US Term Premia around FOMC decisions∗ Nikola Mirkov† First draft: January 2011 This version: December 2011

Abstract This paper studies instantaneous reactions of the term premia implicit in the US interest rates to the Federal Open Market Committee (FOMC) policy rate decisions. The estimation results suggest that the expansionary policy actions, considered as anticipated, cause on average a contemporaneous decline in forward term premia and a rise in the short-rate expectations on the longer-end of the curve. A surprise cut seem to provoke a parallel downward shift of the expected short-rates and a spike in longer term premia. This negative co-movement between the premia and the short-rate expectations, implicit in higher maturities, might provide one explanation of the so called “slope effect” of monetary policy on the US yield curve. The findings are independent of the market price of risk specification or whether the model with a single- or two parameter set is used in the assessment. Nonetheless, allowing a separate set of parameters around policy action days indicates a greater role of the slope factor in explaining the variation of long-term yields around those days. Keywords: term structure, FOMC, policy actions, term premia, Bayesian inference JEL Classifications: E43, E52, C11, G12



This paper is entirely based on the first Chapter of my perspective PhD thesis at the University St.Gallen. I gratefully benefited from key advices and constant support of my doctoral advisor Paul Söderlind. A great thanks to Sylvia Frühwirth-Schnatter, Hanna Samaryna, Marvin Goodfriend, Francesco Audrino, Magnus Dahlquist, Henrik Hasseltoft, Stephan Süss, Pavol Povala, Norman Seeger, Evert Wipplinger, Aleksandar Zdravkovic, the seminar and conference participants at the University St.Gallen and the 28th GdRE Annual International Symposium on Money, Banking and Finance. † Nikola Nikodijevic Mirkov, University St.Gallen, Swiss Institute of Banking and Finance, Rosenbergstr. 52, 9000 St.Gallen, E-mail: [email protected], Tel: +41(0)76.22.98.176

1

1

Introduction

Understanding the implications of a Central Bank policy rate decisions on the yield curve is important for assessing the key monetary policy transmission mechanism - expected path of future short-term interest rates. The rational expectations hypothesis of the term structure1 suggests that the “expectations channel” is the main driver of the long-term interest rates, which influence both aggregate demand and aggregate investments2 . Yet another important element of the long-term interest rates3 represents the time-varying term premium - the difference between the long-term yields and the expected future short-term rates. By using the high-frequency identification strategy from Piazzesi (2005), this study shows that monetary policy in the US seem to affect both short-rate expectations and term premia implicit in the long-term rates. The key assumption of the identification strategy is that the Fed uses all available information, including the yield curve, to deliver a policy rate decision at a certain point in time, and that the subsequent change in yields is considered to be a reaction of the yield curve to the Fed decision. To this end, a three-factor no-arbitrage yield curve model with two separate sets of parameters is estimated using daily data. The first parameter set captures the yield curve dynamics on usual days in the sample, whereas the second set is used for the days when the Federal Open Market Committee (FOMC) releases the policy rate decision. There are totally 82 FOMC policy rate decisions delivered from the beginning of 1999 to the end of 2008, when the Fed funds rate reached the zero-lower bound.4 All decisions are classified into anticipated and surprise policy actions by using the surprise indicator constructed from the Fed funds futures.5 In estimating the model, the level and the slope factor are backed-out from yields, while the implied volatility on longer-maturities interest rate caps, shown to be related to proxies for macroeconomic uncertainty and the term spread, is used as the third state variable. The likelihood function is specified following the ideas from regime-switching term structure models6 and using the Bayesian Markov Chain 1

See Taylor (1995), Rudebusch (1995), Svensson (2003) and Woodford (2003) among many oth-

ers. 2

See for instance Goodfriend (1998), Svensson (2003) and Bernanke (2004). See Duffee (2002), Cochrane and Piazzesi (2009), Kim and Wright (2005), Ang, Bekaert and Wei (2008), Rudebusch, Sack and Swanson (2006), Swanson (2007) and Joslin, Priebsch and Singleton (2010) among many studies. 4 It might be important to notice at this point that, if there is a strong positive relationship between the level of the yield curve and risk premia, as shown in Kim and Singleton (2010), the reactions of the premia might be somewhat compressed by the proximity of the zero lower bound. This was kindly indicated by Prof. Marvin Goodfriend. 5 See Kuttner (2001), Bernanke and Kuttner (2005) and Gurkaynak, Sack and Swanson (2002). 6 Most notably, Ang and Bekaert (2002), Bansal and Zhou (2002) and Dai, Singleton and Yang (2007). 3

2

Monte Carlo (MCMC) method.7 An over-identifying set of restrictions allows for a rather quick convergence of the model with risk-neutral dynamics. This output is then used to select the market price of risk specification mostly preferred by the data. In line with Duffee (2010) and Joslin et al. (2010), both level and slope shocks seem to be priced in yields. Once the model is estimated, every one-day change of the fitted yields, subsequent to an FOMC decision, is decomposed into expected future short-rate change and the term premia change. To my best knowledge, this might be the first study to provide a rationale for the so called “slope effect” of the US monetary policy instrument on the yield curve. The slope effect is an empirical finding that only the short-end of the yield curve moves along the target Fed Funds rate. Fleming and Piazzesi (2005) use the tick data around the FOMC meetings and document that only short rates react to monetary policy surprises. Several other studies estimate a yield curve model on daily US interest rate data and analyse the link between the monetary policy actions and the yield curve. Piazzesi (2005) uses the exact timing of FOMC meetings to estimate a continuous time model of joint distribution of bond yields and interest rate target set by the FOMC. Bauer (2011) looks at the reaction of the US term premia to the macroeconomic news and shows that the short rate expectations, and not risk premia, account for the majority of daily volatility at the long-end of curve. Goukasian and Cialenco (2006) use an extended Nelson-Siegel and Vasicek models to show that the slope of the term structure reacts significantly to the monetary policy actions. Söderlind (2010) uses daily observations of Swiss interest rates and interest rate caps8 and finds that an increase in policy rate causes on average a decrease in term premia for longer maturities. There are arguably two important results in this study. First, the estimated reactions of future short-rate expectations and term premia at the long-end of the curve to the policy rate moves of the Fed seem to produce a slope-effect, because the two negatively co-move. An anticipated Fed decision to reduce the reference rate is followed by an increase in expectations and a fall in term premia, as the markets presumably anticipate less uncertainty in future inflation outlook and price-in the Fed’s commitment to price stability in the medium and long term. On the other side, a surprise cut of the policy rate seems to be followed by a decrease in future rate expectations across maturities and an increase in the term premia immanent to the mid- and long-term yields. Secondly, the separately estimated set of parameters for policy action days unveils that the portion of long-term yields variation explained by the slope factor is much higher on policy action days. The level factor, related to the medium-term inflation expectations,9 is not the main driver of long-term yields around policy action days, 7

Recent works in yield curve estimation using Bayesian MCMC methods include Ang, Dong and Piazzesi (2007) as well as Chib and Ergashev (2009). 8 Söderlind (2010) is also the first study that uses the implied volatility from interest rate caps as an explicit factor in the model estimation. 9 See Mumtaz and Surico (2008) and Rudebusch and Wu (2008).

3

which might point to a greater role of monetary policy instrument in shaping the long-term yields around those days. This paper is organised as follows: the first section introduces the data set; the second section introduces the Gaussian yield curve model and overlays the two-set parameter idea for the non-policy days and for the FOMC rate decisions days; the third section specifies the estimation strategy and describes the Bayesian MCMC algorithm; the fourth part elaborates the results and the final part concludes.

2 2.1

The Data Yields

The dataset includes 2608 daily observations of the U.S. Dollar 6-month Libor and the 1-year to 10-year swap rates from plain vanilla fixed-to-floating swap contracts10 from January 1999 to the end of December 2008. All the yields are converted to continuously compounded assuming semi-annual compounding.11 The 6-month Libor is corrected for the consequences of the credit disruption initiated in August 2007 and lasted until the end of the sample. For this time period, I simply used the 6-month Overnight Indexed Swap (OIS) rate plus the average OIS Libor spread for the period.12 In such a way, the short rate in the sample reflects the average credit conditions throughout the sample and excludes the spike in the 6-month Libor in the days following the Lehman Brothers bankruptcy. Regarding the longer maturity rates, the treasury STRIPS are initially considered as readily available zero-coupon yields that are shown to remove some of the idiosyncratic variation in the yields of individual Treasury notes and bonds.13 Yet STRIPS, as other on-the-run Treasury securities, can “go on special” in the repo market, i.e. can be used as collateral for overnight loans at a rate bellow the general collateral rate. In other words, STRIPS can be traded at a premium respect to other securities such as the off-the-run treasuries. The implications of the Repo specialness on the bonds prices in an interest rate model can be non-trivial as noted in Duffie (1996) and Jordan and Jordan (1997). Besides repo specialness, the swap rates are appealing for other two reasons: they 10

The Libor rates are obtained from daily fixings by the British Bankers Association, while the swap rates are indicative mid-quotes averaged accross many data providers. Both series are available on Bloomberg and the fixing time for the swap rates is set to 17:00 hours New York time. 11 See Hull (2008). 12 Available also on Bloomberg from the beginning of 2001. The average OIS - Libor spread in the sample was 11 basis points. 13 See Sack (2000) for example.

4

are often regarded as “true” constant maturity yield data14 and thus not a subject to approximation error of the bootstrapping and interpolation techniques. In addition, the swap rates imply a limited credit risk premium, as in most cases only the intermediate cash-flows are exchanged. The preliminary data inspection shows that the spread15 between the swap rates and off-the-run treasuries of the corresponding maturity is minor, especially around major financial events in the sample, such as the “Dot-com” bubble burst, 9/11 and the post-Lehman credit disruption in October 2008.

2.2

Anticipated and Surprise Policy Actions

The dataset includes 82 policy meetings of the Federal Open Market Committee (FOMC) that resulted in an interest rate decision. The starting policy action was an interest rate hold delivered on the 3rd of February 1999 due to a loosing growth momentum and contained inflationary pressures.16 The last decision in the sample was made on 16th December 2008 in the aftermath of Lehman Brothers collapse, when the Fed decided to cut the reference rate by 75 basis points to the target range 0 - 1/4 percent. Out of 82 FOMC meetings, 11 decisions are labelled as “surprise changes” of the Federal Funds target rate. Following Kuttner (2001), I first construct a measure of the “surprise element” in Federal Funds target changes using the Federal Funds futures data from Chicago Mercantile Exchange (CME). Secondly, I characterise different policy actions as expected or unexpected. In the construction of the policy surprise indicator, the change in the Fed target rate implied by the current-month futures contract on (monthly) average Federal Funds rate is considered. For a FOMC decision that took place at day d of the month m, the unexpected change in the policy rate, scaled up by a factor that accounts for the number of days in the month affected by the change, is calculated as follows:

∆ i unexpected =

D (F m,d − F m,d −1 ) D−d

(1)

where D is the number of days in the current month and F m,d is the Fed Funds rate implied by the current-month futures contract value. If a policy decision was widely expected, the above change should be close to zero. In order to minimise the effect of month-end noise, I calculate an unscaled change for any decisions that 14

Dai and Singleton (2000). As much as the change in the spread. 16 See Fed Greenbook, February 1999, Part 1. 15

5

came in place in the last 10 calendar days of any month.17 Results are shown in Table 1 in the Appendix. Once constructed the surprise index, a “surprise change” is considered to be any difference calculated in (1) that exceeds a two thirds of the usual 25 basis points move in any direction, namely under -16 and above +16 basis points.18 In such a way, there are 11 FOMC decisions considered as “surprise moves“. Specifically, out of 23 FOMC decisions opting for an interest rate hike, only one seem to have been a surprise hike. It was delivered on 22nd of March 2005 in a series of rate hikes lasting from June 2004 to June 2006. I consider as unexpected holds the FOMC decisions on 19th of March 2002 and 18th of September 2008. The remaining 8 policy actions are considered as surprise target rate cuts and are equally spaced between the dot-com crisis at the beginning of 2000’s and the recent financial crisis. Roughly half of these are delivered after an unscheduled meeting of the FOMC.19 There were overall 23 FOMC decisions to cut the target rate. Finally, a brief comment regarding the FOMC decision on the 16th of December 2008, when the policy rate reached the target range 0 - 25 basis points, is warranted. The scaled one day change of the Fed futures was minus 35 basis points, which means that the futures market anticipated a policy rate cut of a smaller portion i.e. the futures market seem to have been “surprised“ only by the magnitude of the rate cut. In addition to this, Taylor (2010) show that, considering the size of the reserve balances of depository institutions at Federal Reserve banks at that time, the amount of monetary easing seem to have “front-runned” the effective Federal funds rate. For this reason, I do not consider this last rate decision in the sample as a surprise move.

2.3 2.3.1

Implied Volatility from Interest Rate Caps Mechanics

An interest rate cap is a financial derivative that provides protection to the buyer, who has borrowed funds on a floating rate basis, against the risk that interests rate rise in the future. The buyer receives a positive payoff whenever an underlying interest rate (specifically the 3m Libor in our dataset) rises above an agreed level - known as the cap rate. 17

Kuttner (2001) proposes 3 days for the same purpose. 10 days are chosen to bring the measure closer to what previous studies using the tick-by-tick data produced, for example see Fleming and Piazzesi (2005). 18 The two-thirds threshold was chosen as an arguably reasonable portion of the usual policy move, above which the move might be considered as a surprise one. Altering the threshold to 15 or 17 basis points does not change the main result. 19 Namely, on the 3rd of January 2001, the 18th of April 2001, the 17th of September 2001, the 22nd of January 2008 and the 8th of October 2008.

6

One interest rate cap is essentially a portfolio of call options, so called caplets, on all the future short-term rates until the expiry of the cap. For instance, an interest rate cap expiring in 1 year consists of three options on 3-month Libor that start 3, 6 and 9 months ahead, respectively. Each of the options pays a positive difference between the 3-month Libor and the cap rate at every single expiry, i.e. 6, 9 and 12 months ahead, respectively. As different interest rate cap market participants use different option pricing models to determine the buying/selling price of a cap, the premium for an interest rate cap is by convention quoted in units of implied volatility. Specifically, it is the volatility of the underlying short-term rate “priced“ by the market maker and implied by an option pricing model (usually a version of the Black-Scholes-Merton model) used by the market maker. Using again the example of a 1-year interest rate cap, the implied volatility could be thought of as the volatility of the average 3-month Libor one year ahead. Finally, the option data at hand consist of six time series of implied volatilities across the cap maturities of 1-year to 4-year, 7-year and 10-year starting from January 1999.20 Any data point represents a composite quote stemming from many data-providers (market makers) and averaged across cap rates. The Figure 1 shows the joint dynamics of term spread and 10-year implied volatility. As it can be noticed, a relatively high term spread (in levels) seem to be related to a higher implied volatility and vice versa, with correlation coefficient at 0-lag around 0.74. FIGURE 1 ABOUT HERE

2.3.2

The Third Factor Candidate

From what previously said, implied volatility in interest rate caps can be partially considered as uncertainty in future dynamics of the short-term interest rates, over the time-span of the cap maturity.21 Considering the findings in Hördahl and Vestin (2003) and Söderlind and Svensson (1997), the second moment of future short-term rates distribution seem to be an important indicator of market expectations concerning the future monetary policy. The third state variable in the model is the demeaned series of 10 year interest rate cap implied volatility22 and this 20 The data source is Bloomberg and the series is available from December 1995 but the contribution frequency prior to the beginning of 1999 is poor. It also publishes implied volatilities across both maturities and cap rates starting from 2002. As in the case of swap rates, the fixing time is 17:00 hours New York time. 21 Given that the single option caps are not “delta-neutral“ strategies, this implied volatility might be considered as a proxy for the pure uncertainty of the underlying. In other words, the implied volatility quotes from the market makers also include the so called “directional risk“ i.e. the risk that the underlying yield moves upwards or downwards, see Cieslak and Povala (2011). 22 To my best knowledge, the first study that uses implied volatilities from interest rate caps in a yield curve model estimation is Söderlind (2010).

7

section shows that this volatility explains a non-trivial portion of the time-varying “macroeconomic uncertainty“. As the macroeconomic indicators reflecting the final objectives of the Fed,23 I consider the monthly changes in unemployment rate, headline inflation (all-items CPI), industrial production (all-sectors index) and the 10-year Treasury yield. The proxies for the macroeconomic uncertainty of the four indicators plus the federal funds target rate are constructed as squared demeaned monthly changes of the series.24 Next, I regress monthly changes and squared monthly changes in the aforementioned macroeconomic series, namely Fed funds rate, unemployment rate, headline inflation, industrial production and the 10-year yield, on single implied volatilities across maturities. The results of these univariate regressions are shown in Table 2. TABLE 2 ABOUT HERE As the upper panel of the Table 2 illustrates, implied volatilities explain a nontrivial part of changes in the federal funds target rate and the 10-year yield, whereas the proportion of the variation explained increase with the cap maturity. Particularly interesting is the last row of the panel, showing that the implied volatility on the 10-year implied volatility explains roughly 40 percent of the variation in the 10-year yield. The negative sign of the slope coefficient stems from the negative correlation between the level shock in the 10-year yields and the implied volatilities.25 Thus, the effect of the implied volatilities on the yield curve seems to be operating through the negative relation with the level factor. Intuitively, as both economic prospects and credit market conditions are perceived to deteriorate, the money market participants bid up the option caps’ premiums and thus increase the implied volatilities across maturities. At the same time, the central bank eases the policy rate in an attempt to reduce the overall level of the yield curve and the policy rate decision is priced-in on the rates market. The lower panel of the Table 2 shows close relationship of the cap implied volatilities and the uncertainty proxies. It seems that there is a statistically significant link between the 1-year implied volatility and the proxy for unemployment uncertainty, on one side. On the other, the implied volatility from the 10-year implied vola explains a statistically significant portion of the variation in CPI and IP volatilities, while remaining a significant independent variable in regressions against the Fed funds rate (7.1 percent of variation in terms of R-squared) and the 23 The Federal Reserve has a dual mandate of promoting “effectively the goals of maximum employment, stable prices and moderate long-term interest rates“ (Federal Reserve Act, Section 2a. Monetary Policy Objectives). 24 Alternatively, one could use the rolling-window standard deviation as in Blanchard and Simon (2001) or dispersion measures from economic surveys data, see Boero, Smith and Wallis (2008). 25 This is broadly in line with previously reported results in Heidari and Wu (2010) and Almeida, Graveline and Joslin (2011).

8

10-year yield (7.8 percent of variation). It might be perhaps important to emphasize that all the six implied volatility measures explain a statistically significant proportion of the variation in the proxy for inflation uncertainty. Finally, Figure 2 illustrates the link, when the proxies are rolling-window standard deviations of the series, as in Blanchard and Simon (2001). The visual inspection shows some relation of the 10-year implied volatility to unemployment-, industrial production- and inflation standard deviations, whereas the level crosscorrelations are 0.53, -0.39 and -0.04, respectivelly. FIGURE 2 ABOUT HERE

3

The Model

This section introduces the general asset pricing relation, defines the pricing kernel and specifies the Gaussian diffusion process of the underlying state factors with constant volatility. In addition, it derives the bond prices and overlays a simple two-parameter-set design.

3.1

General Setting and State Dynamics

Let us start with the general asset pricing equation26 under physical probability measure P: £ ¤ P n,t = E t M t+1 P n−1,t+1 | I t

(2)

where P n,t is the price of an n-days to maturity zero-coupon bond in time t, M t+1 is the stochastic discount factor and I t is the agents’ current information set. In a risk-neutral world where investors request no risk compensation, the price of the bond P n,t equals: ¤ Q£ P n,t = E t exp(− y1,t )P n−1,t+1 | I t

(3)

where Q is the risk-neutral probability measure and y1,t is the short-term interest rate. The no-arbitrage argument assures that the two above prices are equal, as there exists an equivalent martingale measure Q according to which (3) holds27 and: 26 27

See Campbell, Lo and MacKinlay (1997) See Harrison and Kreps (1979).

9

exp(− y1,t ) = E t [ M t+1 | I t ] = exp(− y1,t )E t [( d Q/ d P) t+1 | I t ]

(4) where d Q/ d P is the Radon-Nykodim derivative28 following a log-normal process: 1 ( d Q/ d P) t+1 = exp − (λ t )0 λ t − (λ t )0 ε t+1 2 µ



(5)

where λ t is the market price of risk associated with the sources of uncertainty ε t 29 . Following Duffee (2002), it is “essentially affine” in the risk factors X t as follows: λ t = λ0 + λ1 X t

(6)

where λ0 and λ1 produce constant prices of risk if λ0 6= 0 and λ1 = 0 or time-varying prices of risk if λ0 6= 0 and λ1 6= 0. The equations (4) to (6) define the pricing kernel. Another fundamental building block of the Gaussian term structure model is the state variable X t which follows a discrete version of constant volatility OrnsteinUhlenbeck process30 . Under the physical probability measure P, the process is:

X t+1 = ( I − Ψ)µ + Ψ X t + Σε t+1

(7)

where the first term on the right-hand side is a vector of factors’ means, Ψ is the VAR matrix, Σ is the covariance matrix, and ε t is an IID N (0, 1).

3.2

Bond Prices

Following Duffie and Kan (1996), the one-period interest rate is an affine function of risk factors X t as:

y1,t = A 1 + B1 X t 28

See Dai et al. (2007). See Ang and Bekaert (2002) and Ang and Piazzesi (2003). 30 See Phillips (1972). 29

10

where the coefficient A 1 corresponds to the average one-period rate in the sample and B1 is a vector of loadings of the risk factors on y1,t . Assuming joint lognormality of bond prices and the pricing kernel in equation (2), the n-periods to maturity nominal bond price is an affine function of the state variables:

© ª 0 P n,t = exp −An,t − Bn,t Xt

(8)

where the coefficients A and B are computed using the usual recursive formula.31 Nonetheless, the two coefficients are time dependent, as the agents can choose between two different sets of parameters to price a n-days to maturity bond P n,t in t: ´ " ³ # j 0 j Ψ − Σ j λ1 B n−1 + B1 Bn,t = ι t ¡ ¢0 Ψ − Σk λ1k B kn−1 + B1

and " An,t = ι t

³ ´ # ,j j j j 0 j A n−1 + ( I − Ψ)µ − Σ j λ0 B n−1 + 21 B n−1 Σ j (Σ j )0 B n−1 + A 1 ¡ ¢0 ,k A kn−1 + ( I − Ψ)µ − Σk λ0k B kn−1 + 12 B n−1 Σk (Σk )0 B kn−1 + A 1

where ι t is a 1 × 2 indicator vector, picking alternatively the j or the k parameter set in every t. It is assumed that agents use the former parameter set to price the bond on a non-policy day and letter in all the days of the FOMC policy rate decision:

j = {non-policy day}

and

k = {FOMC statement day}

As it can be noticed, the parameters in the VAR matrix as well as the parameters µ, A 1 and B1 are common in the two parameter sets. As the time interval is daily and the second parameter set is only used to price the bonds on policy action days, there is arguably no reason to believe that the persistency in the factors dynamics nor the factors mean and the average yield curve level alter significantly on an average policy action day. Finally, given that an agent chooses between the two “models” with separate dynamics, the underlying assumption is that a particular model explains the yield dynamics until infinity. In other words, the persistency of a single “regime” is equal to unity. Yet at a cost of this approximation, there is possibly a multiple benefit. We are able to measure separately the volatilities of the underlying state 31

See for example Campbell et al. (1997) or Cochrane and Piazzesi (2009).

11

variables on the usual days and on the policy days. It is possible to attribute the variation in different yields to the variation in single state variables on the policy action days. Above all, one can allow for different compensations for shocks to the level and to the slope of the term structure on different days.

3.3

No-arbitrage Argument

To assure that the no-arbitrage argument holds, I assume that the state variable X t under the risk-neutral measure Q follows: Q

Q

X t+1 = ( I − Ψ)µQs + ΨQs X t + Σs ε t+1

(9)

where s = { j, k}. Note that the VAR matrix ΨQs is now different in the two sets. Given equations (3) and (4), this implies that the usual no-arbitrage restrictions on the parameters describing the physical and the risk neutral measure in this case are: µQs = µ t − Σst λ0s

and s ΨQs = Ψ t − Σst λ1t

where parameters on the right hand side are estimated while the parameters µQs and ΨQs are derived parameters. This might be important, since in the regimeswitching literature and the stochastic regime-switching setting, Bansal and Zhou (2002), for instance, show that the regime-dependent ΨQs parameter calls for a log-linearisation of the pricing kernel to get the close form solution for the recursive pricing formula. In another important study of Dai et al. (2007), the authors impose a restriction on Q distribution so that ΨQ is regime-independent. In a tworegime switching model, they estimate one of the ΨPt s parameters and derive the other VAR parameter under P and the one-and-only under Q.

3.4

Forward Term Premia

In this section, the model-implied forward term premia are derived. The reason for focusing on this particular definition of the risk premium32 is that most of the 32

The term premium or the risk premium can be equivalently defined as a yield risk premium, a forward risk premium and a return risk premium, see Cochrane and Piazzesi (2009).

12

studies of the U.S. term premia report the forward term premia.33 According to the expectation theory of the term structure34 , an “ n − m period” forward rate n periods ahead, f wd mn,t , is equal to the expected future short rate plus the term premium:

£ ¤ f wd mn,t = E t y1,t+n−1 + f TP n

(10)

where f wd mn,t is the continuously compounded forward rate:

f wd mn,t =

n m yn,t − ym,t n−m n−m

(11)

Once I fit the yield curve and obtain yi,t for every i = {6 m, 1Y , ...10Y } and every t = [1, T ], the forward rates are calculated using (11) and the apropriate expectations part is subtracted. The expected future short term rate between 6-month and 1-year time from t:

E t [ y6m,t+360 ] = A 1 + B01 E t [ X t+180 ] £ ¤ = A 1 + B01 ( I − Ψ)µ( I + Ψ + Ψ2 + ... + Ψ180 ) + Ψ180 X t is subtracted from the 6-month to 1-year forward rate and:

h i E t [ y1Y ,t+360n ] = A 1 + B01 ( I − Ψ)µ( I + Ψ + Ψ2 + ... + Ψ360(n−1) ) + Ψ360(n−1) X t (12)

from any 12-month forward rate starting in n = 2, 3...N years. Further decomposition of the forward term premia to very state variables and parameters of the model can be found in Hördahl, Tristani and Vestin (2006).

4

Estimation

In this section the likelihood function used to construct the joint posterior of parameters and data is derived. In addition, the model is estimated with a simple version the Bayesian Markov-Chain Monte-Carlo (MCMC) method and this section provides the general idea, rationale and the description of the algorithm. 33

For an instructive review, see for instance Rudebusch et al. (2006). See Campbell and Shiller (1991) for an insightful discussion of the expectation theory of the term structure. 34

13

4.1

Likelihood Function

Following the ideas in Chen and Scott (1993), the 6-month and the 10-year yields are set to be observable and the rest to be measured with error. Let yo,t be a vector of stacked observable yields i.e. yields perfectly priced by the model: ¡ ¢0 yo,t = A so + B so X o,t

where A so is a 2 × 1 vector, B so a 2 × 2 matrix of factor loadings and X o,t are the two “latent” factors. The factors are obtained by inverting the observed yields as: ¡ ¢−1 ¡ ¢ X o,t = B so yo,t − A so

To define the X o,t+1 and thus the next day yield yo,t+1 , I follow the timing convention in Dai et al. (2007), according to which the conditional probability of X t+1 satisfies:

f P ( X t+1 | X t ; s t = j ; s t+1 = k) = f P ( X t+1 | X t ; s t = j )

(13)

The formulation differs from the original setting proposed in Hamilton (1989) where:

f P ( X t+1 | X t ; s t = j ; s t+1 = k) = f P ( X t+1 | X t ; s t+1 = k) and the authors introduce it for the reasons of comparability of their model with the continuous-time yield curve models.35 The convention proves to be useful in this setting and it can be shown that the yield yo,t+1 reads: ³ ´0 j c yo,t+1 = µ co + Ψ o yo,t + B ko Σ o ε o,t+1

where: ³ ´0 ³ ´−1 j j Ao µ co = A ko + B ko ( I − Ψ)µ j − B ko Ψ B o

and 35

See Dai et al. (2007).

14

³ ´0 ³ ´−1 co = B k Ψ B oj Ψ o

Consequently, the conditional probability density function of yo,t+1 is:

e

pdf( yo,t+1 | yo,t , s t = j, s t+1 = k) =

³ ´³¡ ¢ ´−1 ³ ´0 co yo,t B k 0 Σ oj (Σ oj )0 B k co yo,t − 12 yo,t+1 −µ co −Ψ yo,t+1 −µ co −Ψ o o

r ¯ ¯¡ ¢ 0 j ¯ ¯ j (2π)T ¯ B ko Σ o (Σ o )0 B ko ¯

(14) Finally, let the yu,t be a vector of the rest N − 2 yields measured with error: ¡ ¢0 yu,t = A su + B su X o,t + ξst ¡ ¢ where ξst is distributed as i.i.d. N 0, ωs2 I . The normally distributed errors are assumed to have the same variance ω j across the unobservable yields, but which is different in the two parameter sets.36 It is also assumed that the errors are uncorrelated cross-section and thus the covariance matrix is a diagonal matrix. Inverting again the relation between yo,t+1 and latent factors we have: d yu,t+1 = µ cu + Ψ u yo,t+1 + ξ t+1

where: ³ ´0 ³ ´−1 µ cu = A ku − B ku B ko A ko

and

³ ´0 ³ ´−1 d Ψ = B ku B ko u

The conditional density of yu,t+1 is equal to:

pdf( yu,t+1 | yo,t+1 , s t = j, s t+1 = k) =

e

³ ´¡ ´´0 ¢ ³³ k k 0 −1 d d − 12 yu,t+1 −µ cu −Ψ yu,t+1 −µ cu −Ψ u yo,t+1 ω I(ω I) u yo,t+1

q

¡ ¢0 (2π)T |ωk I ωk I |

(15) 36 In this way, one can question weather cross-sectional pricing errors have different distributional characteristics on the non-policy versus policy action days.

15

As already mentioned, the state variables are two latent factors and the 10-year implied volatility as the third explicit factor. The likelihood function in logs thus can be written as: lnL

= +

TX −1 t=0 TX −1

¡ ¢ lnpdf yo,t+1 , IVt+1 | yo,t , IVt , s t = j, s t+1 = k ¡ ¢ lnpdf yu,t+1 | yo,t+1 , IVt+1 , s t = j, s t+1 = k

t=0

(16)

4.2

Bayesian Inference

The yield curve implied by the model is a complicated non-linear function of the underlying parameters. As this non-linearity tend to produce a multi-modal likelihood function37 , fitting a yield curve model with a standard maximum likelihood estimation is a daunting task. Bayesian Markov Chain Monte Carlo (MCMC) method seem to be a powerful alternative, providing both efficiency and tractability.

4.2.1

Setting

Let Θ be a vector that collects all the parameters of the model: n o j j Θ = µ, Ψ, Σ j , Σk , λ0 , λ0k , λ1 , λ1k , ω j , ωk , A 1 , B1

The key idea behind Bayesian estimation is to consider the vector as a multivariate random variable, and use the Bayes’ rule to “learn” about the variable given the data:

p (Θ | data) =

p (data | Θ) p (Θ) p (data)

(17)

where p (Θ | data) is the posterior density of Θ, p (data | Θ) is the likelihood function and p (Θ) denotes the prior density of the parameters. The term p (data) is known as “normalising constant“ and it is independent of Θ.38 Consequently, the rule in (17) can be re-written as: 37 38

See Chib and Ergashev (2009). R In particular: p (data) = p(data | Θ) p(Θ) d Θ. See Koop (2003).

16

ln p (Θ | data) ∝ lnL + ln p (Θ) where lnL is the logarithm of the likelihood function defined in equation (16).

4.2.2

Priors

In the estimation, the priors p (Θ) are set to be non-informative or “flat”, so that the posterior density of the model parameters is drawn with equal probability from the pre-defined support interval. Alternatively, one could derive the prior distributions for parameters µ, Ψ and Σs , given the normality assumption of the state VAR process,39 and for ωs given the assumption of the Gaussian measurement error.40 Chib (2001) propose a scaled beta distribution as an alternative to the uniform distribution. Nevertheless, I choose not to impose lower (or higher) probability areas from which the candidate values of parameters are drawn. In such a way, the estimation is almost completely data-driven and proves to be computationally efficient. The parameters’ suport intervals are specified by following the no-arbitrage condition and previous studies. In particular, the eigenvalues of the VAR matrix are set to be positive and less than one and the volatility parameters on the diagonal of Σs are set to be non-negative. The lower bound of the parameters in λ0s vector and λ1s matrix are set as in Chib and Ergashev (2009).

4.2.3

Markov Chain Monte Carlo

I use a simple version of “Metropolis within Gibbs” algorithm41 to draw the parameters from their posterior densities. ¡ The parameter candidates are drawn ¢ from continuous uniform distributions U aΘ , bΘ where the lower and the upper boundaries aΘ and bΘ for each parameter in Θ are specified in Table 4. The algorithm can be described in several steps: Step 1: Set the initial values of parameters Θ0 . Two Markov chains with different starting values are set up. The initial values for the VAR parameters in first chain are obtained from OLS and the data descriptive statistics. The starting values of the market price of risk parameters and of the second chain are chosen arbitrarily.42 39

For instance, see Ang et al. (2007). See for example Mikkelsen (2001). 41 See for example Gilks (1996), Koop (2003) and Lynch (2007). 42 For example, the volatility parameters’ starting values are set to be 3 times larger in the second chain, the Ψ matrix parameters are set to 0.8 and the B1 parameter is set to 0.5 and -0.5 in the 40

17

¡ ¢ Step 2: Draw a candidate log-posterior density ln p Θ∗ | Θmc−1 , data conditional on previously drawn parameters’ values Θmc−1 . The number mc denotes current iteration. The draws are performed separately for every parameter in Θ. For instance, a proposal for the first element in the vector Θ is generated by the following Markov chain:

θ1∗ = θ1mc−1 + ν1U1

where ν• is a scaling factor and U• is an uniformly distributed random number from interval [-1,1]. For the parameters in Ψ, the scale factor ν• is initialised to 0.01, for diagonal elements of Σ matrix and the market price of risk parameters to 0.1, and for the ω parameter to 0.0000143 . The scaling factor is then automatically updated after every 5,000 sweeps44 to obtain the acceptance ratio of approximately 0.5. Step 3: For every parameter in Θ∗ , calculate the difference between the posterior density with the candidate value and the posterior density with the previously drawn parameter value, keeping the other parameter values unchanged. Using again the first element in Θ as an example, the difference reads:

¢ ¡ mc−1 } | Θmc−1 , data α = ln p {θ1∗ , θ2mc−1 , ..., θK ¡ ¢ mc−1 } | Θmc−1 , data − ln p {θ1mc−1 , θ2mc−1 , ..., θK

(18)

Step 4: Draw a random number u ∼ U (0, 1) and accept the single parameter candidate from Step 2, whenever the following holds for the difference in Step 3:

min(0, α) > log( u) Step 5: Repeat the Steps 2 to 5 until the joint posterior density of parameters converge in distribution. The algorithm is ran 100,000 times and the first 40,000 are discarded as the burnin period. The two Markov chains with different starting values for both joint first and the second Markov chain, respectivelly. 43 Proposed in Ang et al. (2007) so that it roughly corresponds to a 30 basis points bid-ask spread on Treasuries. An average spread on the OTC plain vanilla swap market might be similar. See also Skarr (2010). 44 The algorithm is ran for 100,000 times. The scaling factor is updated starting from the 10,000th iteration to the 40,000th iteration.

18

posterior and the single parameters’ posteriors converge to literally the same posterior distributions. Before estimating the entire model, the proposed parametrisation is used to estimate the risk neutral specification. The model under Q converges even quicker45 and thus the algorithm is ran for 50,000 times and the first 20,000 are discarded as burn-in.46

4.3 4.3.1

Econometric Identification Parameters

Solid identification of parameters is an essential part of dynamic term structure models estimation. To this end, I use the standard identification scheme from Dai and Singleton (2000) with a couple of alterations. Regarding the diffusion process for X t , the VAR matrix is set to be upper diagonal47 and set the persistency coefficient between the two latent factors to be zero. Preliminary estimation showed that the Ψ11 parameter, corresponding to the autoregression coefficient of the level factor, is near one. I thus simply set it to 1,48 which is in line with the near co-integration assumption from previous studies.49 Instead of normalising the Σs matrix to unit matrix, I allow for different values of diagonal elements in Σs in two different parameter sets. Finally, the µ vector is normalised to zero. When calculating the parameters of the pricing equation, I impose the usual boundary condition A 0 = B0 = 0. A 1 is normalised to average 6-month Libor in the sample50 , namely 3.64 percent. Assuming that only latent factors move the short-term interest rate, the vector B1 is normalised to [1 1 0]’.51 As the third factor has no loading on the short rate, the last row of vector λ0s and matrix λ1s is zero. Before estimating the entire model, the proposed identification scheme is used to estimate the risk neutral specification. The preliminary analysis point to a rather quick convergence of the parameters defining the Q measure. The mean values of parameters’ posterior distributions are used as starting values in the choice of the market price of risk specification, explained in the following section. 45

On average after 7,000 sweeps, between the non-ZLB sample and the ZLB sample. The scaling factor is also automatically adapted until the 20,000th iteration. 47 Choosing between the lower/upper diagonal Ψ and lower/upper diagonal var-cov matrix Σ is equivallent, see Dai and Singleton (2000), Appendix C. 48 As in Diebold and Li (2006), Söderlind (2010) and Bauer (2011). 49 See for instance Giese (2008) and Jardet, Monfort and Pegoraro (2011). 50 Following Favero, Niu and Sala (2007). 51 Alternativelly, one could normalise the covariance matrix of the diffusion process to a unity matrix and estimate the elements of B1 , see for instance Ang and Piazzesi (2003). 46

19

4.3.2

Market Price of Risk Specification

In the second part of the preliminary analysis, all the parameters of the matrix λ1s for both non-policy and policy days are estimated and the algorithm did not s s s converge after 50,000 sweeps. The parameters λ11 and λ21 on one side, and λ12 s and λ22 on the other side, seem not to be separately identified. To see this, let us consider for example the vector B2s . It is equal to the product:  s s s s Ψ11 − Σ11 λ11 − Σ22 λ21 s s s s  λ12 + P22 − Σ22 λ22 (Ψ − Σs λ1s )0 B1 =  −Σ11 s s s s P13 − Σ11 λ13 + P23 − Σ22 λ23 

(19)

s and Arbitrage restrictions might not help as well, a large positive value of λ11 s s s a large negative value λ21 are interchangeable. The same goes for λ12 and λ21 elements. That said, out of possible 26 combinations to estimate, I isolate 35 following this logic, and estimate every specification running the algorithm 10,000 times. The specifications are then sorted according to the Bayesian Information Criterion and the results are reported in Table 3.

TABLE 3 ABOUT HERE

As it can be noticed, the first 23 specifications have similar values of the criterion and thus produce similar residual sum of squared pricing errors. As there is no obviously preferred one, I look at how many times a parameter was in the specification and thus sum the first 23 rows of the Table 3. The market price of risk specification preferred by the data is:    s s 0 0 λ1,13 λ0,1   s   s λst =  λ0,2  +  0 0 λ1,23  Xt 0 0 0 0 

(20)

Consequently, both “level” and “slope” risks seem to be priced-in, i.e. the expected returns seem to be earned in compensation not only to the “level” shock52 but also in compensation to the “slope“ shock.53 52 53

Similarly to Cochrane and Piazzesi (2009). As in Duffee (2010) and Joslin et al. (2010).

20

5

Results

5.1

Parameters and State Variables

Table 4 presents the estimated transition matrix Ψ in the upper panel and other parameters in the lower panel. As already mentioned, the persistency of the level factor in the daily data is near the unit-root and for that reason it is set to 1 and excluded from estimation. Removing the stationarity restriction on the X l evel do not influence the overall estimation, as the parameter posterior distributions with or without the restriction are literally the same.54 Given the upper panel of Figure 1, this might be a reasonable assumption. TABLE 4 ABOUT HERE The estimates of the Σs matrix point to a relative increase in volatility of the slope factor and a sizeable increase in volatility of the explicit factor on the days of the policy rate decision. The volatility increase around FOMC meeting has been already documented by a wide literature in announcement effect. Estimated Ψ and Σs matrices reproduce the state variables dynamic and I plot the level and the slope factor in the upper panel of Figure 3. The correlation between factors is -0.07 from where it might be inferred that they are “telling“ separate stories about yields’ dynamics. FIGURE 3 ABOUT HERE Finally, the measurement error volatility ωs is higher for the policy action days, which reveals that cross-sectional pricing errors seem to be systematically more volatile on the days of the FOMC decisions. In addition, all the parameters of the MPR specification are negative and not too different between the parameter sets. One might notice though, that the compensation for the “level“ risk around policy days seem to be somewhat higher than on average days.

5.2 5.2.1

Separating non-policy and policy days Pricing Performance

By construction, the 6-month and the 10-year yields are explained perfectly by the model. I therefore look at the cross-sectional fit of the yields and estimate separately the two-parameter-set model and the single-set model, using the entire data sample. Table 5 reports the mean absolute pricing errors. 54

This is also a by-product of the market price of risk specification, see equations (19) and (20).

21

TABLE 5 ABOUT HERE As it can be noticed, the pricing errors of the alternative models are comparable. Both models fit the data increasingly better going from the short- to the long-end of the curve. A slightly better performance of the two-set model around policy action days might be explained by the fact that, the higher volatility of the level and curvature factor plus a more negative market price of risk coefficient for the slope factor translate into a more “pronounced“ implied-volatility-factor loading at the long-end.

5.2.2

Relative importance of factors around policy days

Estimating a separate set of parameters for the policy action days allows for state variables to have different loadings on the yield curve. In addition, we can have an insight into the portion of yield variance explained by the factors on average days and policy action days, separately. On the lower panel of the Figure 3, the explicit factor loadings for two different sets of parameters are illustrated. The loadings are identical for the 6-months yield by construction. Moving towards the long-end, the implied volatility factor has a stronger impact on the yield curve, both on average days in the sample and on the days of the FOMC decisions. Given the shape of the factor loading, the 10-year implied volatility might be considered as the “curvature factor“.55 The negative loadings indicate that an increased uncertainty in future short-rates is followed by a decrease in long-term rates. As the level factor is the main driver of the longer rates overall, there is a negative correlation between the level factor and the implied volatility as well. TABLE 6 ABOUT HERE Table 6 reports the variance decomposition of yields separately for the non-policy days and for the FOMC decisions days. The main driver of medium- and longterm yields on an average day in the sample is the level factor, explaining a half of the variation in medium-term yields and two-thirds of the 10-year yields. On policy action days, diversely, the slope factor explain roughly as much variation in the long-term yields as the level factor and more than a half of the variation in the medium-term yields. This result is driven by the higher slope and curvature factors’ volatilities, estimated in the policy-day set. Accordingly, the explicit factor also becomes more important around policy action days. The variation in the long yields explained by the third factor equals 4.6 percent on an average day and slightly above 10 percent on a day of the FOMC rate call. 55

Having a rather similar shape to the third unobserved factor in Ang and Piazzesi (2003).

22

5.3

Estimated effects of the Policy Rate Announcements

Once the yield curve is fitted, the instantaneous forward rates are calculated and decomposed to expected future short-term rates and the forward term premia. To assure that the estimated term premia behaves similarly to what previous studies have shown, I look at the change of estimated expected rates and the forward term premia during the period June 2004 - February 2005. In the mentioned period, the Fed increased the policy rate by 150 basis points in 6 FOMC meetings. The 6-month Libor increased by roughly 100 basis points while the 10Y swap rate decreased by 64 basis points. This yield curve movement was considered at the time as a “conundrum“ and the subsequent studies showed that, during that time, the future expectations of the short-term rate rose in a parallel fashion across maturities while the long-term premia significantly fell. The term premia and the short-rate expectations reported here are broadly consistent with these findings.56 TABLE 7, FIGURE 5 AND 6 ABOUT HERE That said, the Table 7 and Figures 5 and 6 report the estimated one-day reactions of the term premia to different policy actions,57 while Table 8 and Figure 7 show the estimated one-day average change of the expected future short-term rates after a FOMC decision. The main result, independent from the market price of risk specification58 or the model used to fit the yields, is two-fold. First, the longer term premia seem to fall after an anticipated decision of the FOMC to cut the Fed funds rate. When a move of the Fed is widely anticipated, perceived uncertainty about the future outlook for the economy seem to fall, or at least to abate. The estimated reduction in premia amounts from 3.5 to 11 basis points. At the same time, the long-term expectations of the future short-rate rise, as the markets might expect that the Fed eventually increases the reference rate to fight potential inflationary pressures in the future. The average increase in the short-rate expectations is 2 basis points.59 Secondly, the average response of the forward term premia to an unexpected expansionary move by the Fed is positive and it is followed by a decrease in shortrate expectations. While the forward premia reaction is statistically significant, 56

This result is not reported. The 90 percent credible interval is calculated using the posterior parameters’ distributions. As the parameters enter the model in a complex and non-linear way, I use every 100-th set of parameters along the first Markov chain and after burn-in, to fit the yield curve and then calculate the reaction. 58 Very similar results are obtained for all the 23 specifications reported in Table 3. 59 The reported changes correspond to an average anticipated target Fed funds rate cut of 40 basis points (in 15 occasions). The estimated reduction in premia to an anticipated 25 basis-point cut is thus 2.2 to 6.9, and the estimated increase in expectations 1.3 basis points. 57

23

both one-set and two-set models estimate a downward shift of the short-rate expectations which are not statistically significant, with p-values of the t-statistic slightly above the level of significance of 0.10.60 There is an average 7.4 to 9.3 basis points increase in term premia implicit in the one-year forward rate 3 years in advance, an average 9.8 to 12.2 basis points increase in 5 years and an average 10.8 to 11.5 in 7 years in advance one-year premia, following a surprise interest rate cut of the Fed.61 TABLE 8 AND FIGURE 7 ABOUT HERE The estimated negative co-movement of short-rate expectations and term premia after contractionary policy actions seem to be responsible for the “missing” reaction of the long-term yields to the FOMC decisions after the surprise policy actions found in Fleming and Piazzesi (2005). As it can be noticed, the long-term yields reaction to other decisions is not statistically significant. If the high-frequency identification implemented in this study is arguably one of few reasonable ways to assess the effect of monetary policy on yield curve, this result provides rationale of why interest rate policy in the US might be a slope, and not the level, “tool” in shaping the US yield curve.

6

Conclusion

This study estimates a no-arbitrage Gaussian term structure model which includes an additional set of parameters to capture the yield curve dynamics around the days of the FOMC policy rate decisions, in period from January 1999 - December 2008. Given the potential multi-modality of the likelihood function, a version of Metropolis algorithm is used to estimate the model, while imposing economically meaningful restrictions on paramers’ values and the support interval. No explicit assumptions on the market price of risk are made, but the specification mostly preferred by the data is chosen i.e. the one in which the bond investors are compensated for both the level and the slope shock. The model output indicates that anticipated Fed decisions to cut the policy rate seem to produce a fall in term premia and an increase in future short-rate expectations implicit in long-term forward rates. A surprise cut is estimated to provoke a parallel downward shift in rate expectations and an increase in medium- and long-term forward premia across maturities. Most importantly, this result seem 60

This is given by the fact that the standard deviation of the changes in expectations is somewhat higher than the changes in term premia across the policy rate decisions. 61 Again, the reported changes correspond to an average surprise cut of the policy rate by 56 basis points (in 8 occasions). The estimated reduction in premia to a suprise 25 basis-point cut is thus 3.3 to 4.2 basis points, 4.4 to 5.4 basis points and 4.8 to 5.1, respectivelly.

24

to be independent of the market price of risk specification or weather the singleor two-set model is used in the analysis. The results presented in this study might be further reinforced in several ways. One could elaborate the reaction of the term premia to the policy rate announcements at a higher data frequency. The measured speed of reaction to a FOMC decision might provide an instantaneous insight to the policy maker into the overall market conditions in which the decision is brought or perhaps into a sort of “market assessment“ of the decision. In addition, one could analyse a longer sample of daily observations, because the operating procedures of the Fed has not been changed since the 1994. Also, one could use the output from both empirical- and affine models to decompose the forward rates into expectations and premia part and then look at their reactions. All this might be the subject of author’s future research.

References Almeida, Caio, Jeremy J. Graveline, and Scott Joslin (2011) “Do interest rate options contain information about excess returns?” Journal of Econometrics, Vol. 164, No. 1, pp. 35–44, September. Ang, Andrew and Geert Bekaert (2002) “Regime Switches in Interest Rates,” Journal of Business & Economic Statistics, Vol. 20, No. 2, pp. pp. 163–182. Ang, Andrew, Sen Dong, and Monika Piazzesi (2007) “No-Arbitrage Taylor Rules,” NBER Working Papers 13448, National Bureau of Economic Research, Inc. Ang, Andrew, Geert Bekaert, and Min Wei (2008) “The Term Structure of Real Rates and Expected Inflation,” The Journal of Finance, Vol. 63, No. 2, pp. 797– 849. Ang, Andrew and Monika Piazzesi (2003) “A no-arbitrage vector autoregression of term structure dynamics with macroeconomic and latent variables,” Journal of Monetary Economics, Vol. 50, No. 4, pp. 745–787, May. Bansal, Ravi and Hao Zhou (2002) “Term Structure of Interest Rates with Regime Shifts,” Journal of Finance, Vol. 57, No. 5, pp. 1997–2043, October. Bauer, Michael D. (2011) “Term Premia and the News,” Federal Reserve Bank of San Francisco Working Paper Series. Bernanke, Ben S. (2004) “Gradualism : remarks at an economics luncheon cosponsored by the Federal Reserve Bank of San Francisco (Seattle Branch) and the University of Washington, Seattle, Washington, May 20, 2004,” Speech.

25

Bernanke, Ben S. and Kenneth N. Kuttner (2005) “What Explains the Stock Market’s Reaction to Federal Reserve Policy?” Journal of Finance, Vol. 60, No. 3, pp. 1221–1257, 06. Blanchard, Olivier J. and John A. Simon (2001) “The Long and Large Decline in U.S. Output Volatility,” SSRN eLibrary. Boero, Gianna, Jeremy Smith, and KennethF. Wallis (2008) “Uncertainty and Disagreement in Economic Prediction: The Bank of England Survey of External Forecasters,” Economic Journal, Vol. 118, No. 530, pp. 1107–1127, 07. Calvet, Laurent E., Aldai Fisher, and Liuren Wu (2010) “Dimension - Invariant Dynamic Term Structures,” Working Paper. Campbell, John Y., Andrew Wen-Chuan Lo, and A. Craig MacKinlay (1997) “The econometrics of financial markets,” Princeton University Press, Vol. 2. Campbell, John Y and Robert J Shiller (1991) “Yield Spreads and Interest Rate Movements: A Bird’s Eye View,” Review of Economic Studies, Vol. 58, No. 3, pp. 495–514, May. Chen, Ren-Raw and Louis Scott (1993) “Multi-Factor Cox-Ingersoll-Ross Models of the Term Structure: Estimates and Tests from a Kalman Filter Model,” Journal of Real Estate Finance and Economics, Vol. 27, No. 2. Chib, Siddhartha (2001) “Markov chain Monte Carlo methods: computation and inference,” in J.J. Heckman and E.E. Leamer eds. Handbook of Econometrics, Vol. 5 of Handbook of Econometrics: Elsevier, Chap. 57, pp. 3569–3649. Chib, Siddhartha and Bakhodir Ergashev (2009) “Analysis of Multifactor Affine Yield Curve Models,” Journal of the American Statistical Association, Vol. 104, No. 488, pp. 1324–1337. Cieslak, Anna and Pavol Povala (2011) “Understanding the Term Structure of Yield Curve Volatility,” SSRN eLibrary. Cochrane, John H. and Monika Piazzesi (2009) “Decomposing the Yield Curve,” SSRN eLibrary. Dahlquist, Magnus and Henrik Hasseltoft (2011) “International Bond Risk Premia,” SSRN eLibrary. Dai, Qiang, Kenneth J. Singleton, and Wei Yang (2007) “Regime Shifts in a Dynamic Term Structure Model of U.S. Treasury Bond Yields,” Review of Financial Studies, Vol. 20, No. 5, pp. 1669–1706. Dai, Qiang and Kenneth J. Singleton (2000) “Specification Analysis of Affine Term Structure Models,” The Journal of Finance, Vol. 55, No. 5, pp. 1943–1978.

26

Diebold, Francis X. and Canlin Li (2006) “Forecasting the term structure of government bond yields,” Journal of Econometrics, Vol. 130, No. 2, pp. 337–364, February. Duffee, Greg (2010) “Sharpe ratios in term structure models,” Economics Working Paper Archive 575, The Johns Hopkins University,Department of Economics. Duffee, Gregory R. (2002) “Term Premia and Interest Rate Forecasts in Affine Models,” Journal of Finance, Vol. 57, No. 1, pp. 405–443, 02. Duffie, Darrell (1996) “Special Repo Rates,” The Journal of Finance, Vol. 51, No. 2, pp. pp. 493–526. Duffie, Darrell and Rui Kan (1996) “A Yield-Factor Model Of Interest Rates,” Mathematical Finance, Vol. 6, No. 4, pp. 379–406. Favero, Carlo A, Linlin Niu, and Luca Sala (2007) “Term Structure Forecasting: No-Arbitrage Restrictions vs Large Information Set,” CEPR Discussion Papers 6206, C.E.P.R. Discussion Papers. Fleming, Michael J. and Monika Piazzesi (2005) “Monetary Policy Tick-by-Tick,” working paper. Giese, Julia V. (2008) “Level, Slope, Curvature: Characterising the Yield Curve in a Cointegrated VAR Model,” Economics - The Open-Access, Open-Assessment E-Journal, Vol. 2, No. 28, pp. 1–20. Gilks, W.R (1996) “Full Conditional Distributions in Gilks W.R., Richardson S. and Spiegelhalter D.J. Markov Chain Monte Carlo in Practice,” Chapman and Hall. Goodfriend, Marvin (1998) “Using the term structure of interest rates for monetary policy,” Economic Quarterly, No. Sum, pp. 13–30. Goukasian, Levon and Igor Cialenco (2006) “The Reaction of Term Structure of Interest Rates to the Monetary Policy Actions,” SSRN eLibrary. Gurkaynak, Refet S., Brian Sack, and Eric Swanson (2002) “Market-based measures of monetary policy expectations.” Gurkaynak, Refet S., Brian Sack, and Jonathan H.Wright (2007) “The U.S. Treasury yield curve: 1961 to the present,” Journal of Monetary Economics, Vol. 54, No. 8, pp. 2291–2304, November. Hamilton, James D (1989) “A New Approach to the Economic Analysis of Nonstationary Time Series and the Business Cycle,” Econometrica, Vol. 57, No. 2, pp. 357–84, March. Hamilton, James D. and Jing (Cynthia) Wu (2010) “Identification and Estimation of Affine-Term-Structure Models.”

27

Harrison, J. Michael and David M. Kreps (1979) “Martingales and arbitrage in multiperiod securities markets,” Journal of Economic Theory, Vol. 20, No. 3, pp. 381–408, June. Heidari, Massoud and Liuren Wu (2010) “Market Anticipation of Fed Policy Changes and the Term Structure of Interest Rates,” Review of Finance, Vol. 14, No. 2, pp. 313–342. Hördahl, Peter, Oreste Tristani, and David Vestin (2006) “A joint econometric model of macroeconomic and term-structure dynamics,” Journal of Econometrics, Vol. 131, No. 1-2, pp. 405–444. Hördahl, Peter and David Vestin (2003) “Interpreting Implied Risk-Neutral Densities: The Role of Risk Premia,” ECB Working Paper Series. Hull, John C. (2008) Options, Futures, and Other Derivatives with Derivagem CD (7th Edition): Prentice Hall, 7th edition. Jardet, Caroline, Alain Monfort, and Fulvio Pegoraro (2011) “No-Arbitrage NearCointegrated VAR(p) Term Structure Models, Term Premia and GDP Growth,” SSRN eLibrary. Jordan, Bradford D. and Susan D. Jordan (1997) “Special Repo Rates: An Empirical Analysis,” The Journal of Finance, Vol. 52, No. 5, pp. pp. 2051–2072. Joslin, Scott, Marcel Priebsch, and Kenneth J. Singleton (2010) “Risk Premiums in Dynamic Term Structure Models with Unspanned Macro Risks.” Kim, Don H. and Kenneth J. Singleton (2010) “Term Structure Models and the Zero Bound: An Empirical Investigation of Japanese Yields,” SSRN eLibrary. Kim, Don H. and Jonathan H. Wright (2005) “An arbitrage-free three-factor term structure model and the recent behavior of long-term yields and distant-horizon forward rates,”Technical report. Koop, Gary (2003) “Bayesian Econometrics„” John Wiley and Sons Ltd. Kuttner, Kenneth N. (2001) “Monetary policy surprises and interest rates: Evidence from the Fed funds futures market,” Journal of Monetary Economics, Vol. 47, No. 3, pp. 523–544, June. Lynch, Scott M. (2007) Introduction to Applied Bayesian Statistics and Estimation for Social Scientists: Springer Science+Business Media, 7th edition. Mikkelsen, Peter (2001) “MCMC Based Estimation of Term Structure Models,” Finance Working Papers 01-7, University of Aarhus, Aarhus School of Business, Department of Business Studies. Mumtaz, Haroon and Paolo Surico (2008) “Time-Varying Yield Curve Dynamics and Monetary Policy,” Bank of England Discussion Papers, No. 23, March. 28

Phillips, P. C. B. (1972) “The Structural Estimation of a Stochastic Differential Equation System,” Econometrica, Vol. 40, No. 6, pp. pp. 1021–1041. Piazzesi, Monika (2005) “Bond Yields and the Federal Reserve,” Journal of Political Economy, Vol. 113, No. 2, pp. 311–344, April. Rudebusch, Glenn D. (1995) “Federal Reserve interest rate targeting, rational expectations, and the term structure,” Journal of Monetary Economics, Vol. 35, No. 2, pp. 245–274, April. Rudebusch, Glenn D., Brian P. Sack, and Eric T. Swanson (2006) “Macroeconomic implications of changes in the term premium,”Technical report. Rudebusch, Glenn D. and Tao Wu (2008) “A Macro-Finance Model of the Term Structure, Monetary Policy and the Economy*,” The Economic Journal, Vol. 118, No. 530, pp. 906–926. Sack, Brian (2000) “Using Treasury STRIPS to measure the yield curve,”Technical report. Skarr, Doug (2010) “Understanding Interest Rate Swaps: Math and Pricing,” California Debt and Investment Advisory Commission. Söderlind, Paul (2010) “Reaction of Swiss Term Premia to Monetary Policy Surprises,” Swiss Journal of Economics and Statistics (SJES), Vol. 146, No. I, pp. 385–404, March. Söderlind, Paul and Lars Svensson (1997) “New techniques to extract market expectations from financial instruments,” Journal of Monetary Economics, Vol. 40, No. 2, pp. 383–429, October. Svensson, Lars E.O. (2003) “Monetary policy and learning,” Economic Review, No. Q3, pp. 11–16. Swanson, Eric T. (2007) “What we do and don’t know about the term premium,” FRBSF Economic Letter, No. Jul 20. Taylor, John B (1995) “The Monetary Transmission Mechanism: An Empirical Framework,” Journal of Economic Perspectives, Vol. 9, No. 4, pp. 11–26, Fall. Taylor, John B. (2010) “Getting back on track: macroeconomic policy lessons from the financial crisis,” Review, No. May, pp. 165–176. Woodford, Michael (2003) Interest and Prices: Foundations of a Theory of Monetary Policy: Princeton University Press, 1st edition.

29

7

Appendix - Tables and Figures

Table 1: FOMC Decisions. The reported FOMC meetings that resulted in an interest rate decision including both scheduled and unscheduled meetings. The sample covers 23 decisions to hike the policy rate, 26 cut and 45 hold decisions. Column Surprise reports the unexpected element of every decision extracted from the Fed futures market and following Kuttner (2001). Every absolute value of the surprise indicator that exceeds 2/3 of the 25 basis points categorises a decision as a Surprise Hike/Surprise Cut. Day

Month

Year

3 30 18 30 24 5 16 21 2 21 16 28 22 3 15 19 3 31 20

Feb Mar May Jun Aug Oct Nov Dec Feb Mar May Jun Aug Oct Nov Dec Jan Jan Mar

1999 1999 1999 1999 1999 1999 1999 1999 2000 2000 2000 2000 2000 2000 2000 2000 2001 2001 2001

Decision (bp) 0 0 0 25 25 0 25 0 25 25 50 0 0 0 0 0 -50 -50 -50

30

surprise (bp) 1 -1 10 -8 2 6 11 12 2 -1 8 -2 3 2 0 3 -24 -5 -7

Surprise Hike

Surprise Cut

*

Day

Month

Year

18 15 27 21 17 2 6 11 30 19 7 26 13 24 6 10 29 18 6 25 12 16 28 9 28 16 4 30 10 21 10 14

Apr May Jun Aug Sep Oct Nov Dec Jan Mar May Jun Aug Sep Nov Dec Jan Mar May Jun Aug Sep Oct Dec Jan Mar May Jun Aug Sep Nov Dec

2001 2001 2001 2001 2001 2001 2001 2001 2002 2002 2002 2002 2002 2002 2002 2002 2003 2003 2003 2003 2003 2003 2003 2003 2004 2004 2004 2004 2004 2004 2004 2004

Decision (bp) -50 -50 -25 -25 -50 -50 -50 -25 0 0 0 0 0 0 -50 0 0 0 0 -25 0 0 0 0 0 0 0 25 25 25 25 25

31

surprise (bp) -79 -22 9 -3 -28 -14 -16 -7 3 53 -4 -9 -7 -2 -12 0 2 4 -7 11 0 0 -1 1 3 0 1 -8 7 8 1 0

Surprise Hike

Surprise Cut * *

*

Day

Month

Year

2 22 3 30 9 20 1 13 31 28 10 29 8 20 25 12 31 9 28 7 18 31 11 22 30 18 30 25 5 16 8 29 16

Feb Mar May Jun Aug Sep Nov Dec Jan Mar May Jun Aug Sep Oct Dec Jan May Jun Aug Sep Oct Dec Jan Jan Mar Apr Jun Aug Sep Oct Oct Dec

2005 2005 2005 2005 2005 2005 2005 2005 2006 2006 2006 2006 2006 2006 2006 2006 2007 2007 2007 2007 2007 2007 2007 2008 2008 2008 2008 2008 2008 2008 2008 2008 2008

Decision (bp) 25 25 25 25 25 25 25 25 25 25 25 25 0 0 0 0 0 0 0 0 -50 -25 -25 -75 -50 -75 -25 0 0 0 -50 -50 -75

surprise (bp) 1 29 2 2 0 7 1 0 1 6 1 -7 -5 0 -2 -1 0 1 2 8 -44 9 0 -63 -10 49 -7 -4 -3 21 -17 -10 -35

32

Surprise Hike

Surprise Cut

*

*

* *

* (as anticipated)

Table 2: Macro Uncertainty. The output of univariate regressions with intercept, where monthly changes (upper panel) and squared demeaned monthly changes (lower panel) in the single U.S. macroeconomic variables: Federal Funds rate (FFR), Unemployment rate (UNEMPL), Headline inflation (CPI), Industrial production all-industries (IP) and 10Y yields from STRIPS are regressed on the implied volatilities across the available cap maturities (1Y to 4Y, 7Y and 10Y). The Table reports the regression coefficients and the coefficient of determination of the single univariate regressions. The levels of significance reported are: * 0.1, **0.05 and ***0.01. βˆ

FFR U NEMPL CP I IP y10Y

IV1Y -0,005** 0,003 -0,005 -0,007 -0,004

IV2Y -0,012*** 0,004 -0,018 -0,012 -0,019***

IV3Y -0,016*** 0,005 -0,026* -0,013 -0,038***

IV4Y -0,020*** 0,005 -0,032* -0,024 -0,054***

IV7Y -0,026*** 0,005 -0,039* -0,034 -0,075***

IV10Y -0,033*** 0,006 -0,054* -0,042 -0,095***

0,074 0,009 0,025 0,014 0,358

0,070 0,006 0,021 0,016 0,390

0,074 0,005 0,025 0,016 0,405

IV4Y 0,014*** 0,001 0,121*** 0,077** 0,007*

IV7Y 0,019*** 0,001 0,128*** 0,114*** 0,013***

IV10Y 0,024*** 0,002 0,152*** 0,143*** 0,021***

0,070 0,007 0,036 0,042 0,052

0,071 0,006 0,032 0,042 0,078

R2 FFR U NEMPL CP I IP y10Y

0,030 0,016 0,003 0,007 0,011

0,059 0,016 0,016 0,007 0,094

0,070 0,014 0,023 0,006 0,250 βˆ

FFR 2 U NEMPL2 CP I 2 IP 2 2 y10Y

IV1Y 0,004*** 0,001*** 0,039*** 0,020 0,002

FFR 2 U NEMPL2 CP I 2 IP 2 2 y10Y

0,037 0,036 0,033 0,012 0,008

IV2Y 0,010*** 0,001* 0,097*** 0,034 0,004

IV3Y 0,012*** 0,001 0,110*** 0,055* 0,005

R2 0,068 0,024 0,075 0,014 0,016

0,073 0,015 0,066 0,025 0,016

33

0,073 0,007 0,056 0,034 0,023

Table 3: Candidate Market Price of Risk Specifications Rank 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 Sum 24 ... 35

j

j

j

j

j

j

λ(1,1)

λ(2,2)

λ(1,3)

λ(2,1)

λ(1,2)

λ(2,3)

1 0 1 0 0 1 1 0 0 0 0 1 0 0 0 0 0 0 1 0 1 0 1 8 0 ... 0

0 0 0 0 0 0 0 0 1 0 0 1 0 0 0 1 1 1 0 0 1 0 1 7 0 ... 0

1 1 1 1 0 0 0 0 1 0 1 1 0 0 1 1 1 0 1 1 0 1 1 14 0 ... 1

0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 1 0 0 0 0 1 0 4 1 ... 1

0 0 0 0 0 0 0 0 0 1 1 0 1 1 1 0 0 0 1 1 0 1 0 8 0 ... 0

1 0 0 1 0 0 1 1 1 0 0 1 1 1 1 0 1 0 1 1 0 0 0 12 0 ... 1

34

BIC -8.41 -8.40 -8.39 -8.39 -8.39 -8.37 -8.34 -8.34 -8.20 -8.19 -8.17 -8.17 -8.16 -8.13 -8.12 -8.08 -8.07 -8.06 -8.04 -8.01 -8.00 -7.97 -7.91 -7.39 ... -2.30

Table 4: Parameters. Estimated parameters and numerical standard errors (in parenthesis), along with the support interval and the acceptance ratio are reported. The columns IF denote the estimated inefficiency factor computed as P 1 + 2 Ll=1 ρ ( l ) where ρ ( l ) is the autocorrelation at lag l in the Markov chain sequence of a parameter, and L is the value at which the autocorrelation function goes to zero. Further details can be found in Chib (2001). LB

UB

Ψ1,1

-

-

Ψ2,2

0.001

0.999

Ψ1,3

-0.999

0.999

Ψ2,3

-0.999

0.999

Ψ3,3

0.001

0.999

LB

UB

j Σ1,1

0.001

15.000

Σ2,2

j

0.001

15.000

Σ3,3

j

0.001

15.000

λ0,1

j

-100.000

-0.001

λ0,2

j

-100.000

-0.001

λ1,13

j

-100.000

100.000

λ1,23

j

-100.000

100.000

ωj

0.001

10.000

Common Parameters θ

AccRatio

IF

1.000 0.772 0.506 70.9 (0.002) 0.000 0.511 113.4 (0.000) -0.068 0.491 129.9 (0.001) 0.996 0.521 16.7 (0.002) Regime-Dependent Parameters Non-policy Days Policy Days θ AccRatio IF θ AccRatio 0.050 0.503 7.3 0.038 0.524 (0.000) (0.002) 0.187 0.523 114.4 0.195 0.552 (0.002) (0.010) 0.350 0.504 4.0 0.418 0.553 (0.003) (0.023) -0.002 0.582 59.5 -0.116 0.520 (0.002) (0.099) -1.897 0.465 59.5 -1.777 0.522 (0.031) (0.178) -0.102 0.439 472.8 -0.112 0.458 (0.007) (0.040) -0.241 0.394 427.2 -0.237 0.421 (0.007) (0.025) 0.370 0.525 9.0 0.395 0.539 (0.004) (0.022)

35

IF 9.2 106.8 7.5 59.8 63.6 414.9 348.7 5.0

Table 5: Mean Absolute Pricing Errors. The values are reported for the singleand the two-set model across maturities and the pricing errors of the fitted models regard the entire dataset. The estimated one-set model is separately estimated and not embedded in the two-set model. In-Sample (Jan99 - Dec08) 1Y 3Y 5Y Policy Days (82) Two-Set One-Set - Anticipated Moves (71) Two-Set One-Set - Surprise Moves (11) Two-Set One-Set All Days (2608) / Two-Set One-Set

7Y

9Y

20.07 19.99

19.14 19.04

9.19 9.22

6.38 6.56

1.89 2.04

18.29 18.18

18.08 17.96

8.63 8.63

5.97 6.16

1.76 1.91

31.58 31.73

26.01 26.02

12.83 13.05

8.99 9.17

2.74 2.83

18.00 17.99

18.45 18.39

8.70 8.64

6.14 6.09

1.92 1.90

Table 6: Variance Decomposition. The Table reports the variance decomposition of selected yields in proportions. The variance is decomposed by dividing each sinj gle state variable shock j to an n-periods yield: MSE n = B0n Σ j B n + B0n ΨΣ j ΨB n , where Σ j is a K × K matrix with zeros and a non-zero j j element corresponding to the volatility of state variable j ; with the overall Mean Squared Error of forecasting the states 1 period ahead (upper panel): MSE n = B0n ΣB n + B0n ΨΣΨB n ; or equvallently, 60 periods ahead (lower panel). Non-Policy Days

Policy Days

X l evel X slope IV

6m Libor 0.289 0.690 0.022

5Y 0.483 0.469 0.047

1 Day 10Y 6m Libor 0.627 0.172 0.327 0.794 0.046 0.035

X l evel X slope IV

6m Libor 0.376 0.605 0.019

5Y 0.546 0.410 0.044

3 Months 10Y 6m Libor 0.672 0.234 0.286 0.735 0.043 0.031

36

5Y 0.317 0.595 0.088

10Y 0.448 0.452 0.101

5Y 0.367 0.547 0.086

10Y 0.487 0.415 0.098

Table 7: One-Day Average Change of the Term Premia in basis points. The table reports the difference between the mean one-day forward premia change on a policy action day and the mean one-day premia change on a non-policy day. The premia are estimated using both the single-set model (upper panel) and the twoparameter-set model (lower panel). The levels of significance reported are: *0.1, **0.05 and ***0.01 and come from an independent two-sample t-test. Occasion/Maturity All Decisions (82) - Hikes (23) - Cuts (23) – Anticipated Cuts (15) – Surprise Cuts (8) - Holds (36)

Occasion/Maturity All Decisions (82) - Hikes (23) - Cuts (23) – Anticipated Cuts (15) – Surprise Cuts (8) - Holds (36)

One-Set Model 1Y 3Y -0,13 -0,36 (0,70) (0,71) -0,19 -0,58 (0,66) (0,65)

5Y -0,46 (0,72) -0,74 (0,66)

7Y -0,48 (0,72) -0,79 (0,67)

10Y -0,46 (0,74) -0,79 (0,68)

-3.38* (0,10) 9.81* (0,08) 0,04 (0,99)

-4.32* (0,05) 10.77* (0,09) 0,17 (0,96)

-5.51** (0,02) 11,35 (0,10) 0,39 (0,91)

5Y 1,86 (0,15) 1,5 (0,38)

7Y 0,18 (0,89) -0,19 (0,91)

10Y -6.14*** 0,00 -6.51*** (0,00)

-0,95 (0,62) 12.16*** (0,01) 0,04 (0,99)

-3,57* (0,10) 11.46* (0,07) 0,17 (0,96)

-11.15*** (0,00) 5,68 (0,40) 0,39 (0,91)

-0,65 -2,11 (0,22) (0,17) 2.48* 7.4* (0,07) (0,08) -0,04 -0,06 (0,96) (0,98) Two-Set Model 1Y 3Y 0,04 1,48 (0,89) (0,13) -0,05 1,18 (0,91) (0,37) -0,44 (0,39) 2.66** (0,05) -0,04 (0,96)

-0,18 (0,90) 9.27*** (0,00) -0,06 (0,98)

37

Table 8: One-Day Average Change of Future Short-Rate Expectations in basis points. The table reports the difference between the average 1-day change of future short-rate expectations on a policy action day and the average 1-day expectations change on a non-policy day for the single-set model (upper panel) and the two-set model (lower panel). The levels of significance reported are: *0.1, **0.05 and ***0.01 and stem from an independent two-sample t-test. Occasion/Maturity All Decisions (82) - Hikes (23) - Cuts (23) – Anticipated Cuts (15) – Surprise Cuts (8) - Holds (36)

Occasion/Maturity All Decisions (82) - Hikes (23) - Cuts (23) – Anticipated Cuts (15) – Surprise Cuts (8) - Holds (36)

One-Set Model 1Y 3Y -0,26 -0,10 (0,65) (0,89) 1,12*** 1,05 (0,00) (0,11) -0,22 1,22 (0,80) (0,22) -6,68 -7,82 (0,13) (0,12) 0,16 0,01 (0,81) (0,99) Two-Set Model 1Y 3Y -0,27 -0,12 (0,64) (0,87) 1,13*** 1,08* (0,00) (0,10) -0,24 (0,77) -6,69 (0,13) 0,16 (0,81)

38

1,13 (0,24) -7,85 (0,12) 0,01 (0,99)

5Y 0,00 (0,99) 1,00 (0,25)

7Y 0,03 (0,97) 0,99 (0,30)

10Y 0,04 (0,95) 0,98 (0,32)

2.01* (0,09) -8,44 (0,12) -0,07 (0,96)

2.29* (0,07) -8,66 (0,11) -0,10 (0,95)

2.41* (0,07) -8,76 (0,11) -0,11 (0,95)

5Y -0,03 (0,97) 1,05 (0,23)

7Y 0,00 (0,99) 1,04 (0,27)

10Y 0,01 (0,98) 1,03 (0,29)

1.89* (0,10) -8,49 (0,12) -0,07 (0,96)

2.16* (0,08) -8,72 (0,11) -0,10 (0,95)

2.28* (0,07) -8,82 (0,11) -0,11 (0,95)

Figure 1: The upper panel illustrates the U.S. yield curve dynamics from the beginning of 1999 to the end of 2008. The vertical green (solid) lines represent the FOMC decisions to hike the target rate, while the red (dashed) lines are days when the Fed opted for an interest rate cut. The lower panel shows the term spread (blue solid) on the left scale versus the 10-year implied volatility (red dashed) on the right scale.

39

Figure 2: The figure reports 24-month rolling window standard deviations of the US unemployment rate (blue solid), industrial production (green solid) and consumer price index (red solid) on the left-hand side and the 10-year implied volatility (black dashed line) on the right-hand side.

40

Figure 3: The upper panel shows the estimated level (solid blue) and slope (dashed green) factors backed out from 6m Libor and 10Y swap rate, respectively. The lower panel reports the explicit factor loadings for all-days parameters (solid red) and FOMC decision days parameters (dashed red).

41

Figure 4: The figure reports the parameters’ posterior densities. The algorithm is ran for 100,000 times and the first 40,000 sweeps are considered as the “burn-in” period.

42

Figure 5: The four panels report the two-parameters-set modelled reaction of the forward term premia to all decisions in the sample (upper left), all decisions to hike Fed funds rate (upper right), anticipated interest rate cuts (lower left) and surprise interest rate cuts (lower right) together with the 90 percent credible interval. The credible interval is calculated by using every 100-th sweep from one of the Markov chains after burn-in.

43

Figure 6: The four panels report the single-parameter-set modelled reaction of the forward term premia to all decisions in the sample (upper left), all decisions to hike Fed funds rate (upper right), anticipated interest rate cuts (lower left) and surprise interest rate cuts (lower right) together with the 90 percent credible interval.

44

Figure 7: The four panels report the single-parameter-set modelled reaction of the expected future short-term interest rate to all decisions in the sample (upper left), all decisions to hike Fed funds rate (upper right), anticipated interest rate cuts (lower left) and surprise interest rate cuts (lower right) together with the 90 percent credible interval.

45

US Term Premia around FOMC decisions

called “slope effect” of the US monetary policy instrument on the yield curve. ..... 22To my best knowledge, the first study that uses implied volatilities from interest rate .... where Q is the risk-neutral probability measure and y1,t is the short-term interest ...... for Social Scientists: Springer Science+Business Media, 7th edition.

488KB Sizes 1 Downloads 118 Views

Recommend Documents

US Term Premia around FOMC decisions
... Aleksandar Zdravkovic, the seminar and conference participants at the ..... One interest rate cap is essentially a portfolio of call options, so called caplets, on ...... Kim, Don H. and Jonathan H. Wright (2005) “An arbitrage-free three-factor

US FOMC - No tapering yet
Fed Chairman Ben Bernanke reiterated that the tapering plan was never a ..... Securities and Exchange Board of India (SEBI) ... to any direct, indirect or consequential losses, loss of profits and damages) of any reliance thereon or usage thereof. ..

Term Premia and Inflation Uncertainty: Empirical ...
−360. BC. −628. −521. −107. Australia. −237. 2. −630. OLS. −621. −53. −567 ... OLS and BC term premia conform to these views following the methods of Wright.

Term Structure of Consumption Risk Premia in the ...
Jul 4, 2016 - risk sensitivities of exchange rates to alternative current period shocks. .... Second, my model has stochastic variance: I account for the variation in the ... baskets, buys τ− period foreign risk-free bonds, and at time t+τ ......

around & around & around by roy starkey
Mar 19, 2015 - By Erling Vikanes. An extraordinary tale of a extraordinary life of an extraordinary person. Roy Starkey shows us that nothing is impossible for the one who dares. I wish that this book will be an inspiration for todays younger generat

Risk premia and unemployment fluctuations
Jun 1, 2018 - We further study to which extent our analysis could be affected by the presence of ... Firms post vacancies at a per-period cost κ, and the vacancies are ... allows us to construct a lower bound on the required volatility of risk ...

Gold – FOMC-led cycle lows are rising -
However, the cycle lows have risen after each of the past hikes. ... special rate of 3% for gold, which replaces the excise duty (1%) and state VAT .... or may be the only market maker in relation to such investments, or provide, or have provided ...

Inflation Experiences on the FOMC - Econometrics Laboratory, UC ...
Feb 27, 2017 - 2. In this paper, we analyze how lifetime experiences influence monetary policy views and the ... crtical-of-policies-set-to-retire-in-march.html. 2 ...

Inflation Experiences on the FOMC - Econometrics Laboratory, UC ...
Feb 27, 2017 - of birth, place of birth, gender, highest degree earned, the program ...... for having earned a PhD, a JD, an MBA, or a Master's degree as ... hand, most comprehensive in accounting for unobserved person-specific determinants.