗<br />

= −θΩ xe<br />

t−1 + λθ I 1<br />

1 + b Ωxx t−1 − λ¯µB 1 γ ∗ + ˜π 0 .<br />

(K.22)<br />

where ˜π 0 is the normalized expected profit when θ = 0:<br />

˜π 0 = (1 − λ¯γ)γ ∗ . (K.23)<br />

From (K.20), λθ I 1<br />

1+¯b Ωxx t−1 = λθI<br />

1+¯b γ2 ∗ 1−e−2θt<br />

2θ<br />

becomes:<br />

Ω xe<br />

t<br />

= N Iλγ 2 ∗<br />

2(1+¯b) (1 − e−2θt ). The differential equation for<br />

dΩ xe<br />

t<br />

dt<br />

= −θΩ xe<br />

t−1 + D 1<br />

(<br />

1 − e<br />

−2θt ) + D 2 ,<br />

D 1 = N Iλγ 2 ∗<br />

2(1 + ¯b) , D 2 = ˜π 0 − λ¯µ ¯B 1 γ ∗ .<br />

with<br />

(K.24)<br />

This is a first order ODE with solution:<br />

Ω xe<br />

t<br />

= ( D 1 + D 2<br />

)1 − e −θt<br />

θ<br />

e −θt − e −2θt<br />

+ D 1 . (K.25)<br />

θ<br />

Next, we compute the HFT expected profit at t = 0 in the smooth regime:<br />

˜π θ = 1<br />

σw<br />

2 E<br />

= 1 E<br />

=<br />

σw<br />

2 ∫ 1<br />

0<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

(w t − p t )dx t<br />

(<br />

wt−1 − p t−1 + dw t − λdy t<br />

) ( γ ∗ dw t − θx t−1 dt)<br />

(<br />

−θΩ xe<br />

t−1 + γ ∗ − λγ ∗¯γ<br />

)<br />

dt.<br />

(K.26)<br />

33


Therefore,<br />

From (K.25),<br />

˜π θ<br />

= ˜π 0 −<br />

∫ 1<br />

0<br />

θΩ xe<br />

t dt. (K.27)<br />

θΩ xe<br />

t = N Iλγ∗<br />

2 (<br />

2(1 + ¯b) 1 − e −2θt) + (˜π 0 − λ¯µ ¯B<br />

) (<br />

1 γ ∗ 1 − e −θt) , (K.28)<br />

with D 1 <strong>and</strong> D 2 as in (K.24). We compute<br />

˜π θ = ˜π 0 e −θt +λ¯µ ¯B 1 γ ∗<br />

∫ 1<br />

0<br />

(<br />

1 − e<br />

−θt ) dt − N Iλγ 2 ∗<br />

2(1 + ¯b)<br />

which is as stated in the Theorem. This expression has a limit when θ → ∞:<br />

∫ 1<br />

0<br />

(<br />

1 − e<br />

−2θt ) dt, (K.29)<br />

lim<br />

θ→∞ ˜πθ<br />

= λ¯µ ¯B 1 γ ∗ − N Iλγ∗<br />

2 . (K.30)<br />

2(1 + ¯b)<br />

The Extreme Regime<br />

Here, HFTs have the following mean-reverting strategy:<br />

dx Θ t = −Θx t−1 + γ ∗ dw t , (K.31)<br />

where x t is the inventory of each of the N I HFTs. Denote by<br />

Ω xx<br />

t<br />

Ω xw<br />

= E( x 2 )<br />

t<br />

σwdt 2 ,<br />

t = E( )<br />

x t w t<br />

σwdt<br />

2<br />

Ωxe t = E( x t (w t − p t ) )<br />

σwdt<br />

2 , X t = E( )<br />

x t ˜dwt<br />

σwdt<br />

2<br />

, Ω xp<br />

t = E( )<br />

x t p t<br />

.<br />

σwdt 2 , Z t = E( )<br />

x t−1 dy t<br />

σwdt<br />

2<br />

(K.32)<br />

Recall that<br />

φ = 1 − Θ.<br />

(K.33)<br />

We write x t in AR(1)-form, with autoregressive coefficient φ:<br />

x t = x t−1 + dx t = (1 − Θ)x t−1 + γ ∗ dw t = φx t−1 + γ ∗ dw t<br />

∑<br />

∞<br />

= γ ∗ (1 − Θ) i dw t−i .<br />

i=1<br />

(K.34)<br />

Note that the last equality is the moving average representation. We also write the aggregate<br />

order flow at t:<br />

dy t = −Θ I x t−1 + ¯γdw t + ¯µ ˜dw t−1 + du t , Θ I = N I Θ. (K.35)<br />

34


Using (K.34), we compute<br />

Ω xw<br />

t = E( w t x t<br />

)<br />

σ 2 wdt<br />

∑<br />

∞<br />

= γ ∗ (1 − Θ) i = γ ∗<br />

Θ .<br />

i=1<br />

(K.36)<br />

The recursive formula for Ω xx<br />

t<br />

is<br />

Ω xx<br />

t = E( (x t ) 2)<br />

σ 2 wdt<br />

= E( φ 2 x 2 t−1 + 2φx t−1γ ∗ dw t + γ 2 ∗(dw t ) 2)<br />

σ 2 wdt<br />

= φ 2 Ω xx<br />

t−1 + γ 2 ∗.<br />

(K.37)<br />

From this, we obtain the usual variance formula for the AR(1) process:<br />

Ω xx<br />

t = γ2 ∗<br />

1 − φ 2 = γ 2 ∗<br />

Θ(1 + φ) . (K.38)<br />

Since dy t = −Θ I x t−1 + ¯γdw t + ¯µ ˜dw t−1 + du t , we express Z t as a function of X t−1 :<br />

Z t = E( x t−1 dy t<br />

)<br />

σ 2 wdt<br />

= −Θ I Ω xx<br />

t−1 + ¯µX t−1 = −N I<br />

γ 2 ∗<br />

1 + φ + ¯µX t−1. (K.39)<br />

The recursive formula for X t is:<br />

X t = E( x t ˜dwt<br />

)<br />

σ 2 wdt<br />

= −ρφZ t + γ ∗ − ργ ∗¯γ<br />

= E( (φx t−1 + γ ∗ dw t )(dw t − ρdy t ) )<br />

σ 2 wdt<br />

(K.40)<br />

= N I ρφ γ2 ∗<br />

1 + φ − ρφ¯µX t−1 + γ ∗ − ργ ∗¯γ.<br />

The coefficient ρφ¯µ = bφ ∈ (0, 1), hence we apply Lemma I.2 to compute<br />

X t<br />

= N I ρφγ2 ∗<br />

1+φ + γ ∗(1 − ρ¯γ)<br />

. (K.41)<br />

1 + φρ¯µ<br />

We also compute<br />

Z t<br />

= −Θ I Ω xx<br />

t−1 + ¯µX t−1<br />

= ¯µ N I ρφγ2 ∗<br />

1+φ + γ ∗(1 − ρ¯γ) γ∗<br />

2 − N I<br />

(K.42)<br />

1 + φρ¯µ<br />

1 + φ<br />

= ¯µγ ∗(1 − ρ¯γ)<br />

1 + φρ¯µ − N γ∗<br />

2 I<br />

(1 + φ)(1 + φρ¯µ) .<br />

35


The recursive formula for Ω xp<br />

t is<br />

Ω xp<br />

t = E( (p t−1 + λdy t )(φx t−1 + γ ∗ dw t ) )<br />

σ 2 wdt<br />

= φΩ xp<br />

t−1 + λφZ t + λγ ∗¯γ.<br />

(K.43)<br />

Since 1 − φ = Θ ∈ (0, 1), we apply Lemma I.2 to compute<br />

Ω xp<br />

t<br />

= λφZ t + λγ ∗¯γ<br />

. (K.44)<br />

Θ<br />

From (K.36) <strong>and</strong> (K.44), we get<br />

−ΘΩ xe<br />

t<br />

= − ( ΘΩ xw<br />

t<br />

− ΘΩ xp )<br />

t = −γ∗ + λφZ t + λγ ∗¯γ. (K.45)<br />

We compute the expected profit of a HFT:<br />

˜π Θ = 1<br />

σw<br />

2 E<br />

= 1 E<br />

=<br />

=<br />

σw<br />

2 ∫ 1<br />

0<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

(w t − p t )dx t<br />

(<br />

wt−1 − p t−1 + dw t − λdy t<br />

)(<br />

γ∗ dw t − Θx t−1<br />

)<br />

(<br />

−ΘΩ xe<br />

t−1 + γ ∗ − λ¯γγ ∗ + λΘZ t<br />

)dt<br />

( (−γ∗<br />

+ λφZ t + λγ ) )<br />

∗¯γ + γ ∗ − λ¯γγ ∗ + λΘZ t dt<br />

= λZ<br />

( ¯µγ∗ (1 − ρ¯γ)<br />

= λ<br />

1 + φρ¯µ − N γ 2 )<br />

∗<br />

I<br />

.<br />

(1 + φ)(1 + φρ¯µ)<br />

(K.46)<br />

This is the same as the formula stated in the Theorem. One can see that by setting Θ = 0, or<br />

equivalently by setting φ = 1, we get the same value as the limit of the smooth regime profit<br />

when θ → ∞.<br />

K.2 Proof of Proposition 5<br />

In the smooth regime, let θ = 0 (no inventory management). Recall that γ ∗ = γ. We use the<br />

notations from the proof of Theorem 1. Since ¯γ <strong>and</strong> ¯µ are the equilibrium values, we compute<br />

1 − λ¯γ = 1 − λ ρ a = 1 − a+b<br />

a(1+b) a = 1−a<br />

1+b = 1<br />

N F +1 . Then,<br />

π Sm,θ=0<br />

σ 2 w<br />

= γ(1 − λ¯γ) = γ 1 − a<br />

1 + b ≈ γ<br />

N F<br />

.<br />

(K.47)<br />

36


Next, compute the HFT profit in the extreme regime:<br />

π Ex,Θ<br />

σ 2 w<br />

= λ λ<br />

1 − a<br />

ρ 1 + φb bγ − ρ γ ργ<br />

(1 + φ)(1 + φb) = λ ρ<br />

(<br />

γ λ<br />

N F<br />

=<br />

b(1 + b)<br />

N F (1 + φb) ρ N F + 1 −<br />

1 + b<br />

1 + φb<br />

)<br />

.<br />

a<br />

1 + φ<br />

bγ<br />

1 + N F<br />

−<br />

λ<br />

ρ γ a<br />

N F<br />

(1 + φ)(1 + φb)<br />

(K.48)<br />

√<br />

From Section 3, we know that when N F is large, a ≈ 1; b ≈ b ∞ = 5−1<br />

2<br />

, which satisfies<br />

b ∞ (b ∞ + 1) = 1; <strong>and</strong> λ ρ ≈ 1. Thus, the profit is approximately (within O ( 1<br />

)<br />

N F N F<br />

)<br />

π Ex,Θ<br />

σ 2 w<br />

≈<br />

γ<br />

N F (1 + φb ∞ )<br />

φ<br />

1 + φ . (K.49)<br />

The other formulas in the Proposition follow from the above formulas, by specializing to the<br />

cases Θ = 0 (φ = 1) or Θ = 1 (φ = 0).<br />

L<br />

Robust Trading Strategies<br />

In this section, we prove Propositions 10 <strong>and</strong> 11 from Section 6.3.<br />

L.1 Proof of Proposition 10<br />

Speculators’ Optimal Strategy (γ, φ)<br />

The expected profit at t = 0 of the i’th w-speculator is<br />

π i w,0<br />

= E<br />

=<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

⎛<br />

⎝w t + e t − p t−1 − λ t<br />

((γ w,t i + γw,t −i )dw ∑N e<br />

) ⎞<br />

t + γ j e,t de t + du t<br />

⎠ γw,tdw i t<br />

γw,tσ i wdt 2 − λ t γw,t<br />

i (<br />

γ<br />

i<br />

w,t + γw,t) −i σ<br />

2<br />

w dt,<br />

j=1<br />

(L.1)<br />

where γ −i<br />

w,t denotes the aggregate coefficient of the other N w − 1 w-speculators.<br />

This is a<br />

pointwise quadratic optimization problem, with solution λ t γw,t i = 1−λtγ−i w,t<br />

2<br />

. Since this is true<br />

for each w-speculator, we obtain that all γw,t i 1<br />

are equal to γ w,t =<br />

λ . Similarly, all t(N w+1) γj e,t<br />

1<br />

are equal to γ e,t =<br />

λ t(N e+1)<br />

. Combining these two equations, we get<br />

γ i w,t = γ w,t =<br />

Dealer’s Pricing Rule (λ)<br />

1<br />

λ t (N w + 1) , γj e,t = γ e,t =<br />

1<br />

λ t (N e + 1) .<br />

(L.2)<br />

The dealer takes the speculators’ strategies as given, thus assumes that the aggregate order<br />

flow is of the form dy t = N w γ t dw t + N e φ t de t + du t . To set λ t , the dealer sets p t such that the<br />

37


market is efficient, which implies<br />

dp t = λ t dy t , with λ t = Cov(w t + e t , dy t )<br />

Var(dy t )<br />

=<br />

N w γ w,t σw 2 + N e γ e,t σe<br />

2<br />

Nwγ 2 w,t 2 σ2 w + Ne 2 γt,e 2 σ2 e + σu<br />

2 . (L.3)<br />

This implies (N w λ t γ w,t ) 2 σ 2 w+(N e λ t γ e,t ) 2 σ 2 e+λ 2 t σ 2 u = N w λ t γ w,t σ 2 w+N e λ t γ e,t σ 2 e. But N w λ t γ w,t =<br />

N w<br />

N w+1 <strong>and</strong> N eλ t γ e,t =<br />

Ne<br />

N . Hence, e+1 λ2 t σu 2 =<br />

Nw<br />

(N w+1) 2 σ 2 w + Ne<br />

(N e+1) 2 σ 2 e, which implies<br />

λ t = λ =<br />

(<br />

N w σw<br />

2 N e σ 2 ) 1/2<br />

e<br />

(N w + 1) 2 σu<br />

2 +<br />

(N e + 1) 2 σu<br />

2 . (L.4)<br />

This proves the stated formulas.<br />

L.2 Proof of Proposition 11<br />

The trading strategy of the β-speculator is of the form:<br />

dx t = β t (w t−1 − p t−1 )dt + γ w dw t (L.5)<br />

where γ w is the equilibrium (constant) value. Denote the aggregate coefficients by<br />

¯γ w = N w γ w , ¯γ e = N e γ e . (L.6)<br />

Define the following (normalized) covariances<br />

Σ t = E( (w t − p t ) 2)<br />

σ 2 w<br />

, Ω t = E( e t p t<br />

)<br />

σ 2 w<br />

, ˜σ e 2 = σ2 e<br />

σw<br />

2 . (L.7)<br />

Since w 0 = e 0 = p 0 , we have<br />

Σ 0 = Ω 0 = 0. (L.8)<br />

The normalized expected profit of the β-speculator at t = 0 is:<br />

˜π β = 1 ∫ 1<br />

(<br />

(<br />

) )<br />

σw<br />

2 E w t + e t − p t−1 − λ β t (w t−1 − p t−1 )dt + ¯γ w dw t + ¯γ e de t + du t ×<br />

0<br />

)<br />

×<br />

(β t (w t−1 − p t−1 )dt + γ w dw t<br />

=<br />

∫ 1<br />

0<br />

= ˜π 0 +<br />

)<br />

(γ w − λγ w¯γ w + β t Σ t−1 − β t Ω t−1 dt<br />

∫ 1<br />

0<br />

)<br />

(β t Σ t−1 − β t Ω t−1 dt,<br />

(L.9)<br />

where ˜π 0 is the normalized profit of the β-speculator when β = 0, i.e.,<br />

˜π 0 = γ w − λγ w¯γ w =<br />

γ w<br />

N w + 1 = 1<br />

λ(N w + 1) 2 .<br />

(L.10)<br />

38


Since dw t − dp t = −λβ t (w t−1 − p t−1 )dt + (1 − λ¯γ w )dw t − λ¯γ e de t − λdu t , Σ t satisfies:<br />

dΣ t<br />

dt<br />

= 1<br />

σ 2 wdt E( 2(w t−1 − p t−1 )(dw t − dp t ) + (dw t − dp t ) 2)<br />

= −2λβ t Σ t−1 + (1 − λ¯γ w ) 2 + (λ¯γ e ) 2˜σ 2 e + λ 2˜σ 2 u.<br />

(L.11)<br />

This is a first order ODE with solution:<br />

Σ t<br />

B t =<br />

∫ t<br />

= D Σ e −2λBt e 2λBτ dτ,<br />

∫ t<br />

0<br />

0<br />

with<br />

β τ dτ, D Σ = (1 − λ¯γ w ) 2 + (λ¯γ e ) 2˜σ 2 e + λ 2˜σ 2 u.<br />

(L.12)<br />

We recall now a formula we derived in the computation of λ:<br />

(λ¯γ w ) 2 + (λ¯γ e ) 2˜σ 2 e + λ 2˜σ 2 u = λ¯γ w + λ¯γ e˜σ 2 e. (L.13)<br />

Then, we compute<br />

D Σ = 1 − λ¯γ w + λ¯γ e˜σ 2 e =<br />

1<br />

N w + 1 + N e<br />

N e + 1 ˜σ2 e.<br />

(L.14)<br />

By integrating (L.11) over [0, 1] (<strong>and</strong> using Σ 0 = 0), we also compute<br />

∫ 1<br />

0<br />

β t Σ t−1 dt = D Σ − Σ 1<br />

. (L.15)<br />

2λ<br />

Since dp t = λβ t (w t−1 − p t−1 )dt + λ¯γ w dw t + λ¯γ e de t + λdu t , Ω t satisfies:<br />

dΩ t<br />

dt<br />

= 1<br />

σ 2 wdt E( p t−1 de t + e t−1 dp t + de t dp t<br />

)<br />

= −λβ t Ω t−1 + λ¯γ e˜σ 2 e.<br />

(L.16)<br />

This is a first order ODE with solution:<br />

Ω t<br />

∫ t<br />

= D Ω e −λBt e λBτ dτ,<br />

D Ω = λ¯γ e˜σ 2 e = N e<br />

N e + 1 ˜σ2 e.<br />

0<br />

with<br />

(L.17)<br />

By integrating (L.16) over [0, 1] (<strong>and</strong> using Ω 0 = 0), we also compute<br />

∫ 1<br />

0<br />

β t Ω t−1 dt = D Ω − Ω 1<br />

. (L.18)<br />

λ<br />

39


We compute<br />

∫ 1<br />

∫ 1<br />

Σ 1 = D Σ e −2λ(B 1−B t) dt, Ω 1 = D Ω e −λ(B 1−B t) dt.<br />

Combining the formulas above, we compute<br />

˜π β<br />

B t =<br />

= ˜π 0 + D Σ<br />

2λ<br />

∫ t<br />

0<br />

0<br />

(<br />

1 −<br />

∫ 1<br />

0<br />

β τ dτ, D Σ =<br />

)<br />

e −2λ(B 1−B t) dt − D Ω<br />

λ<br />

(<br />

1 −<br />

0<br />

∫ 1<br />

0<br />

)<br />

e −λ(B 1−B t) dt ,<br />

1<br />

N w + 1 + N e<br />

N e + 1 ˜σ2 e, D Ω = N e<br />

N e + 1 ˜σ2 e.<br />

(L.19)<br />

(L.20)<br />

If we define<br />

ε t = e −λ(B 1−B t) = e −λ(∫ 1<br />

t βτ dτ) ∈ [0, 1], (L.21)<br />

we compute<br />

˜π β<br />

= ˜π 0 + 1 λ<br />

= ˜π 0 + 1<br />

2λ<br />

∫ 1<br />

( ) ( )<br />

1 + ε t<br />

1 − εt D Σ − D Ω dt<br />

2<br />

( ) ( 1 + ε t<br />

1 − εt<br />

N w + 1 − (1 − ε t)N e˜σ e<br />

2<br />

N e + 1<br />

0<br />

∫ 1<br />

0<br />

)<br />

dt.<br />

(L.22)<br />

This completes the proof.<br />

40

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!