diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzhnfm" "b/data_all_eng_slimpj/shuffled/split2/finalzzhnfm" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzhnfm" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nMuch of our knowledge of gas giant exoplanets comes from those transiting in front of bright stars on close-in orbits. These planets are easier to detect and are more favorable to follow-up observations such as radial velocity measurements. Their atmospheric characterization through transit spectroscopy is also facilitated by the large amount of flux received from bright stars. The vast majority of these planets have been discovered by dedicated ground-based photometric surveys, the most prolific being WASP \\citep{Pollacco2006, Collier2007} and HAT \\citep{Bakos2002, Bakos2004}. These two projects detected \\simi80 hot Jupiters around stars of magnitude $V \\leq 11$. The \\textit{CoRoT}\\xspace and \\textit{Kepler}\\xspace missions provided little addition to this sample because they target fainter stars and much smaller sky areas.\nThe XO project \\citep{McCullough2005} aims at detecting transiting exoplanets around bright stars from the ground using small telescopes. The project started in 2003 and discovered five close-in gas giant planets, XO-1b to XO-5b \\citep{McCullough2006, Burke2007, JohnsKrull2008, McCullough2008, Burke2008}. A second version of XO was deployed in 2011 and 2012 and operated nominally from 2012 to 2014. This yielded the discovery of the hot Jupiter XO-6b orbiting a bright, hot, and fast rotating star \\citep{Crouzet2017}, and of other planet candidates.\nIn this chapter, we provide an overview of this second version of XO. First, we present the goals of the project, the instrumental setup, and the observation strategy. Then, we detail the various steps of the data reduction pipeline, and we review the instrumental performances. Finally, we present the search for periodic signals in the lightcurves, the identification and follow-up observations of transit candidates, and we compare the detection yield to the number of detected planets.\n\n\n\n\\section{Goals of the project}\n\\label{sec: Goals}\n\nThe XO project aims at detecting transiting exoplanets around bright stars ($V < 11$) including some with long orbital periods ($P > 10$ days). Bright host stars ensure that follow-up observations can be achieved to confirm the planet candidates, and make their atmospheric characterization possible. Besides, whereas close-in transiting giant planets have been discovered and studied by dozens, only a few transiting ones with long orbital periods and bright host stars are known. The long period severely decreases the transit probability making such detections challenging. As a result, constraints on the physical properties of long period giant exoplanets have been obtained only recently from a statistical study of \\textit{Kepler}\\xspace objects, and their atmospheric properties are largely unknown \\citep[see][and the chapter entitled \\underline{Hot Jupiter Populations from Transit and RV Surveys} of this book]{Santerne2016}. \n\n\n\n\n\\section{Instrumental setup}\n\\label{sec: Instrumental setup}\n\nXO is composed of three identical units installed at Vermillion Cliffs Observatory, Kanab, Utah, at Observatorio del Teide, Tenerife, Canary Islands, and at Observatori Astron\\`omic del Montsec, near \\`Ager, Spain. Each unit is composed of two 10 cm diameter and 200 mm focal length Canon telephoto lenses equipped with Apogee E6 1024$\\times$1024 pixels CCD cameras, mounted on a German-Equatorial Paramount ME mount, and protected by a shelter with a computer-controlled roof (Figure \\ref{fig: XO units}). \nThe basic hardware is described in \\citet{McCullough2005} but was changed a bit in the transition from the first version of XO at a single observatory at Haleakala, Hawaii, to the three observatories in the small roll-off roof enclosures. The main change was from back-illuminated SITe CCDs with parallel port interface to front-side illuminated Kodak CCDs with Ethernet interface, as a cost compromise (the manufacturer SITe stopped making the inexpensive back-side illuminated CCDs).\nAll six lenses and cameras operate in a network configuration: they point toward the same fields of view over the night at the three locations. The CCDs are read in time delayed integration (TDI): pixels are read continuously while stars move along columns on the detector. The scan and read rates are adjusted to maintain round PSFs (Point Spread Functions). This setup results in strips of $43^{\\circ}.2\\times7^{\\circ}.2$ on the sky instead of square images. This technique maximizes the number of observed bright stars by enlarging the effective field of view. The exposure time to acquire a full strip is 5.3 minutes, which corresponds to 53 seconds on each $7^{\\circ}.2\\times7^{\\circ}.2$ square subimage. The nominal PSF FWHM (Full Width Half Maximum) is 1.2 pixels and varies in practice between 1 and 2.5 pixels (see Section \\textit{\\nameref{sec: FWHM variations}}). The units can be controlled remotely and operate robotically. Data from a weather sensor are recorded and analyzed in real time; they are used to send instructions such as opening or closing the roof, running the observations, etc. A webcam sensitive in the optical and infrared is mounted inside each enclosure and can be accessed from a web interface. A computer running on Linux is installed inside each unit to control the operations and store the data. An IDL program commands the operations of the unit and runs the observations. This program as well as other daily tasks are launched using crontab. \n\n\n\n\\begin{figure}\n\\includegraphics[width=5cm]{crouzet_fig1.png}\n\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig2.png}\n\\caption{XO units at Vermillion Cliffs Observatory, Kanab, Utah (left) and at the Observatorio del Teide, Canary Islands (right). A third unit is installed at the Observatori Astron\\`omic del Montsec near \\`Ager, mainland Spain.}\n\\label{fig: XO units}\n\\end{figure}\n\n\n\\newpage\n\n\\section{Observations}\n\\label{sec: Observations}\n\nThis section presents the observation strategy that we adopted for XO and gives an overview of the operations.\n\n\n\\subsection{Observation strategy}\n\\label{sec: Observation strategy}\n\nWe observed two strips with respective centers at $RA = 90^{\\circ}$ and $RA = 270^{\\circ}$ and $Dec$ extending from $+90^{\\circ}$ to $+54^{\\circ}$. This corresponds to an effective sky area of $520^{\\circ 2}$. We observed these strips for twice nine months between 2012 and 2014, one strip at a time depending on their observability. An example of one night observations with XO is shown in Figure \\ref{fig: XO observations}. Darks and flats are taken at the beginning and end of the night, science strips are taken the rest of the night. The scan direction is either North or South with a switch occurring around the meridian crossing. Under nominal conditions, we collect from 60 to 100 TDI strips each night depending on the season. \n\n\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig3.png}\n\\caption{Example of observing night with XO. These data were taken by one camera of the unit located at the Observatorio del Teide, Canary Islands, during the night of June 8, 2013. Only a sample of the data is shown. The TDI mode yields long strips instead of square images. The operations are summarized at the bottom from dusk (left) to dawn (right). The blue triangle near the meridian crossing indicates parts of the strips that are contaminated by one of the enclosure's walls (see Section \\textit{\\nameref{sec: Data selection}}).}\n\\label{fig: XO observations}\n\\end{figure}\n\n\n\n\\subsection{Data quality check}\n\nWe developed a data quality check software program that analyzes the observations after each night. All science and dark files are listed and their size is checked. Files with a very small size are corrupted (usually they contain a header with no data) and are removed from the list.\nThen, for each science image (strip), we measure its size in pixel units, the corresponding number of 1024$\\times$1024 sub-images, the median intensity, the sky background, and the number of point sources. We also compute the FWHM in the $x$ and $y$ directions using 100 bright, non-saturated stars. The median intensity and the number of point sources are also computed for each 1024$\\times$1024 sub-image. This program is ran for both cameras at each site.\nIn addition, we summarize the weather data and the CPU temperature over the night. We also report the status of the network power strip (NPS) and of the enclosure roof, the last commands that were issued at the end of the observations, and the available disk space.\nThe results are sent via email (Figure \\ref{fig: diagnostics}) with a separate alert email if the NPS is not in a nominal state. This program runs every day at each site.\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig4.png}\n\\caption{Example of nightly diagnostics file for the night of March 2, 2013 for the XO unit at Observatori Astron\\`omic del Montsec. ``Michael'' and ``Walt'' are the nicknames of the two cameras. Each camera recorded 101 TDI strips; the diagnostics information for only three strip is shown for clarity (as indicated by `...'). The weather sensor information, NPS state, enclosure's roof state, CPU temperature, and Sun altitude are also reported.}\n\\label{fig: diagnostics}\n\\end{figure}\n\n\n\n\\subsection{Data management}\n\nThe data are stored on the computers at each site and are transferred periodically on servers at Space Telescope Science Institute (STScI) through internet using a $rsync$ command. A 8 TB storage space was allocated to the project. The full analysis of the data is performed on the STScI servers.\n\n\n\n\n\\section{Data reduction}\n\\label{sec: Data reduction}\n\nThe data reduction pipeline is divided in two phases. The first phase consists of reducing the data of each night. The second phase consists of gathering the data taken on the same field over all nights and to build the lightcurves. Data from the six cameras are reduced independently and the lightcurves are merged at the very end of the data reduction.\nThe pipeline is composed of \\textit{IDL} routines that are launched with \\textit{Bash} programs. The \\textit{Bash} commands iterate on each night and then on each field. The main advantage is that if \\textit{IDL} crashes, the \\textit{Bash} program just moves on to the next night or field.\n\n\n\n\\subsection{Carving the strips}\n\n\nFor each night, the strips are carved into $7^{\\circ}.2\\times7^{\\circ}.2$ square images. A World Coordinate System (WCS) solution is found using the astrometry.net software program \\citep{Lang2010} plus a six parameter astrometric solution. We use a two step process: first, each square image is centered on the coordinates of a predefined field (listed in Table \\ref{tab: xo fields}), then a precise astrometric solution is found. Observing in TDI mode allows us to observe these square fields more efficiently than simply cycling through them using standard telescope pointing and CCD readout, as done by other transit surveys.\n\n\n\n\\begin{table}\n\\begin{center}\n\\caption{Central coordinates of the fields of view observed by XO. Because we observe in TDI mode, fields with the same $RA$ are observed within one scan. Only a small amount of data is obtained for fields 10 to 12.}\n\\label{tab: xo fields}\n\\vspace{3mm}\n\\begin{tabular}{cccccccccccccc}\n\\hline\n\\hline\nField & 00 & 01 & 02 & 03 & 04 & 05 & 06 & 07 & 08 & 09 & 10 & 11 & 12 \\\\\n\\hline\n$RA \\rm \\; [^{\\circ}]$ & 270 & 270 & 270 & 270 & 270 & 90 & 90 & 90 & 90 & 270 & 270 & 90 & 90 \\\\\n$Dec \\rm \\; [^{\\circ}]$ & +90 & +83 & +76 & +69 & +62 & +83 & +76 & +69 & +62 & +55 & +48 & +55 & +48 \\\\\n\\hline\n\\hline\n\\end{tabular}\n\\end{center}\n\\end{table}\n\n\n\n\n\\subsection{Image calibration}\n\n\nAbout five dark frames and five flat frames are taken for each camera at the beginning and end of the night. They come as strips of the same size as the science images, but because we use the TDI mode, the final darks and flats are 1-D vectors with 1024 elements corresponding to the columns.\nFor each dark frame, an outlier resistant mean is computed for each column after excluding the first 1024 rows. The resulting 1-D vectors that have a mean value between 500 and 2000 ADU (Analog to Digital Unit) are averaged to create a final dark. Those with higher or lower median values are not used, as they are contaminated by parasitic light for example. One dark is created for each night. In the absence of dark frames, we use the dark of the previous or the following night.\n\n\n\nOne flat is built for each camera. We use the TDI images recorded during twilight and calibrate them with the dark. We eliminate pixels that are close to saturation ($>$~50,000 ADU) as well as a region of $3\\times3$ pixels around them.\nWe eliminate stars in the following way. We build a median-filtered image where each pixel is replaced by the median of the $15\\times15$ surrounding pixels, and subtract it to the initial image. We compute the outlier-resistant standard deviation of the residual image. Pixels that are above three times this standard deviation are flagged and are removed from the initial image. This yields an image equivalent to the initial image but with the stars removed.\nThen, we divide each row by its median in order to give the same weight to all rows. We compute an outlier-resistant mean of the image along the columns (starting at row 1024) and normalize the resulting vector by its median. This yields a flat vector. We average the flat vectors obtained from each twilight image to build a flat vector for each day (Figure \\ref{fig: flat-fields}), and we average the daily flat vectors obtained from the first months of observations. \nPermanent warm columns are identified using the darks and are replaced by the average of the 10 neighboring valid columns (here a ``column'' refers to one vector element). We suppress high frequency variations by replacing each element by the average of its 10 neighbors. We normalize this vector by the median of the 50 central elements. This yields a masterflat that we use for all the observations. This procedure is repeated for all six camera.\nFinally, the science image calibration is performed by subtracting the dark vector to each row and dividing it by the flat vector.\n\n\n\nThe strips are affected by warm columns. There are typically 10 warm columns per image except for one CCD that produces about 100 warm columns. About a third are permanent and two thirds vary from image to image. We correct for them in each image in the following way. We compute a smoothed version of each row using a 11-pixel running median and subtract it to the original row. Then, we compute the median of each column in the residual image. Those exceeding a threshold are considered as warm; we use a threshold of 15 ADU for the six CCDs after a visual inspection. For these columns, this median is subtracted to the same column in the original image in order to remove the excess counts. We record the number of corrected columns, their indices, and their original values in the image header.\n\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig5.pdf}\n\\caption{Example of flats built from twilight images taken during March, 2013 with one camera, zoomed in between columns 400 and 560. Because we use the TDI mode, flats are one-dimensional vectors with 1024 elements. We normalize and average these flats to built a masterflat for each camera and use it for all the observations.}\n\\label{fig: flat-fields}\n\\end{figure}\n\n\n\n\\subsection{Flux extraction}\n\n\nWe select around 6000 target stars in each field for the photometry using an image taken under excellent conditions (the ``reference image'', Figure \\ref{fig: reference image}). Stellar fluxes are extracted by circular aperture photometry using the \\textit{Stellar Photometry Software} program \\citep[\\textit{SPS,}][]{Janes1993}. We compute the sky background for each star using an annulus around the aperture and subtract it. To measure the flux, \\textit{SPS} does not simply sum the pixels inside the aperture: instead, it computes an intensity profile by interpolating between pixels, and measures the flux by integrating this profile over the aperture size. In the original version of the XO pipeline, only one aperture of 3 pixel radius was used for all the stars. We improved this by optimizing the aperture for each star, as described next. For this process, we use only high quality data that were selected after a first flagging procedure and visual inspection of the star -- epoch magnitude arrays (see Sections \\textit{\\nameref{sec: Building the lightcurves}} and \\textit{\\nameref{sec: Data selection}}). \nFor each star, we build lightcurves for nine apertures with radii ranging from 2 to 10 pixels, calculate their RMS, and select the aperture that leads to the smallest RMS. We use the point-to-point RMS rather than the standard RMS so the aperture optimization is not affected by long term trends. This is justified because point sources are identified in each image and the apertures are placed at their centroids.\nThe fields of view are crowded so the best aperture for a given star may be affected by other sources, and it shows large fluctuations for stars of similar magnitudes. To suppress these fluctuations, we build a function that sets the optimal aperture for each stellar magnitude. We divide the magnitude range [6, 15] into 100 bins, calculate the mean best aperture in each bin, and fit this distribution by a fourth order polynomial (Figure \\ref{fig: aperture function}). The last step is an iterative process to have the magnitude of each star correspond to that measured in its optimal aperture. \nWe run this process for two representative fields (00 and 02) and average them to obtain a final aperture function. We build an aperture function for each camera. Differences between cameras are due to different focus or stability, for example. Finally, the optimal aperture for each star is given by the aperture function rounded to the closest integer (because only a small number of apertures can be used, due to a limitation of \\textit{SPS}).\n\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig6.png}\n\\caption{Reference image for field 06. Point sources identified in all the images of this field are matched against this image. The image size is 1024$\\times$1024 pixels and $7^{\\circ}.2\\times7^{\\circ}.2$ on the sky. The gray scale at the bottom indicates the intensity in ADU (most bright pixels are beyond the scale limit). The star hosting the transiting hot Jupiter XO-6b is indicated by a green circle.}\n\\label{fig: reference image}\n\\end{figure}\n\n\n\\begin{figure}\n\\includegraphics[width=5.7cm]{crouzet_fig7.png}\n\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig8.png}\n\\caption{Best photometric aperture as a function of stellar $R$ magnitude for two cameras, at Vermillion Cliffs Observatory (left) and at Observatori Astron\\`omic del Montsec (right). Each star is represented by a black dot. A mean is performed in each magnitude interval (red points) and a fourth order polynomial is fitted to define an aperture function (blue line). We use this function to define an aperture for each star.}\n\\label{fig: aperture function}\n\\end{figure}\n\n\n\n\\subsection{Building the lightcurves}\n\\label{sec: Building the lightcurves}\n\nThe strip carving, astrometry, image calibration, and flux extraction are performed for each night. We obtain a star coordinate -- magnitude file for each image. Because the \\textit{SPS} software identifies point sources in each image before doing the photometry (it does not handle input coordinates), we must correlate the sources found in each image to build the lightcurves. This correlation is made against the reference image. We include an image distortion correction performed in $x, y$ rather than in $RA, Dec$ because the observed fields are close to the celestial pole which yields degeneracies in $RA$. Then, we build a star -- epoch array that gathers the magnitudes recorded for all stars at all epochs. Similar star -- epoch arrays are built for the magnitude error, $RA$, $Dec$, $x$, $y$, sky background, airmass, and PSF cross-correlation parameter. We obtain one set of such arrays for each night, field, and camera.\n\n\nThe second phase of the pipeline is ran for each camera and each field. We combine all the nights together: we correlate the star coordinates from night to night and build one set of star -- epoch arrays covering all the observations. Examples of such arrays are shown in Figure \\ref{fig: cal arrays}. Because we are interested in bright stars, we truncate these arrays and keep only the 2000 brightest stars in each field, which corresponds to a limit $R$ magnitude around 12.5 and a photometric precision of 2-3\\%. We use only these stars in the rest of the pipeline.\n\n\n\n\n\\subsection{Photometric calibrations}\n\nThe next steps are photometric calibrations that are essential to improve the lightcurves from the raw photometry. These calibrations are applied independently for each scan direction (the strips are observed scanning North or South which yield different systematic effects). First, we compute the mean magnitude of each lightcurve and subtract it. From this point, the lightcurves consist of residual magnitudes rather than absolute magnitudes and we work with the residual magnitude arrays. We select reference stars in the following way. For a given star $s$, we subtract the photometric time series of $s$ from those of all the stars and evaluate the mean absolute deviation (MAD) of the resulting time series. Stars are sorted by increasing MAD; the first index is excluded because it corresponds to star $s$, and the following $N$ stars are kept as reference stars. Then, we build a reference lightcurve using an outlier-resistant mean of the $N$ reference stars, and subtract it to the lightcurve of star $s$. This yields a calibrated lightcurve for star $s$. We use $N = 10$. Lightcurves of stars with a bright nearby reference star are usually of better quality. For example, XO-6 has a reference star of magnitude $R = 9.6$ located at 5.6 arcmin (14 pixels) separation. This reference star makes XO-6b a good target for atmospheric characterization by differential spectrophotometry from the ground, if both stars can be placed on the detector. \nThen, for each epoch, we remove a 3rd order polynomial corresponding to low-frequency variations of the magnitude residuals in the CCD's \\textit{x,y} plane, also called ``L-flats\". We also remove a linear dependence of each lightcurve with airmass. At the end of the process, the mean magnitude of each lightcurve is subtracted again.\n\n\n\\subsection{Data selection}\n\\label{sec: Data selection}\n\n\nParts of the data are affected by poor weather, high sky background, instrumental malfunctions, etc., and must be removed. A lot of effort is put into inspecting the magnitude arrays, identifying the low quality parts, searching for the causes, defining criteria and procedures to flag them, and adjusting the flagging parameters. We summarize below the main causes of low quality data and the data selection procedures. Figure \\ref{fig: cal arrays} shows an example of star -- epoch magnitude arrays before and after the flagging.\n\n\n\\begin{itemize}\n\n \\item The main cause of bad data is poor weather. To quantify this, we calculate the standard deviation of the magnitude residuals at each epoch, $\\sigma_{ep}$, using stars in the magnitude range [8.3, 10]. We flag epochs with $\\sigma_{ep}$ above a given threshold that we set manually for each camera and each field (Figure ~\\ref{fig: histo sigep}). These thresholds are between 1.9\\% and 4.5\\% with an average at 2.3\\%.\n \n \\item We flag data with a sky background larger than 10,000 ADU and those with an airmass larger than 3.\n \n \\item We flag epochs where one of the enclosure's wall is visible in the images, which occurs when the observed field is at low altitude. Although part of the image could be used, some reference stars may be missing so we discard the whole image. These images show a large difference of sky background between the part pointing at the sky and that pointing at the wall. We identify them by calculating the standard deviation of the sky background across the image normalized to its mean. Those with a value larger than a given threshold are flagged.\n \n \\item We flag epochs where less than half of the stars have a valid magnitude measurement. \n\n \\item We flag epochs where the cross-correlation parameter averaged over all stars is lower than a threshold value, ranging from 0.61 to 0.74 depending on the camera and field. This parameter is a crude measurement of the PSF shape. In particular, it identifies images that are affected by imperfect tracking, which have a low correlation parameter (see Section \\textit{\\nameref{sec: Mount stability}}).\n\n\\end{itemize}\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig9.png}\n\\caption{Star -- epoch data arrays generated by the XO pipeline. Stars and epochs are sorted on the $x$ and $y$ axis by increasing magnitude and time, respectively. The full arrays contain 2000 stars and around 12,000 epochs. A subset of 475 stars $\\times$ 1640 epochs is displayed here, and corresponds to field 02 observed with one camera at the Observatorio del Teide between March 12 and April 25, 2013. From left to right: magnitude residuals before flagging, magnitude residuals after flagging, sky background, airmass, cross-correlation function. In the magnitude residual arrays, stars that have been discarded appear as white columns, variable stars appear as alternatively dark and bright columns, and epochs with an increased dispersion $\\sigma_{ep}$ appear as black and white regions (before they are flagged).}\n\\label{fig: cal arrays}\n\\end{figure}\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig10.pdf}\n\\caption{Histogram of the standard deviation of the magnitude residuals at each epoch, $\\sigma_{ep}$, for field 00 observed with one camera at Vermillion Cliffs Observatory over the whole duration of the observations (black line). The tail of the histogram corresponds to data taken under poor conditions, generally due to bad weather, that yield larger values of $\\sigma_{ep}$. A threshold (red line) is set manually for each field and each camera to flag these data. A similar approach is used for other flagging parameters.}\n\\label{fig: histo sigep}\n\\end{figure}\n\n\n\n\n\\subsection{Correction of systematic effects and final lightcurves}\n\nWe correct the remaining systematic effects in the cleaned residual magnitude arrays using the \\textit{SysRem} algorithm \\citep[\\textit{Systematic Removal},][]{Tamuz2005}. We correct for 10 eigenfunctions using 20 iterations, still per camera and field independently. Finally, we combine the lightcurves and other parameters obtained from the six cameras, interleaving the epochs and sorting them by ascending Julian date. This yields one set of star -- epoch arrays for each field.\n\n\n\n\n\\section{Instrumental performances}\n\\label{sec: Instrumental performances}\n\nIn this section, we review some technical issues that we encountered along the observations and present the photometric performances of the instruments.\n\n\\subsection{FWHM variations}\n\\label{sec: FWHM variations}\n\nAfter analyzing the first months of observations, we found significant variations of the PSF FWHM that are well correlated with the ambient temperature (Figure \\ref{fig: fwhm variations}). This effect is common and due to the lenses. Such PSF variations affect the flux measurements. To minimize this effect, the lenses were surrounded by a thermal regulation system made of flat and soft resistor strips for heating, a thermal probe, and a synthetic isolator, but this system was not efficient enough. After the first nine months, we increased the heating power and the isolation layer which reduced these FWHM fluctuations.\n\n\n\\begin{figure}\n\\includegraphics[width=5.7cm]{crouzet_fig11.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig12.png}\n\\includegraphics[width=5.7cm]{crouzet_fig13.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig14.png}\n\\includegraphics[width=5.7cm]{crouzet_fig15.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig16.png}\n\\caption{Variations of the PSF FWHM as a function of the ambient temperature during the first months of observations for the six cameras. Top: Vermillion Cliffs Observatory; middle: Observatorio del Teide; bottom: Observatori Astron\\`omic del Montsec. These variations were minimized afterwards by improving the temperature control system.}\n\\label{fig: fwhm variations}\n\\end{figure}\n\n\n\n\\subsection{Mount stability}\n\\label{sec: Mount stability}\n\n\nSome series of images show elongated PSFs or even two point sources separated by a few pixels instead of a single star (Figure \\ref{fig: elongated fwhm}). The cause is a mechanical defect in the mount resulting in an imperfect tracking. This behavior is seen only for a small fraction of the data taken at Observatorio del Teide and was corrected after the first nine months of observations. The affected images are discarded following a procedure described in Section \\textit{\\nameref{sec: Data selection}}. \n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig17.png}\n\\caption{Zoom on a standard image (left) and on one with elongated PSFs due to a mount defect (right). This behavior is seen only for a small fraction of the data taken at Observatorio del Teide.}\n\\label{fig: elongated fwhm}\n\\end{figure}\n\n\n\n\\subsection{Interferences}\n\n\nAnother defect is the presence of electrical interferences that create divisions within a strip, each with a specific background level, usually separated by sharp horizontal boundaries, and which vary from one strip to another (Figure \\ref{fig: interferences}). This effect is seen for a small fraction of the data taken at Observatorio del Teide and for a very small fraction of the data taken at Vermillion Cliffs Observatory. We solved this effect by improving the electrical isolation of the systems. No clear consequence is identified on the photometry probably because the excess counts are removed while subtracting the sky background, so we keep these data as they are.\n\n\n\\begin{figure}\n\\includegraphics[width=6.15cm]{crouzet_fig18.png}\\hfill\n\\includegraphics[width=5.4cm]{crouzet_fig19.png}\n\\caption{Example of images affected by electrical interferences at the Observatorio del Teide. Left: effect on the strips taken during the night of February 25, 2013. Some strips are affected at the beginning and end of the night. The scale in ADU is indicated at the bottom. Right: effect on a square image (obtained after carving the parent strip) and projection along the $y$ direction downwards.}\n\\label{fig: interferences}\n\\end{figure}\n\n\n\n\\subsection{Photometric precision}\n\\label{sec: Photometric precision}\n\n\nTo evaluate the photometric precision of the instruments, we calculate the standard deviation of the lightcurves over all the observations and report them in a RMS -- magnitude diagram. An example is shown in Figure \\ref{fig: RMS diagram} for one camera and one field. We obtain a precision between 1\\% and 3\\% for stars of magnitude 9 to 12.5 considering all the exposures, on a timescale of 6.3 minutes. This precision is twice larger than the theoretical prediction, on average. We also calculate the point-to-point RMS which does not consider correlated noise. The point-to-point RMS agrees well with the theoretical predictions for stars fainter than magnitude 11 and is slightly larger for brighter stars. The difference between the true and point-to-point RMS indicates that lightcurves are dominated by correlated noise, which is often the case for ground-based time-series photometric observations. For XO, the large amount of collected data by six different cameras from three different sites on the same fields averages these variations and reduces the RMS drastically when folding the lightcurves at given periods, as shown in Figures \\ref{fig: xo poster}, \\ref{fig: lightcurves}, and \\ref{fig: lightcurves long}. For example, after folding the lightcurve of the hot Jupiter XO-6b at the orbital period of the planet, the RMS is 1.3 mmag on a 30 min timescale (the ingress and egress duration) and 0.8 mmag on a 1.5 hour timescale (half the transit duration). \n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig20.png}\n\\caption{RMS -- magnitude diagram for the stars of field 00 obtained from one camera at Vermillion Cliffs Observatory over the full duration of the observations (twice nine months). Each point represents the standard deviation of the lightcurve as a function of the stellar $R$ magnitude. The true RMS (black dots) and the point-to-point RMS (red dots) are calculated from the full unbinned lightcurves with one point taken every 6.3 minutes on average. We also show the point-to-point RMS on a one-hour timescale obtained after binning the lightcurves with a 10-point boxcar (gray dots). Several noise components computed theoretically for the unbinned lightcurves are indicated: Poisson noise (dashed line), sky background noise (dash-dot line), read-out noise (dotted line), a systematic noise floor arbitrarily set at 0.4\\% (dash-dot-dot line), and total noise (plain line).}\n\\label{fig: RMS diagram}\n\\end{figure}\n\n\n\n\n\\section{Planet search}\n\\label{sec: Planet search}\n\nIn this section, we present the search for transiting exoplanets from the extracted lightcurves. Several steps are necessary: searching for transit signals, selecting viable planet candidates, and confirming or rejecting them through follow-up observations. Then, we compare the number of detected planets with the expected yield. \n\n\n\\subsection{Search for periodic signals}\n\\label{sec: Search for periodic signals}\n\nWe search for periodic signals in the lightcurves using the BLS algorithm \\citep[Box Least Square,][]{Kovacs2002}. This program searches for square shapes in the lightcurves representing transit-like events. Because we are interested in short and long period planets, we perform the search over a wide period range: $0.4 < P < 100$ days. The frequency comb is set by several parameters: the expected transit duration $\\tau$, the total extent of the observations $t_{obs}$ (two years), and the maximum number of lightcurve bins $N_{bin}$ that BLS can handle. The transit duration depends on $P$ but also on the stellar density $\\rho_\\star$. We use the known exoplanet population discovered by the ground-based transit surveys WASP and HAT to bracket the range of stellar densities and adopt a range $0.15 < \\rho_\\star < 5 \\rm \\; g\\,cm^{-3}$. This bracketing is necessary to limit the search to plausible systems.\n\nThe transit duration $\\tau$ varies widely within the period search range so a unique frequency comb is not appropriate (again it would yield too many unphysical detections). We divide the period range into 15 intervals with sizes following approximately a logarithmic scale. For each interval, we calculate the limits of the plausible transit duty cycles $q$, where $q\\, = \\, \\tau \/ P$, and use them to build a specific frequency comb (Figure \\ref{fig: transit duty cycle}).\nEach frequency comb has a constant spacing in $\\Delta f \/ f$, which we set to keep a maximum timing error over $t_{obs}$ as one fourth of the transit duration. We also set the number of lightcurve bins to keep a minimum of four in-transit bins. This ensures that the transit events overlap well with each other when folding the lightcurve. We run the BLS search in each interval separately. During this search, we refine the frequency comb with nine times more frequencies which we apply locally, first around the forty best frequencies, then around the best frequency. For each lightcurve, BLS returns the period that yields the largest power. In addition, we calculate several parameters corresponding to that signal: $q$, $\\varphi_{min}$ and $\\varphi_{max}$ (the minimum and maximum in-transit phases), the transit depth $\\delta$, and $\\alpha$. This latter parameter is equivalent to the signal to noise ratio of the transit: $\\alpha = \\delta \/ \\sigma \\times \\sqrt{N \\times q}$, where $\\sigma$ is the standard deviation of the residual lightcurve after subtracting the transit signal, $N$ is the number of data points, and $N \\times q$ is the number of in-transit data points (assuming that the points are equally distributed).\n \nWe run BLS twice. After the first run, the histogram of identified periods shows several peaks that are mostly centered around integer values below 15 days, which indicate aliases of the day-night cycle. Another peak around 29 days is present and is probably due to the Moon cycle. We define filters manually to exclude periods around these peaks and perform a second run.\n\nFinally, we combine the 15 intervals into 3 period ranges: 0.4 -- 1 day, 1 -- 10 days, and 10 -- 100 days. For each lightcurve, we compare the results obtained for the intervals within each range and keep the period that yields the largest power in the BLS spectrum.\nAll the lightcurves end up with a best period but most of them do not contain a valid signal. We keep only those with $\\alpha > 15$ and sort them by decreasing $\\alpha$ (lightcurves with the most significant signals are ranked first). This yields around 300 lightcurves out of 2000 for each field. We developed an interactive program in $IDL$ to inspect these selected lightcurves and to identify transit candidates, as displayed in Figure \\ref{fig: xo poster}. From this search, we identified hundreds of variable stars and a few tens of transiting planet candidates. Examples of lightcurves of variable objects are shown in Figures~\\ref{fig: lightcurves} and ~\\ref{fig: lightcurves long}.\n\n\nTo evaluate the transit search efficiency, we inject fake transits in the lightcurves. We choose 20 lightcurves of stars of various magnitudes and assign them different periods as well as transit parameters randomly chosen within physically plausible ranges. We also inject the transits of the planets discovered by the WASP and HAT surveys in the lightcurves of stars of similar magnitudes (these planets have periods generally below 10 days). This yields a total of 167 lightcurves with injected transits that we concatenate at the end of the star -- epoch magnitude array, and that we analyze in the same way as regular lightcurves. More than 90\\% of the injected planets are recovered after the BLS search and the transit candidate selection.\n\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig21.pdf}\n\\caption{Transit duty cycle $q$ as a function of orbital period $P$ for different stellar densities $\\rho_\\star$. Values of $\\rho_\\star$ from top to bottom are 0.15, 0.3, 1.41 (Sun), 3, and 5 $\\rm g\\,cm^{-3}$ and are represented by a red dashed, red dotted, green plain, blue dotted, and blue dashed line respectively. We split the period search range into 15 intervals and define the allowed duty cycles in each of them using the extremum values obtained for the stellar densities 0.15 and 5 $\\rm g\\,cm^{-3}$, as indicated by the boxes. A specific frequency comb is built for each interval.}\n\\label{fig: transit duty cycle}\n\\end{figure}\n\n\n\\begin{figure}\n\\includegraphics[width=\\columnwidth]{crouzet_fig22.png}\n\\caption{Image of the interactive interface that is used to analyze lightcurves and identify transit candidates. This example is that of the hot Jupiter XO-6b \\citep{Crouzet2017}. The interface shows the lightcurve folded at the best period (top left) with individual data points in black and a binning in yellow, the BLS spectrum (top right), the full lightcurve (middle left) where the transits are indicated in green, a zoom on the transit (middle right), the centroid variations, a visible and an infrared image of the star's neighborhood, the results of a transit fit, and a list of parameters (BLS outputs, transit parameters, magnitudes, color, spectral type, Simbad information, etc.).}\n\\label{fig: xo poster}\n\\end{figure}\n\n\n\\begin{figure}\n\\includegraphics[width=5.7cm]{crouzet_fig23.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig24.png}\n\\includegraphics[width=5.7cm]{crouzet_fig25.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig26.png}\n\\includegraphics[width=5.7cm]{crouzet_fig27.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig28.png}\n\\includegraphics[width=5.7cm]{crouzet_fig29.png}\\hfill \n\\includegraphics[width=5.7cm]{crouzet_fig30.png}\n\\caption{Lightcurves of eclipsing objects and variable stars with short periods obtained with XO. Individual data points obtained with a time interval of 6.3 minutes are shown in black and a binning is shown in red. The period in days and the $R$ magnitude are indicated on each plot. From left to right and top to bottom: a transiting planet candidate with a 12 day period, a probable eclipsing binary, a semi-detached eclipsing binary, two variable stars, three variable stars with beats (with the lowest period indicated).}\n\\label{fig: lightcurves}\n\\end{figure}\n\n\n\\begin{figure}\n\\includegraphics[width=5.7cm]{crouzet_fig31.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig32.png}\n\\includegraphics[width=5.7cm]{crouzet_fig33.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig34.png}\n\\includegraphics[width=5.7cm]{crouzet_fig35.png}\\hfill\n\\includegraphics[width=5.7cm]{crouzet_fig36.png} \n\\caption{Same as Figure \\ref{fig: lightcurves} for long period objects. From left to right and top to bottom: an eccentric eclipsing binary with a 88 day period, a heartbeat eccentric eclipsing binary with a 20 day period, a 24 day period variable star with amplitude variations, and three variable stars with 115 day, 31 day, and 20 day periods.}\n\\label{fig: lightcurves long}\n\\end{figure}\n\n\n\n\n\\subsection{Follow-up observations}\n\\label{sec: Follow-up observations}\n\nFollow-up observations are necessary to confirm or reject transit candidates. We built a large follow-up team of amateur and professional astronomers who run these observations using facilities reported in Table \\ref{tab: follow-up facilities}. We perform photometry at the predicted transit ephemerides to check the reality of the signals; most candidates are not detected and are rejected. Then, we perform multi-color photometry: we observe transits in different bandpasses to identify eclipsing binaries, which have a color-dependent transit depth. We also compare the odd and even transit depths, which usually differ for eclipsing binaries, and search for secondary eclipses that may be unseen in the XO lightcurves. Then, valid candidates are sent for radial velocity observations with the SOPHIE spectrograph at the Observatoire de Haute-Provence, France \\citep{Bouchy2009}. In practice, follow-up observations are very time consuming, so the candidates are ordered by priority and the type of follow-up observation is carefully chosen. For example, a few candidates show a very clear signal in the XO data and have all the properties of a transiting planet; these are sent directly for radial velocities. Others are sent to multi-color photometry: one transit observed in two colors can be enough to identify an eclipsing binary. Finally, some objects require specific observations. This was the case for the hot Jupiter XO-6b which orbits a fast rotating F5 star: the radial velocity precision was limited to about 70 $\\rm m \\,s^{-1}$ due to the stellar rotation, which was too large to confirm the presence of the planet. We performed Rossiter-McLaughlin observations with SOPHIE and detected the planet's signature in Doppler tomography, which confirmed its planetary nature \\citep{Crouzet2017}. Follow-up observations of other transit candidates are underway.\n\n\n\\begin{table}\n\\begin{center}\n\\caption{Observatories and telescopes used for follow-up observations of XO transit candidates. The diameters of the primary mirrors are given in inches and cm.}\n\\label{tab: follow-up facilities}\n\\begin{tabular}{lll}\n\\hline\n\\hline\nObservatory & Telescope & Purpose \\\\\n\\hline\nObservatoire de Haute-Provence, France & 76\" (193 cm), SOPHIE spectrograph & Radial velocities \\\\\nHereford Arizona Observatory, Arizona, USA & Celestron 11\" (28 cm), Meade 14\" (36 cm) & Photometry \\\\\nActon Sky Portal, Massachusetts, USA & 11\" (28 cm) & Photometry \\\\\nObservatori Astron\\`omic del Montsec, Catalonia, Spain & Joan Or\\'o Telescope 31\" (80 cm) & Photometry \\\\\nObservatoire de Nice, France & Schaumasse 16\" (40 cm) & Photometry \\\\\nVermillion Cliffs Observatory, Kanab, Utah, USA & 24'' (60 cm) & Photometry \\\\\nElgin Observatory, Elgin, Oregon, USA & 12\" (30 cm) & Photometry \\\\\n\\hline\n\\hline\n\\end{tabular}\n\\end{center}\n\\end{table}\n\n\n\\subsection{Planet detection yield and discoveries}\n\\label{sec: Yield and planet detection}\n\nWe estimate the yield of exoplanet discovery with XO and compare it to the number of detected planets. The aim is to inform us on the observation and data analysis efficiency and not to perform an accurate completeness study. Thus, we compute only simple estimates.\nWe use the yield simulations developed by \\citet{Sullivan2015} for the NASA \\textit{TESS}\\xspace mission \\citep[\\textit{Transiting Exoplanet Survey Satellite,}][]{Ricker2014}, which in turn are based off results from the \\textit{Kepler}\\xspace mission. We consider only planets with orbital periods $P < 20$ days, because they are easier to detect due to the decreasing transit probability for longer orbital periods, as evidenced by the WASP \\& HAT planet distributions peaking around 3-4 days.\nThe \\textit{TESS}\\xspace simulations cover 95\\% of the entire sky, \\textit{i.e.}\\xspace 38,000\\ensuremath{^{\\circ2}}\\xspace. The number of transiting planets with $P < 20$ days and radii greater than four times that of Earth ($R_p > 0.4 \\;R_{Jup}$) orbiting stars with Cousins $I$ band magnitude $I_c < 10$ over this area is expected to be \\simi100, and \\simi300 for $I_c < 11$ \\citep[][Figures 11 \\& 22]{Sullivan2015}. The XO units concentrate on 520\\ensuremath{^{\\circ2}}\\xspace (10 fields of 52\\ensuremath{^{\\circ2}}\\xspace each). Thus, the yield of such planets is 1.4 for $I_c < 10$ and 4.1 for $I_c < 11$.\nIn this calculation we assume ideal monitoring, \\textit{i.e.}\\xspace no limitation due to the observing window or instrumental precision. For \\textit{TESS}\\xspace, this is well justified as the monitoring will be continuous for at least 30 days with a precision that is much better than necessary for the detection of transiting giant exoplanets (in other words, \\textit{TESS}\\xspace will detect nearly every planet with those characteristics, at least around the selected target stars). For XO, the observing window per field extends to approximately the nighttime of eighteen months (twice nine months, two strips, two longitudes considering that Observatorio del Teide and Observatori Astron\\`omic del Montsec have a similar longitude), and the precision of the full lightcurves folded at short periods and binned on timescales that are relevant to transits is at the millimagnitude level (see Section \\textit{\\nameref{sec: Photometric precision}} and Figures \\ref{fig: xo poster} and \\ref{fig: lightcurves}). Thus, transiting close-in giant planets should be detected, if present. \nWe found two such planets with the XO units so far, XO-6b \\citep{Crouzet2017} and XO-7b \\citep{Crouzet-inprep}, which host stars have $I_c$ magnitudes of 10.12 and 10.27 respectively. This number is in line with the approximate yield. Thus, the observations and data analysis are as efficient as one could expect. A re-analysis of the XO data to find new transiting close-in gas giant planets would be of low gain and we will simply pursue the follow-up observations of the candidates we have identified. Although transiting planets with longer periods are more challenging to detect, we found a few candidates with $P > 20$ days which are also under follow-up observations.\n\n\n\n\n\\section{Conclusion}\n\\label{sec: Conclusion}\n\nIn this chapter, we presented an overview of the second version of the XO project: instrumental setup, operations, data reduction, instrumental performances, search for transit signals, planet yield and discoveries. We observed two strips covering an effective sky area of $520^{\\circ2}$ for twice nine months using the CCDs in time-delayed integration, and we extracted the lightcurves of \\simi20,000 bright stars up to magnitude $R \\approx 12.5$. The precision is at the millimagnitude level when folding the lightcurves on timescales that are relevant to transits. In addition, this setup allows us to detect long period signals, up to $P \\approx 100$ days. We identified several hundreds of variable stars and a few tens of planet candidates. The transiting hot Jupiter \\mbox{XO-6b} orbiting a fast rotating star has been discovered from this work \\citep{Crouzet2017}. Another planet, XO-7b, has been confirmed \\citep{Crouzet-inprep} and other transit candidates are under follow-up observations. The XO observations have been discontinued in anticipation of the NASA \\textit{TESS}\\xspace mission and the work on XO is now dedicated to the follow-up and study of individual objects.\n\n\n\n\n\\section{Acknowledgments}\nN.C. gratefully acknowledges Peter R. McCullough as the founder and Principal Investigator of the XO project. The XO project was supported by NASA grant NNX10AG30G.\nThis research has made use of the Extrasolar Planets Encyclopaedia (exoplanet.eu) and the Simbad database (simbad.u-strasbg.fr\/simbad\/).\nSoftware: astrometry.net \\citep{Lang2010}, Stellar Photometry Software \\citep{Janes1993}.\n\n\n\n\\bibliographystyle{spbasicHBexo}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nFor real values of $\\,s$, $s>1$, the Riemann zeta function is defined as $\\,\\zeta(s) := \\sum_{n=1}^\\infty{{\\,1\/n^s}}$.\\footnote{In this domain, this series converges according to the integral test. For $\\,s=1$, one has the harmonic series $\\:\\sum_{n=1}^\\infty{1\/n}$, which diverges to infinity.} For $\\,s=2k$, $\\,k \\in \\mathbb{Z}$, $k>0$, Euler (1740) did find that~\\cite{Euler} \n\\begin{equation}\n\\zeta(2k) = \\frac{2^{2k-1} \\, \\left|B_{2k}\\right|}{(2k)!} \\, \\: \\pi^{2k} \\, ,\n\\label{eq:Euler}\n\\end{equation}\nwhere $\\,B_{k}\\,$ is the $k$-th Bernoulli number.\\footnote{The (rational) numbers $B_{k}$ are the coefficients of ${\\,z^k\/k!}\\,$ in the Taylor series expansion of ${\\,z\/(e^z-1)}$, $|z| < 2\\,\\pi$.}\n\nAs a consequence, since $\\,B_2 = 1\/6\\,$ one has $\\,\\zeta(2) = \\pi^2\/6$, which is the Euler solution to the Basel problem (see Ref.~\\cite{Euler0} and references therein).\n\nBy noting that the series expansion approach introduced by Dancs and He (2006) on seeking for an Euler-type formula for $\\,\\zeta{(2k+1)}$, see Ref.~\\cite{Dancs}, could be modified in a manner to furnish similar formulas for $\\,\\zeta(2k)$, here in this note I show that the substitution of $\\,\\sin{(n \\pi)}\\,$ by $\\,\\cos{(n \\pi)}\\,$ in the Dancs-He initial series in fact yields a series expansion which can be reduced to a finite sum involving only even zeta values. From the first few terms of this sum, I have found an elementary proof of $\\,\\zeta{(2)} = {\\,\\pi^2\/6}\\,$ and a recurrence formula for $\\zeta{(2k)}$. The proofs are elementary in the sense they do not involve complex analysis, Fourier series, or multiple integrals.\\footnote{For non-elementary proofs, see, e.g., Refs.~\\cite{Kalman,Apostol} and references therein.}\n\n\n\\section{Elementary evaluation of $\\,\\zeta{(2)}$}\n\nFor any real $\\epsilon > 0$ and $u \\in [1,1+\\epsilon]$, we begin by taking into account the following Taylor series expansion considered by Dancs and He in Ref.~\\cite{Dancs}:\n\\begin{equation}\n\\frac{2 \\, e^t}{e^t + u} = \\sum_{m=0}^{\\infty}{\\phi_m(u) ~ \\frac{t^m}{\\,m!}} \\, ,\n\\label{eq:phiserie}\n\\end{equation}\nwhich converges absolutely for $\\,|t| < \\pi$.\n\nFrom the generating function for the Euler polynomial $\\,E_m(x)$, namely ${\\,2 \\, e^{x\\,t}\/(e^t + 1)} = \\sum_{m=0}^{\\infty}{E_m(x)\\, \\frac{t^m}{m!}}\\,$, it is clear that $\\,\\phi_m(1) = E_m(1)$, for all nonnegative integer values of $\\,m$. For $u>1$, we have\n\\begin{equation}\n\\phi_m(u) = -2 \\, \\sum_{n=1}^{\\infty}{\\frac{n^{\\,m}}{(-u)^{\\,n}}} \\, .\n\\label{eq:phi}\n\\end{equation}\nLet us take this series as our definition of $\\,\\phi_{-m}(u)$, $m$ being a positive integer. Therefore\n\\begin{equation}\n\\phi_{-m}(1) = -2 \\, \\sum_{n=1}^\\infty{\\frac{(-1)^n}{n^m}} = -2 \\: \\zeta^{*}{(m)} = 2\\,(1-2^{1-m})\\,\\zeta{(m)}\n\\label{eq:phinegative}\n\\end{equation}\nfor all integer $\\,m>1$.\n\nNow, let\n\\begin{equation*}\nf(u) := \\sum_{n=1}^{\\infty}{\\frac{\\left({\\,1\/u}\\right)^n}{n^2}}\n\\end{equation*}\nbe an auxiliary function, with $u$ belonging to the same domain as above. Since $\\,\\cos{(n \\pi)} = (-1)^n$, then $\\,f(u)\\,$ can be written in the form\n\\begin{equation*}\nf(u) = \\sum_{n=1}^{\\infty}{(-1)^n \\, \\frac{\\cos(n\\,\\pi)}{u^n\\,n^2}} \\, .\n\\end{equation*}\nOn expanding $\\,\\cos{(n \\pi)}\\,$ in a Taylor series, one has\n\\begin{equation*}\nf(u) = \\sum_{n=1}^{\\infty}{\\left[\\frac{(-1)^n}{u^n\\,n^2} \\cdot \\sum_{j=0}^{\\infty}{(-1)^j \\, \\frac{(n \\pi)^{2 j}}{(2 j)!}}\\right]} = \\sum_{j=0}^{\\infty}{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, \\sum_{n=1}^{\\infty}{(-1)^n \\, \\frac{n^{2 j}}{u^n\\,n^2}}} ,\n\\end{equation*}\nin which the change of sums justifies by Fubini's theorem. By writing the last series in terms of $\\,\\phi_m(u)$, one has\n\\begin{eqnarray}\nf(u) = \\sum_{j=0}^{\\infty}{(-1)^j \\, \\frac{\\pi^{2 j}}{(2 j)!} \\, \\frac{\\phi_{2j-2}(u)}{(-2)}} \\, .\n\\label{eq:marcia}\n\\end{eqnarray}\n\nThis is sufficient for proofing our first result.\n\\newline\n\n\\begin{teo}[Short evaluation of $\\,\\zeta{(2)}\\,$]\n\\label{teo:z2}\n\\begin{equation*}\n\\sum_{n=1}^\\infty{\\frac{1}{\\,n^2}} = \\frac{\\pi^2}{6} \\, .\n\\end{equation*}\n\\end{teo}\n\n\\begin{prova}\n\\; By taking the limit as $u \\rightarrow 1^{+}$ on both sides of Eq.~\\eqref{eq:marcia}, one has\n\\begin{eqnarray}\n\\lim_{u \\rightarrow 1^{+}} \\sum_{n=1}^{\\infty}{\\frac{1}{u^n \\, n^2}} = -\\frac12 \\, \\phi_{-2}(1) \\, + \\frac12 \\, \\frac{\\pi^2}{2!} \\, \\phi_0{(1)} -\\frac12 \\, \\sum_{j=2}^{\\infty}{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, \\phi_{2j-2}(1)} \\, ,\n\\end{eqnarray}\nwhich, in face of the value of $\\phi_{-2}(1)$ stated in Eq.~\\eqref{eq:phinegative}, implies that\n\\begin{equation}\n\\sum_{n=1}^{\\infty}{\\frac{1}{n^2}} = -\\frac12 \\, \\left[ 2 \\left(1-2^{1-2}\\right) \\zeta{(2)} \\right] \\, +\\frac{\\pi^2}{4} \\, E_0{(1)} -\\frac12 \\, \\sum_{j=2}^{\\infty}{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, E_{2j-2}(1)} \\, .\n\\end{equation}\nSince $E_0(1)=1$ and $E_m(1)=0$ for all $m>0$, the right-hand side of this equation reduces to $\\,-\\frac12 \\, \\zeta{(2)} +{\\,\\pi^2 \/ 4}$, which implies that\n\\begin{equation*}\n\\zeta{(2)} = -\\frac12 \\, \\zeta{(2)} +\\frac{\\pi^2}{4} \\, ,\n\\end{equation*}\nand then $\\: \\dfrac{3}{2} \\: \\zeta{(2)} = \\dfrac{\\pi^2}{4}\\,$. \n\\begin{flushright} $\\Box$ \\end{flushright}\n\\end{prova}\n\n\n\\section{Recurrence formula for $\\zeta{(2 k)}$}\n\nInterestingly, our approach can be easily adapted to treat higher even zeta values by changing the exponent of $\\,n\\,$ from $\\,2\\,$ to $\\,2 k$. The result is the following recurrence formula for even zeta values.\n\n\\begin{teo}[Recurrence for $\\,\\zeta{(2 k)}\\,$]\n\\label{teo:z2k}\n\\; For any positive integer $\\,k$,\n\\begin{equation*}\n\\left( 4 -\\frac{4}{2^{2k}} \\right) \\zeta{(2k)} = \\sum_{m=1}^{k-1}{ \\frac{(-1)^{k-m+1}}{(2k-2m)!} \\left( 2 -\\frac{4}{2^{2m}} \\right) \\pi^{2k-2m} \\, \\zeta{(2m)}} \\,-(-1)^k \\frac{\\pi^{2k}}{(2k)!} \\, .\n\\label{eq:recorrencia}\n\\end{equation*}\n\\end{teo}\n\n\\begin{prova}\n\\; We begin by defining $\\,f_k{(u)} := \\sum_{n=1}^{\\infty}{\\left({\\,1\/u}\\right)^n \/ n^{2 k}}\\,$. Again, since $\\,\\cos{(n \\pi)} = (-1)^n$, we may write\n\\begin{eqnarray}\nf_k{(u)} &=& \\sum_{n=1}^{\\infty}{(-1)^n \\, \\frac{\\cos(n\\,\\pi)}{u^n\\,n^{2k}}} = \\sum_{n=1}^{\\infty}{\\frac{(-1)^n}{u^n\\,n^{2k}} \\, \\sum_{j=0}^{\\infty}{(-1)^j \\, \\frac{(n \\pi)^{2 j}}{(2 j)!}}} \\nonumber \\\\\n&=& \\sum_{j=0}^{\\infty}{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, \\sum_{n=1}^{\\infty}{(-1)^n \\, \\frac{n^{2 j}}{u^n\\,n^{2k}}}}\\, .\n\\end{eqnarray}\nOn rewriting the last series in terms of $\\phi_m(u)$, one finds\n\\begin{equation*}\nf_k{(u)} = \\sum_{j=0}^{\\infty}{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, \\frac{\\phi_{2j-2k}(u)}{(-2)}} \\, -\\frac12 \\, \\sum_{j=0}^{k-1}{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, \\phi_{2j-2k}(u)} \\,- \\frac12 \\, \\sum_{j=k}^\\infty{(-1)^j \\frac{\\pi^{2 j}}{(2 j)!} \\, \\phi_{2j-2k}(u)} \\, .\n\\end{equation*}\nNow, on substituting $\\,m = j -k\\,$ in the above series, one has\n\\begin{eqnarray}\nf_k{(u)} = -\\frac12 \\, \\sum_{m=-k}^{-1}{(-1)^{m+k} \\, \\frac{\\pi^{2m+2k}}{(2m+2k)!} \\, \\phi_{2m}(u)} -\\frac12 \\, \\sum_{m=0}^\\infty{(-1)^{m+k} \\, \\frac{\\pi^{2m+2k}}{(2m+2k)!} \\, \\phi_{2m}(u)} \\nonumber \\\\\n= -\\frac12 \\, (-1)^k \\left[ \\, \\sum_{\\widetilde{m}=1}^{k}{ \\frac{(-1)^{\\widetilde{m}}\\,\\pi^{2k-2\\widetilde{m}}}{(2k-2\\widetilde{m})!} \\, \\phi_{-2\\widetilde{m}}(u)}\n+ \\sum_{m=0}^\\infty{\\frac{(-1)^{m} \\, \\pi^{2m+2k}}{(2m+2k)!} \\, \\phi_{2m}(u)} \\right] \\! . \\;\n\\label{eq:2seriesM}\n\\end{eqnarray}\nThe limit as $u \\rightarrow 1^{+}$, taken on both sides of Eq.~\\eqref{eq:2seriesM}, yields\n\\begin{equation}\n\\lim_{u \\rightarrow 1^{+}} \\sum_{n=1}^{\\infty}{\\frac{1}{u^n \\, n^{2k}}} = -\\frac12 \\, (-1)^k \\left[ \\, \\sum_{m=1}^k{ \\frac{(-1)^m \\, \\pi^{2k-2m}}{(2k-2m)!} \\, \\phi_{-2m}(1)}\n+\\sum_{m=0}^\\infty{\\frac{(-1)^m \\, \\pi^{2m+2k}}{(2m+2k)!} \\, \\phi_{2m}(1)} \\right]\\!.\n\\label{eq:aux}\n\\end{equation}\nFrom Eq.~\\eqref{eq:phinegative}, one knows that $\\phi_{-2m}(1) = 2 \\left( 1 -2^{1-2m} \\right) \\zeta{(2m)}$. For nonnegative values of $m$, one has $\\phi_{2m}(1) = E_{2m}(1) = 0$, the only exception being $\\,\\phi_0(1) = E_0(1)=1$. This reduces Eq.~\\eqref{eq:aux} to\n\\begin{equation*}\n\\sum_{n=1}^{\\infty}{\\frac{1}{n^{2k}}} = -(-1)^{k} \\sum_{m=1}^k{\\frac{(-1)^m \\, \\pi^{2k-2m}}{(2k-2m)!} \\, \\left( 1 -2^{1-2m} \\right) \\zeta{(2m)}} -(-1)^k\\, \\frac{\\pi^{2k}}{2\\,(2k)!} \\, .\n\\end{equation*}\nBy extracting the last term of the sum and isolating $\\,\\zeta{(2k)}$, one finds\n\\begin{equation*}\n\\left( 2 -\\frac{2}{2^{2k}} \\right) \\, \\zeta{(2k)} = (-1)^{k+1} \\sum_{m=1}^{k-1}{\\frac{(-1)^m \\, \\pi^{2k-2m}}{(2k-2m)!} \\, \\left( 1 -2^{1-2m} \\right) \\zeta{(2m)}} -(-1)^k\\, \\frac{\\pi^{2k}}{2\\,(2k)!} \\, .\n\\end{equation*}\nA multiplication by $2$ on both sides yields the desired result.\n\\begin{flushright} $\\Box$ \\end{flushright}\n\\end{prova}\n\n\nThe first few even zeta values can be readily obtained from this recurrence formula. For $\\,k=1$, the sum in the right-hand side is null and one has\n\\begin{equation*}\n3 \\,\\zeta(2) = -(-1)\\,\\frac{\\pi^2}{2} \\, ,\n\\end{equation*}\nwhich simplifies to $\\,\\zeta(2) = \\pi^2\/6$, in agreement to Theorem~\\ref{teo:z2}.\n\nFor $k=2$, one has\n\\begin{equation*}\n\\frac{15}{4} \\: \\zeta(4) = \\frac{\\pi^2}{2!} \\: \\zeta(2) - \\frac{\\pi^4}{ 4!} \\, .\n\\end{equation*}\nBy substituting the value of $\\,\\zeta(2)$, above, and multiplying both sides by $4$, one finds\n\\begin{equation}\n15 \\: \\zeta(4) = \\frac{\\pi^4}{3} -\\frac{\\pi^4}{6} = \\frac{\\pi^4}{6} \\, ,\n\\label{eq:z4}\n\\end{equation}\nwhich implies that $\\,\\zeta(4) = \\pi^4 \/ 90$.\n\nNote that, by writing the recurrence formula in Theorem~\\ref{teo:z2k} in the form\n\\begin{equation}\n\\left( 1 -\\frac{1}{2^{2k}} \\right) \\, \\frac{\\zeta{(2k)}}{\\pi^{2k}} = \\sum_{m=1}^{k-1}{ \\, \\frac{(-1)^{k-m+1}}{(2k-2m)!} \\left( \\frac12 -\\frac{1}{2^{2m}} \\right) \\frac{\\zeta{(2m)}}{\\pi^{2m}}} -\\frac{(-1)^k}{4\\,(2k)!} \\, ,\n\\end{equation}\nit is straightforward to show, by induction on $k$, that the ratio ${\\, \\zeta{(2k)} \/ \\pi^{2k}}\\,$ is a rational number for every positive integer $\\,k$, without making use of Euler's formula for $\\zeta{(2k)}$, see Eq.~\\eqref{eq:Euler}, and Bernoulli numbers. In fact, this was the original motivation that has led the author to study the properties of the Dancs-He series expansions. The proofs developed here could well be modified to cover other special functions of interest in analytic number theory.\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nGoal of this paper is to provide a\none-dimensional symmetry result for a phase transition\nequation in a genuinely nonlocal regime in three spatial dimensions.\nThat is, we consider a fractional Allen-Cahn equation of the type\n\\begin{equation}\\label{ALLEN}\n(-\\Delta)^s u=u-u^3\\end{equation} in~$\\mathbb{R}^3$, with\n$s\\in\\left(0,\\frac12\\right)$, and, under boundedness and monotonicity\nassumptions, we prove that~$u$ depends only on one variable, up to a rotation.\n\nIn this setting,\nas customary, for~$s\\in(0,1)$, we consider the fractional Laplace operator\ndefined by\n$$ (-\\Delta)^s u(x):=c_{n,s}\\int_{\\mathbb{R}^n}\\frac{2u(x)-u(x+y)-u(x-y)}{|y|^{n+2s}}\\,dy,$$\nwith\n\\begin{equation}\\label{cns} c_{n,s}:=\\frac{2^{2s-1}\\,s\\,\\Gamma\\left( \\frac{n}2+s\\right)}{\n\\pi^{\\frac{n}2}\\,\\Gamma(1-s)},\\end{equation}\nbeing~$\\Gamma$ the Euler's Gamma Function.\n\nMoreover, we say that~$u$ is $1$D if there exist~$u_o:\\mathbb{R}\\to\\mathbb{R}$ and~$\\omega_o\\in\nS^{n-1}$ such that~$u(x)=u_o(\\omega_o\\cdot x)$ for any~$x\\in\\mathbb{R}^n$.\nThen, our main result in this paper is the following:\n\n\\begin{theorem}\\label{MAIN}\nLet~$n\\le3$, $s\\in\\left(0,\\frac12\\right)$ and~$u\\in C^2(\\mathbb{R}^n,[-1,1])$\nbe a solution of~$(-\\Delta)^s u=u-u^3$ in~$\\mathbb{R}^n$, \nwith~$\\partial_{x_n}u>0$ in~$\\mathbb{R}^n$.\nThen, $u$ is $1$D.\n\\end{theorem}\n\nRecently, a result similar to that in Theorem~\\ref{MAIN} has been established\nin Theorem~1.4 of~\\cite{DSVarxiv}, under the additional assumption that\n\\begin{equation} \\label{LIMIT}\n\\lim_{x_n\\to-\\infty} u(x',x_n)=-1\\quad{\\mbox{ and }}\\quad\n\\lim_{x_n\\to+\\infty} u(x',x_n)=1.\\end{equation}\nTherefore, Theorem~\\ref{MAIN} here\nis the extension of Theorem~1.4 of~\\cite{DSVarxiv}\nin which it is not necessary to assume the limit condition~\\eqref{LIMIT}.\\medskip\n\nWe recall that equation~\\eqref{ALLEN}\nrepresents a phase transition subject to long-range interactions, see e.g.\nChapter~5 in~\\cite{bucur} for a detailed description of the model.\nIn particular, the states~$u=-1$ and~$u=1$ would correspond to the\n``pure phases'' and equation~\\eqref{ALLEN} models\nthe coexistence between intermediate phases and studies\nthe separation between them. \n\nAt a large scale, the separation between phases\nis governed by the minimization of a limit interface, which can be either\nof local or nonlocal type,\nin dependence of the fractional parameter~$s\\in(0,1)$,\nwith a precise bifurcation occurring at\nthe threshold~$s=\\frac12$. \nMore precisely, as proved in~\\cite{MR2948285, MR3133422},\nif~$u$ is a local energy minimizer for equation~\\eqref{ALLEN}\nand~$u_\\varepsilon(x):=u(x\/\\varepsilon)$, as~$\\varepsilon\\searrow0$\nwe have that~$u_\\varepsilon$ approaches a ``pure phase''\nstep function with values in~$\\{-1,1\\}$. That is, we can write, up to subsequences,\n$$ \\lim_{\\varepsilon\\searrow0} u_\\varepsilon = \\chi_E-\\chi_{\\mathbb{R}^n\\setminus E},$$\nand the set~$E$ possesses a minimal interface criterion, depending on~$s$.\nMore precisely, in the ``weakly nonlocal regime'' in which~$s\\in\\left[\\frac12,1\\right)$,\nthe set~$E$ turns out to be a local minimizer for the classical\nperimeter functional: in this sense, on a large scale, the\nweakly nonlocal regime is indistinguishable with respect to the classical case\nand, in spite of the fractional nature of equation~\\eqref{ALLEN},\nits limit interface behaves in a local fashion\nwhen~$s\\in\\left[\\frac12,1\\right)$.\n\nConversely, in the ``genuinely nonlocal regime''\nin which~$s\\in\\left( 0,\\frac12\\right)$,\nthe set~$E$ turns out to be a local minimizer for the nonlocal perimeter functional\nwhich was introduced in~\\cite{MR2675483}. That is, the interface\nof long-range phase transitions when~$s\\in\\left(0, \\frac12\\right)$\npreserves its nonlocal features at any arbitrarily large scale,\nand, as a matter of fact, the scaling properties of the associated\nenergy functional preserve this nonlocal character as well.\nNeedless to say, the persistence of the nonlocal\nproperties at any scale and the somehow unpleasant scaling of the associated\nenergies provide a number of difficulties in the analysis\nof long-range phase coexistence models.\\medskip\n\nIn particular, symmetry properties\nof the solutions of equation~\\eqref{ALLEN}\nhave been intensively studied, also in view of a celebrated\nconjecture by E. De Giorgi in the classical case, see~\\cite{DG}.\nThis classical conjecture asks whether or not bounded and monotone\nsolutions of phase transitions equations are necessarily $1$D.\nIn the fractional framework, a positive answer\nto this problem was known in dimension~$2$\n(see~\\cite{MR2177165} for the case~$s=\\frac12$\nand~\\cite{SV09, cabre-TAMS, MR3035063} for the full range~$s\\in(0,1)$).\nAlso, in dimension~$3$, a positive answer was known\nonly in the weakly nonlocal regimes~$s=\\frac12$\nand~$s\\in\\left(\\frac12,1\\right)$, see~\\cite{MR2644786,MR3148114}.\nSee also~\\cite{2016arXiv161009295S} for a very recent contribution\nabout symmetry results for equation~\\eqref{ALLEN}\nin the weakly nonlocal regime --\nas a matter of fact, the lack of ``good energy estimates''\nprevented the extension of the techniques of these articles\nto the strongly nonlocal regime~$s\\in\\left(0,\\frac12\\right)$.\nIn this sense, our Theorem~\\ref{MAIN} aims at overcoming\nthese difficulties, by relying on the very recent paper~\\cite{DSVarxiv},\nwhich has now taken into account the weakly nonlocal regime\nfor equation~\\eqref{ALLEN}.\\medskip\n\nAfter this work was completed we have \nalso received a preliminary version of the article \\cite{cabre}, \nin which symmetry results for fractional Allen-Cahn equations \nwill be obtained also in the setting of stable solutions.\n\\medskip\n\nThe rest of the paper is organized as follows.\nIn Section~\\ref{SEC:1}, we recall the notion of local minimizers\nand we introduce an equivalent minimization problem in an extended space:\nthis part is rather technical, but absolutely non-standard,\nsince the lack of decay of our solution and the strongly nonlocal condition~$s\\in\\left(0,\\frac12\\right)$\nmake the energy diverge, hence the standard extension methods\nare not available in our case and we will need to introduce a suitable energy\nrenormalization procedure.\n\nIn Section~\\ref{SEC:2}, we relate stable and minimal solutions in the one-dimensional\ncase, by relying also on some layer solution theory of~\\cite{cabre-sire-AIHP}.\n\nIn Section~\\ref{CLASS:A} we consider the profiles of the solution\nat infinity and we establish their minimality and symmetry properties.\n\nIn Section~\\ref{CLASS:B}, we discuss the minimization properties\nunder perturbation which do not overcome the limit profiles,\nand in Section~\\ref{CLASS:C} we recover the minimality of a solution from that\nof its limit profiles. The proof of Theorem~\\ref{MAIN}.\nis contained in Section~\\ref{CLASS:D}.\n\nFor completeness, in Section~\\ref{UL810} we also provide a variant of Theorem~\\ref{MAIN}\nthat gives minimality and symmetry results under the assumption that the limit profiles\nare two-dimensional.\n\\medskip\n\nIt is worth to point out that the setting in Sections~\\ref{SEC:1}--\\ref{CLASS:C}\nis very general and it applies to all the fractional powers~$s\\in(0,1)$, hence\nit can be seen as a useful tool to deal with a class of problems also in extended spaces,\nso to recover minimal properties of the solution from some knowledge of the limit profiles.\n\n\n\\section{Local minimizers in $\\mathbb{R}^n$ and extended local minimizers in~$\\mathbb{R}^{n+1}_+$}\\label{SEC:1}\n\nEquation~\\eqref{ALLEN} lies in the class of semilinear fractional equations of the type\n\\begin{equation}\\label{ALLEN-GEN} (-\\Delta)^s u=f(u).\\end{equation}\n{F}rom now on,\nwe will denote by~$f$ a bistable nonlinearity, namely,\nwe assume that~$f(-1)=f(1)=0$, and there exist~$\\kappa > 0$\nand~$c_\\kappa > 0$ such that~$ f'(t)<-c_\\kappa$ for\nany~$t\\in[-1,-1+\\kappa]\\cup[1-\\kappa,1]$. We also assume that\n\\begin{equation} \\label{BISTABLE}\n\\int_0^1 f(\\sigma)\\,d\\sigma>0>\\int_{-1}^0 f(\\sigma)\\,d\\sigma\n.\\end{equation}\nTo ensure that the solution is sufficiently regular\nin our computations, we assume that\n\\begin{equation}\\label{QUE}\nf\\in C^{1,\\alpha}_{\\rm loc}(\\mathbb{R})\\end{equation}\nwith~$\\alpha\\in(0,1)$\nand~$\\alpha>1-2s$.\nWe remark that, in view\nof~\\eqref{QUE} here and Lemma~4.4 in~\\cite{cabre-sire-AIHP}, bounded\nsolutions of~\\eqref{ALLEN-GEN} are automatically in~$C^2(\\mathbb{R}^n)$,\nwith bounded second derivatives.\n\nThe prototype for such bistable nonlinearity is, of course, the case in which~$f(t)=t-t^3$.\nAlso,\nto describe the energy framework of nonlocal phase transitions, given~$s\\in(0,1)$, $v:\\mathbb{R}^n\\to\\mathbb{R}$\nand~$\\omega\\subset\\mathbb{R}^n$, we consider the functional\n$$ {\\mathcal{F}}_\\omega(v):=\n\\frac{c_{n,s}}{2}\\iint_{Q_\\omega}\\frac{|v(x)-v(y)|^2}{|x-y|^{n+2s}}\\,dx\\,dy+\\int_\\omega\nF(v(x))\\,dx,$$\nwhere\n$$ Q_\\omega:= (\\omega\\times\\omega)\\cup\n(\\omega\\times\\omega^c)\\cup(\\omega^c\\times\\omega)$$\nand\n$$ F(t):= -\\int_{-1}^t f(\\tau)\\,d\\tau.$$\n\n\\begin{definition}\\label{D:L} We say that~$u$ is a local minimizer\nif, for any~$R>0$ and any~$\\varphi\\in C^\\infty_0(B_R)$, it holds that~$\n{\\mathcal{F}}_{B_R}(u)\\le{\\mathcal{F}}_{B_R}(u+\\varphi)$.\n\\end{definition}\n\nNow we describe an extended problem\nand relate its local minimization to the one in Definition~\\ref{D:L}\n(see~\\cite{caffasil}).\nFor this, we set~$a:=1-2s\\in(-1,1)$ and~$\\mathbb{R}^{n+1}_+:=\\mathbb{R}^n\\times(0,+\\infty)$.\nThen, given any~$V:\\mathbb{R}^{n+1}_+\\to\\mathbb{R}$ and~$\\Omega\\subset\\mathbb{R}^{n+1}$, we define\n$$ {\\mathcal{E}}_\\Omega(V):=\n\\frac{\\tilde c_{n,s}}{2}\\int_{\\Omega^+}z^a |\\nabla V(x,z)|^2\\,dx\\,dz+\\int_{\\Omega_0}\nF(V(x,0))\\,dx,$$\nwhere~$\\Omega^+:=\\Omega\\cap\\mathbb{R}^{n+1}_+$ and~$\\Omega_0:=\\Omega\\cap\\{z=0\\}$.\nHere, we used the notation~$\n(x,z)\\in\\mathbb{R}^n\\times(0,+\\infty)$ to denote the variables of~$\\mathbb{R}^{n+1}_+$ and~$\\tilde c_{n,s}>0$\nis a normalization constant.\nGiven~$R>0$, we also denote\n\\begin{eqnarray*}&&{\\mathcal{B}}_R:= B_R\\times(-R,R)=\\big\\{(x,z)\\in\n\\mathbb{R}^{n}\\times\\mathbb{R} {\\mbox{ s.t. }} x\\in B_R {\\mbox{ and }} |z|0\\}=\nB_R\\times(0,R)=\\big\\{(x,z)\\in\n\\mathbb{R}^{n}\\times\\mathbb{R} {\\mbox{ s.t. }} x\\in B_R {\\mbox{ and }} z\\in(0,R)\\big\\}.\\end{eqnarray*}\nIn this setting, we have the following notation:\n\\begin{definition}\\label{D:L:2} We say that~$U$ is an extended local minimizer\nif, for any~$R>0$ and any~$\\Phi\\in C^\\infty_0({\\mathcal{B}}_R)$, it holds that~$\n{\\mathcal{E}}_{{\\mathcal{B}}_R}(U)\\le{\\mathcal{E}}_{{\\mathcal{B}}_R}(U+\\Phi)$.\n\\end{definition}\n\nThe reader can compare Definitions~\\ref{D:L} and~\\ref{D:L:2}. Also, given~$v\\in L^\\infty(\\mathbb{R}^n)$,\nwe consider the $a$-harmonic extension of~$v$ to~$\\mathbb{R}^{n+1}_+$ as the function~$E_v:\n\\mathbb{R}^{n+1}_+\\to\\mathbb{R}$\nobtained by convolution with the Poisson kernel of order~$s$. More explicitly, we set\n\\begin{equation}\\label{0djhicseodfeeyyeyeye} P(x,z):= \\bar c_{n,s} \\frac{z^{2s}}{(|x|^2+z^2)^{\\frac{n+2s}2}}.\\end{equation}\nIn this framework, $\\bar c_{n,s}$ is a positive normalization constant such that\n$$ \\int_{\\mathbb{R}^n} P(x,z)\\,dx=1,$$\nsee e.g.~\\cite{MR3461641}. Then we set\n$$ E_v(x,z):=\\int_{\\mathbb{R}^n} P(x-y,z)\\,v(y)\\,dy=\n\\int_{\\mathbb{R}^n} P(y,z)\\,v(x-y)\\,dy.$$\nWe remark that when~$v\\in C^\\infty_0(\\mathbb{R}^n)$, the function~$E_v$ can also be obtained\nby\nminimization\nof the associated Dirichlet energy, namely\n\\begin{equation} \\label{DIRmi}\n\\inf_{{V\\in C^\\infty_0(\\mathbb{R}^{n+1})}\\atop{V(x,0)=v(x)}}\n\\int_{\\mathbb{R}^{n+1}_+}z^a |\\nabla V(x,z)|^2\\,dx\\,dz\n=\n\\int_{\\mathbb{R}^{n+1}_+}z^a |\\nabla E_v(x,z)|^2\\,dx\\,dz,\\end{equation}\nsee Lemma 4.3.3 in~\\cite{bucur}. Nevertheless,\nwe want to consider here the more general framework in which~$v$\nis bounded, but not necessarily decaying at infinity, and this will produce a number\nof difficulties, also due to the lack of ``good'' functional settings.\n\nWe also remark that the setting in~\\eqref{DIRmi} and the normalization\nconstant~$\\tilde c_{n,s}$\nare compatible with the choice of the constant in~\\eqref{cns},\nsince, for any $v\\in C^\\infty_0(\\mathbb{R}^n)$,\n\\begin{equation}\\label{COMP:AK}\n\\frac{c_{n,s}}{2}\\iint_{\\mathbb{R}^{2n}}\\frac{|v(x)-v(y)|^2}{|x-y|^{n+2s}}\\,dx\\,dy=\n\\frac{\\tilde c_{n,s}}{2}\\int_{\\mathbb{R}^{n+1}_+}z^a |\\nabla E_v(x,z)|^2\\,dx\\,dz,\n\\end{equation}\nsee e.g. formula~(4.3.15) in~\\cite{bucur}.\n\nSince these normalization constants will not play any role in the following computations,\nwith a slight abuse of notation, for the sake of simplicity,\nwe just omit them in the sequel.\n\nIn our setting, for functions~$v$ with no decay at infinity,\nformula~\\eqref{COMP:AK} does not make sense, since both the terms could diverge.\nNevertheless, we will be able to overcome this difficulty by an energy\nrenormalization procedure, based on the formal substraction of the infinite energy.\nThe rigorous details of this procedure are discussed in the\nfollowing\\footnote{We observe that Proposition~\\ref{0OAPQO182:P}\nhere is also related to the extension method in Lemma~7.2\nin~\\cite{MR2675483}, where suitable trace and extended energies\nare compared in the unit ball:\nin a sense, since\nProposition~\\ref{0OAPQO182:P} here\ncompares energies defined in the whole of the space, it can be viewed\nas a ``global'', or ``renormalized'', version of\nLemma~7.2\nin~\\cite{MR2675483}. }\nresult:\n\n\\begin{proposition}\\label{0OAPQO182:P}\nLet~$n\\ge2$ and~$s\\in(0,1)$.\nFor any~$v\\in W^{2,\\infty}(\\mathbb{R}^n)$\nand any~$\\varphi\\in C^\\infty_0(\\mathbb{R}^n)$, it\nholds that\n\\begin{eqnarray*} +\\infty&>& \n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v+\\varphi)(x)-(v+\\varphi)(y)|^2-\n|v(x)-v(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\\\\&\n=& \\lim_{R\\to+\\infty} \\inf_{{\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n.\\end{eqnarray*}\n\\end{proposition}\n\n\\begin{proof} For concreteness, we\nsuppose that the support of~$\\varphi$\nlies in~$B_1$.\nWe take~$\\tau\\in C^\\infty_0(B_2,\\,[0,1])$,\nwith~$\\tau=1$ in~$B_1$, and we let\n\\begin{equation}\\label{v k} \\tau_k(x):=\\tau\\left(\\frac{x}{k}\\right)\n\\quad{\\mbox{ and }}\\quad\nv_k(x):=\\tau_k(x)\\,v(x).\\end{equation}\nWe also set~$w_k:= v-v_k$.\nIn this way, $w_k$ is bounded, uniformly\nLipschitz and\nvanishes in~$B_k$.\nIn particular,\n\\begin{equation}\\label{9:PAA}\n\\frac{\\big|w_k(x)-w_k(y)\\big|\\,\\big|\n\\varphi(x)-\\varphi(y)\\big|\n}{|x-y|^{n+2s}}\\le\n\\frac{C\\,\\min\\{1,|x-y|\\}\\,\\big|\n\\varphi(x)-\\varphi(y)\\big|\n}{|x-y|^{n+2s}}\\in L^1(\\mathbb{R}^{2n}).\\end{equation}\nFurthermore, we have that\n\\begin{eqnarray*}\n&& \\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v+\\varphi)(x)-(v+\\varphi)(y)|^2-\n|v(x)-v(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&&\\qquad\n-\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v_k+\\varphi)(x)-(v_k+\\varphi)(y)|^2-\n|v_k(x)-v_k(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\n\\\\ &=&2\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{\\big(v(x)-v(y)\\big)\\big(\n\\varphi(x)-\\varphi(y)\\big)\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&&\\qquad\n-2\\iint_{\\mathbb{R}^{2n}}\n\\frac{\\big(v_k(x)-v_k(y)\\big)\\big(\n\\varphi(x)-\\varphi(y)\\big)\n}{|x-y|^{n+2s}}\\,dx\\,dy\n\\\\ &=&\n2\\iint_{\\mathbb{R}^{2n}}\n\\frac{\\big(w_k(x)-w_k(y)\\big)\\big(\n\\varphi(x)-\\varphi(y)\\big)\n}{|x-y|^{n+2s}}\\,dx\\,dy\n.\\end{eqnarray*}\nThis, \\eqref{9:PAA}\nand the Dominated Convergence\nTheorem give that\n\\begin{equation}\\label{K9:00:01}\n\\begin{split}&\n\\lim_{k\\to+\\infty}\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v_k+\\varphi)(x)-(v_k+\\varphi)(y)|^2-\n|v_k(x)-v_k(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\n\\\\&\\qquad=\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v+\\varphi)(x)-(v+\\varphi)(y)|^2-\n|v(x)-v(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\n.\\end{split}\\end{equation}\nAlso,\n$$ \\frac{\\big|v(x)-v(y)\\big|\\,\\big|\n\\varphi(x)-\\varphi(y)\\big|\n}{|x-y|^{n+2s}}\\le\n\\frac{C\\,\\min\\{1,|x-y|\\}\\,\\big|\n\\varphi(x)-\\varphi(y)\\big|\n}{|x-y|^{n+2s}}\\in L^1(\\mathbb{R}^{2n}),$$\nand therefore\n\\begin{equation}\\label{K9:00:02}\n\\begin{split}&\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v+\\varphi)(x)-(v+\\varphi)(y)|^2-\n|v(x)-v(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&\\qquad=\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{\n\\big(v(x)-v(y)\\big)\\,\\big(\n\\varphi(x)-\\varphi(y)\\big)\n}{|x-y|^{n+2s}}\\,dx\\,dy+\n\\iint_{\\mathbb{R}^{2n}}\\frac{|\\varphi(x)-\\varphi(y)|^2}{|x-y|^{n+2s}}\\,dx\\,dy\n<+\\infty\n.\\end{split}\\end{equation}\nOn the other hand, recalling~\\eqref{0djhicseodfeeyyeyeye}, \n\\[\nz^a |\\nabla P(x,z)|\\le\n\\frac{Cz(|x|+z)}{(|x|^2+z^2)^{\\frac{n+2s+2}{2}}}\n+\\frac{C}{(|x|^2+z^2)^{\\frac{n+2s}{2}}}\n\\le\n\\frac{C}{(|x|^2+z^2)^{\\frac{n+2s}{2}}}\n.\\]\nThis implies that\n\\begin{eqnarray*}\n&& z^a |\\nabla E_\\varphi(x,z)|=\n\\left| z^a \\int_{\\mathbb{R}^n} \\nabla\nP(x-y,z)\\,\\varphi(y)\\,dy\n\\right|\\\\&&\\qquad\\le\nC \\int_{B_1} |\\nabla P(x-y,z)|\\,dy\n\\le C\\int_{B_1}\n\\frac{dy}{(|x-y|^2+z^2)^{\\frac{n+2s}{2}}}\n.\\end{eqnarray*}\nHence, for any~$x\\in\\mathbb{R}^n\\setminus B_2$,\n\\begin{equation}\\label{HENCE1}\nz^a |\\nabla E_\\varphi(x,z)|\n\\le C\\int_{B_1}\n\\frac{dy}{(|x|^2+z^2)^{\\frac{n+2s}{2}}}\n= \\frac{C}{(|x|^2+z^2)^{\\frac{n+2s}{2}}}\n\\end{equation}\nand, if~$x\\in B_2$,\n\\begin{equation}\\label{HENCE2}\nz^a |\\nabla E_\\varphi(x,z)|\\le\nC\\int_{B_1}\n\\frac{dy}{(0+z^2)^{\\frac{n+2s}{2}}}\n\\le\n\\frac{C}{z^{n+2s}}\n.\\end{equation}\nWe also notice that\n\\begin{equation}\\label{6bis}\\begin{split}& z^a\\,|\\nabla E_\\varphi(x,z)|\\le\nz^a\\,\\left| \\int_{B_1} P(y,z)\\,\\nabla\\varphi(x-y)\\,dy\\right|\n\\\\&\\qquad\\le C z^a\\int_{\\mathbb{R}^n} P(y,z)\\,dy\\le Cz^a\\in L^1(B_2\\times(0,2)).\\end{split}\\end{equation}\nIn addition, $\\| E_{v_k}\\|_{L^\\infty(\\mathbb{R}^{n+1}_+)}\\le\n\\|v_k\\|_{L^\\infty(\\mathbb{R}^n)}$;\nas a consequence\nof these observations, setting\n\\begin{equation}\\label{elle R}\nL_R:=\n(\\partial B_R)\\times (0,R)\\quad{\\mbox{\nand }}\\quad U_R:= B_R\\times\\{R\\}, \\end{equation}we have that\n\\begin{equation}\\label{889AO}\\begin{split}&\n\\lim_{R\\to+\\infty}\n\\int_{L_R\\cup U_R }\nz^a |E_{v_k}(x,z)|\\,|\n\\partial_\\nu \nE_{\\varphi}(x,z)|\\,d{\\mathcal{H}}^{n}(x,z)\n\\le\\lim_{R\\to+\\infty}\n\\int_{L_R\\cup U_R }\n\\frac{C\\,d{\\mathcal{H}}^{n}(x,z)}{\n(|x|^2+z^2)^{\\frac{n+2s}{2}}}\n\\\\ &\\qquad\\qquad\\le\\lim_{R\\to+\\infty}\n\\int_{L_R\\cup U_R }\n\\frac{C\\,d{\\mathcal{H}}^{n}(x,z)}{\nR^{n+2s}} \\le \\lim_{R\\to+\\infty}\n\\frac{CR^n}{R^{n+2s}}=0,\n\\end{split}\\end{equation}\nwhere~$\\partial_\\nu$ denotes the\nexternal normal derivative to the boundary\nof~${\\mathcal{B}}_R^+$.\n\nAccordingly,\n\\begin{eqnarray*}\n&& \\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v_k+\\varphi}(x,z)|^2\n-|\\nabla E_{v_k}(x,z)|^2\\big)\\,dx\\,dz\\\\\n&=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{\\varphi}(x,z)|^2\n+2\\nabla E_\\varphi(x,y)\\cdot\n\\nabla E_{v_k}(x,z)\\big)\\,dx\\,dz\n\\\\ &=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a |\\nabla E_{\\varphi}(x,z)|^2\n+2{\\rm div}\\,\\big(z^a\nE_{v_k}(x,y)\n\\nabla E_{\\varphi}(x,z)\\big) \\,dx\\,dz\\\\\n&=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a |\\nabla E_{\\varphi}(x,z)|^2\\,dx\\,dz\n+2\\int_{\\partial {\\mathcal{B}}_R^+ }\nz^a E_{v_k}(x,z)\n\\partial_\\nu \nE_{\\varphi}(x,z)\\,d{\\mathcal{H}}^{n}(x,z)\\\\\n&=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a |\\nabla E_{\\varphi}(x,z)|^2\\,dx\\,dz\n+2\\int_{B_R} v_k(x)\\,(-\\Delta)^s\\varphi(x)\\,dx\n\\\\ &=&\n\\int_{ \\mathbb{R}^{n+1}_+}\nz^a |\\nabla E_{\\varphi}(x,z)|^2\\,dx\\,dz\n+2\\int_{\\mathbb{R}^n} v_k(x)\n\\,(-\\Delta)^s\\varphi(x)\\,dx.\n\\end{eqnarray*}\nSimilarly,\n\\begin{eqnarray*}\n&& \\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_{v}(x,z)|^2\\big)\\,dx\\,dz\\\\\n\\\\&=&\n\\int_{ \\mathbb{R}^{n+1}_+}\nz^a |\\nabla E_{\\varphi}(x,z)|^2\\,dx\\,dz\n+2\\int_{\\mathbb{R}^n} v(x)\n\\,(-\\Delta)^s\\varphi(x)\\,dx.\\end{eqnarray*}\nSince~$|v_k (-\\Delta)^s\\varphi|\\le|v(-\\Delta)^s\\varphi|\\le\\frac{C}{1+|x|^{n+2s}}\\in L^1(\\mathbb{R}^n)$,\nwe thus obtain that\n\\begin{equation}\\label{89:0128esd}\n\\begin{split}\n&\\lim_{k\\to+\\infty}\n\\int_{\\mathbb{R}^{n+1}_+}\nz^a\\big(|\\nabla E_{v_k+\\varphi}(x,z)|^2\n-|\\nabla E_{v_k}(x,z)|^2\\big)\\,dx\\,dz\n\\\\=\\;&\\lim_{k\\to+\\infty}\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v_k+\\varphi}(x,z)|^2\n-|\\nabla E_{v_k}(x,z)|^2\\big)\\,dx\\,dz\n\\\\\n=\\;&\\lim_{k\\to+\\infty}\n\\left[\\int_{ \\mathbb{R}^{n+1}_+}\nz^a |\\nabla E_{\\varphi}(x,z)|^2\\,dx\\,dz\n+2\\int_{\\mathbb{R}^n} v_k(x)\n\\,(-\\Delta)^s\\varphi(x)\\,dx\n\\right]\\\\ =\\;&\n\\int_{ \\mathbb{R}^{n+1}_+}\nz^a |\\nabla E_{\\varphi}(x,z)|^2\\,dx\\,dz\n+2\\int_{\\mathbb{R}^n} v(x)\n\\,(-\\Delta)^s\\varphi(x)\\,dx\\\\\n=\\;&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_{v}(x,z)|^2\\big)\\,dx\\,dz.\n\\end{split}\\end{equation}\nMoreover, from~\\eqref{COMP:AK}, we know that\n\\begin{eqnarray*}\n&& \n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v_k+\\varphi)(x)-(v_k+\\varphi)(y)|^2-\n|v_k(x)-v_k(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v_k+\\varphi}(x,z)|^2\n-|\\nabla E_{v_k}(x,z)|^2\\big)\\,dx\\,dz\n.\\end{eqnarray*}\nPutting together this with~\\eqref{K9:00:01},\n\\eqref{K9:00:02} and~\\eqref{89:0128esd}, we\nconclude that\n\\begin{eqnarray*} +\\infty&>& \n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v+\\varphi)(x)-(v+\\varphi)(y)|^2-\n|v(x)-v(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&=&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz.\\end{eqnarray*}\nTherefore, to complete the proof of the desired claim,\nit remains to show that\n\\begin{equation}\\label{0OAPQO182}\n\\begin{split}\n&\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n\\\\\n=\\;& \\lim_{R\\to+\\infty} \\inf_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n.\\end{split}\\end{equation}\nTo this end,\nwe take~$\\theta\\in C^\\infty_0({\\mathcal{B}}_1,\\,[0,1])$\nwith~$\\theta=1$ in~${\\mathcal{B}}_{1\/2}$\nand we set~$\\theta_R(x,z):=\\theta\\left(\\frac{x}{R},\\frac{z}{R}\\right)$\nand~$\\Phi_R:=E_\\varphi \\theta_R$.\nWe observe that\n$$ |E_\\varphi(x,z)|\\le \\int_{B_1}\n\\frac{C z^{2s}}{(|x-y|^2+z^2)^{\\frac{n+2s}2}}\\,dy$$\nand therefore, if~$|x|\\ge2$,\n\\begin{equation*} \n|E_\\varphi(x,z)|\\le \n\\frac{C z^{2s}}{(|x|^2+z^2)^{\\frac{n+2s}2}}\\le\n\\frac{C z^{2s}}{(1+|x|^2+z^2)^{\\frac{n+2s}2}}\n.\\end{equation*}\nSimilarly, if~$|x|\\le2$ and~$z\\ge1$,\n\\begin{equation*}\n|E_\\varphi(x,z)|\\le\\int_{ B_1}\n\\frac{C z^{2s}}{(0+z^2)^{\\frac{n+2s}2}}\\,dy\\le\n\\frac{C z^{2s}}{(1+|x|^2+z^2)^{\\frac{n+2s}2}},\n\\end{equation*}\nup to renaming~$C$.\nAlso, if~$|x|\\le2$ and~$z\\in(0,2)$,\n$$ |E_\\varphi(x,z)|\\le C\\int_{\\mathbb{R}^n}P(x-y,z)\\,dy=C.\n$$\nIn view of these estimates, we have that\n\\begin{equation}\\label{STIME:EPHI1}\n\\begin{split}& \\int_{\\mathbb{R}^{n+1}_+} z^a |E_\\varphi(x,z)|^2\\,dx\\,dz\n\\le\nC+\\int_{\\mathbb{R}^{n+1}_+\\setminus {\\mathcal{B}}_2} z^a |E_\\varphi(x,z)|^2\\,dx\\,dz\\\\\n&\\qquad\\le C+\\int_{\\mathbb{R}^{n+1}_+\\setminus {\\mathcal{B}}_2} \n\\frac{C z^{1+2s}}{(1+|x|^2+z^2)^{{n+2s}}}\n\\,dx\\,dz\\\\&\\qquad\n\\le\nC+\\int_{\\mathbb{R}^{n+1}_+\\setminus {\\mathcal{B}}_2} \n\\frac{C }{(1+|x|^2+z^2)^{{n+s-\\frac12}}}\n\\,dx\\,dz\\le C,\n\\end{split}\\end{equation}\nsince~$2n+2s-1>n+1$.\n\nSimilarly, recalling~\\eqref{HENCE1},\n\\eqref{HENCE2} and~\\eqref{6bis},\n\\begin{equation*}\n\\begin{split}& \\int_{\\mathbb{R}^{n+1}_+} z^a |\\nabla E_\\varphi(x,z)|^2\\,dx\\,dz\n\\\\ \\le\\;& C+\\int_{ \\{(x,z):\\, x\\in B_2,\\, z>1\\}}\nz^{-a} \\big(z^a |\\nabla E_\\varphi(x,z)|\\big)^2\\,dx\\,dz\n+\\int_{ \\{(x,z): \\, x\\in \\mathbb{R}^n\\setminus B_2,\\, z\\in(0,1]\\}}\nz^{-a} \\big(z^a |\\nabla E_\\varphi(x,z)|\\big)^2\\,dx\\,dz\n\\\\&\\qquad+\\int_{ \\{(x,z):\\, x\\in \\mathbb{R}^n\\setminus B_2,\\,z>1\\}}\nz^{-a} \\big(z^a |\\nabla E_\\varphi(x,z)|\\big)^2\\,dx\\,dz\n\\\\ \\le\\;& C+\n\\int_{ \\{(x,z):\\, x\\in B_2,\\, z>1\\}}\n\\frac{C}{z^{2n+1+2s}}\\,dx\\,dz\n+\\int_{ \\{(x,z):\\, x\\in \\mathbb{R}^n\\setminus B_2,\\, z\\in(0,1]\\}}\n\\frac{C}{z^{1-2s} (|x|^2+0)^{n+2s}}\\,dx\\,dz\\\\&\\qquad\n+\\int_{ \\{(x,z):\\, x\\in \\mathbb{R}^n\\setminus B_2,\\,z>1\\}}\n\\frac{C}{(|x|^2+z^2)^{n+2s}}\\,dx\\,dz\n\\\\ \\le\\;&C,\n\\end{split}\n\\end{equation*}\nand therefore\n\\begin{equation}\\label{STIME:EPHI2}\n\\int_{\\mathbb{R}^{n+1}_+\\setminus{\\mathcal{B}}_{R\/2}} z^a |\\nabla E_\\varphi(x,z)|^2\\,dx\\,dz\\le\\delta(R),\n\\end{equation}\nwith~$\\delta(R)$ infinitesimal as~$R\\to+\\infty$.\n\nIn addition,\n\\begin{eqnarray*} \\Big| |\\nabla \\Phi_R|^2-|\\nabla E_\\varphi|^2\\Big|\n&\\le& E_\\varphi^2|\\nabla\\theta_R|^2\n+|\\nabla E_\\varphi|^2 |1-\\theta_R^2|\n+2|E_\\varphi| |\\theta_R|\n|\\nabla E_\\varphi||\\nabla\\theta_R|\\\\\n&\\le& \\frac{C E_\\varphi^2}{R^2 }+\nC|\\nabla E_\\varphi|^2 \\chi_{\\mathbb{R}^{n+1}_+\\setminus {\\mathcal{B}}_{R\/2}}.\n\\end{eqnarray*}\nTherefore, in light of~\\eqref{STIME:EPHI1} and~\\eqref{STIME:EPHI2},\n\\begin{eqnarray*}&& \\left|\n\\int_{\\mathbb{R}^{n+1}_+} z^a |\\nabla \\Phi_R(x,z)|^2\\,dx\\,dz\n-\n\\int_{\\mathbb{R}^{n+1}_+} z^a |\\nabla E_\\varphi(x,z)|^2\\,dx\\,dz\\right|\\\\ &\\le&\n\\frac{C}{R^2}\\int_{\\mathbb{R}^{n+1}_+} z^a\n|E_\\varphi(x,z)|^2\\,dx\\,dz+\nC\\int_{\\mathbb{R}^{n+1}_+\\setminus {\\mathcal{B}}_{R\/2}} z^a\n|\\nabla E_\\varphi(x,z)|^2 \\,dx\\,dz \\\\&\\le&\n\\frac{C}{R^2}+\\delta(R).\n\\end{eqnarray*}\nConsequently, we find that\n\\begin{equation}\\label{PAal203948:PP}\n\\begin{split}\n& \\lim_{R\\to+\\infty} \\inf_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n\\\\ \\le\\;&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi_R)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n\\\\\n=\\;&\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Phi_R(x,z)|^2 +2\\nabla E_v(x,z)\\cdot\\nabla\\Phi_R(x,z)\\big)\\,dx\\,dz\n\\\\ \n\\le\\;&\n\\lim_{R\\to+\\infty} \\left[ \\frac{C}{R^2}+\\delta(R)+\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_\\varphi(x,z)|^2 +2\\nabla E_v(x,z)\\cdot\\nabla\nE_\\varphi(x,z)\\big)\\,dx\\,dz\\right.\\\\&\\qquad\n\\left.+2\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi_R(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right]\\\\\n\\le\\;&\n\\lim_{R\\to+\\infty} \\left[ \\frac{C}{R^2}+\\delta(R)+\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_\\varphi(x,z)|^2 +2\\nabla E_v(x,z)\\cdot\\nabla\nE_\\varphi(x,z)\\big)\\,dx\\,dz\\right.\\\\&\\qquad\\left.\n+2\\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right]\n.\\end{split}\\end{equation}\nNow we claim that\n\\begin{equation}\\label{PAal203948}\n\\lim_{R\\to+\\infty} \\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\left|\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right|\n=0.\n\\end{equation}\nTo check this, we recall the notation in~\\eqref{elle R}\nand observe that\n\\begin{equation}\\label{012owe2eudyf8i:00}\n\\begin{split}\n&\\lim_{R\\to+\\infty}\\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\left|\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right|\\\\\n=\\,&\n\\lim_{R\\to+\\infty}\\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\\left|\n\\int_{ {\\mathcal{B}}_R^+ } {\\rm div}\\,\\Big(\nz^a \\Phi(x,z) \\nabla E_v(x,z)\\Big)\\,dx\\,dz\n-\\int_{ {\\mathcal{B}}_R^+ } {\\rm div}\\,\\Big(\nz^a E_v(x,z)\\nabla E_\\varphi(x,z)\\Big)\\,dx\\,dz\\right|\\\\\n=\\,&\n\\lim_{R\\to+\\infty}\\left| \\int_{B_R} \\varphi(x)\\, (-\\Delta)^s v(x)\\,dx\n- \\int_{B_R} v(x)\\, (-\\Delta)^s \\varphi(x)\\,dx\\right.\\\\\n&\\quad\\qquad\\left.-\\int_{L_R\\cup U_R} \nz^a E_v(x,z)\\partial_\\nu E_\\varphi(x,z)\\,d{\\mathcal{H}}^n(x,z)\\right|\n.\\end{split}\n\\end{equation}\nAlso, as in~\\eqref{889AO}, we have that\n\\begin{equation*}\n\\lim_{R\\to+\\infty}\n\\int_{L_R\\cup U_R }\nz^a |E_{v}(x,z)|\\,|\n\\partial_\\nu \nE_{\\varphi}(x,z)|\\,d{\\mathcal{H}}^{n}(x,z)=0.\n\\end{equation*}\nHence, fixing~$k$ and\nrecalling the notation in~\\eqref{v k}, we derive from~\\eqref{012owe2eudyf8i:00}\nthat\n\\begin{eqnarray*}\n&&\\lim_{R\\to+\\infty}\\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\\left|\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right| \\\\\n&=& \\left|\\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s v(x)\\,dx\n- \\int_{\\mathbb{R}^n} v(x)\\, (-\\Delta)^s \\varphi(x)\\,dx\\right| \\\\\n&=& \\left|\\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s v_k(x)\\,dx\n- \\int_{\\mathbb{R}^n} v_k(x)\\, (-\\Delta)^s \\varphi(x)\\,dx\n\\right.\\\\\n&&\\qquad+\\left. \\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s w_k(x)\\,dx\n- \\int_{\\mathbb{R}^n} w_k(x)\\, (-\\Delta)^s \\varphi(x)\\,dx\\right|\n\\\\ &=&\\left|\n\\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s w_k(x)\\,dx\n- \\int_{\\mathbb{R}^n} w_k(x)\\, (-\\Delta)^s \\varphi(x)\\,dx\\right|.\n\\end{eqnarray*}\nSince~$|(-\\Delta)^s \\varphi(x)|\\le\\frac{C}{1+|x|^{n+2s}}\\in L^1(\\mathbb{R}^n)$,\nwe can take the limit as~$k\\to+\\infty$ and use the Dominated Convergence Theorem,\nto obtain that\n\\begin{equation}\\label{udofjvw9etgierufgh}\n\\begin{split}\n&\\lim_{R\\to+\\infty}\\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\left|\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right|\\\\\n=\\;&\\lim_{k\\to+\\infty}\\left| \\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s w_k(x)\\,dx\n- \\int_{\\mathbb{R}^n} w_k(x)\\, (-\\Delta)^s \\varphi(x)\\,dx\\right|\\\\\n=\\;&\\lim_{k\\to+\\infty}\\left| \\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s w_k(x)\\,dx\\right|\n\\\\ =\\;&\n\\lim_{k\\to+\\infty}\\left| \\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s v(x)\\,dx\n-\\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s (\\tau_k v)(x)\\,dx\\right|\n\\\\ =\\;& \\left|\n\\int_{\\mathbb{R}^n} \\varphi(x)\\, (-\\Delta)^s v(x)\\,dx\n-\n\\lim_{k\\to+\\infty}\\left[ \\int_{\\mathbb{R}^n} \\tau_k(x)\\,\\varphi(x)\\, (-\\Delta)^s v(x)\\,dx\\right.\\right.\\\\ &\n\\qquad\\left.\\left.+\n\\int_{\\mathbb{R}^n} v(x)\\,\\varphi(x)\\, (-\\Delta)^s \\tau_k(x)\\,dx+\n\\int_{\\mathbb{R}^n} \\varphi(x)\\, B(\\tau_k, v)(x)\\,dx\\right]\\right|\n\\\\ =\\;&\\lim_{k\\to+\\infty}\\left|\\frac{1}{k^{2s}}\n\\int_{B_1} v(x)\\,\\varphi(x)\\, (-\\Delta)^s \\tau\\left(\\frac{x}{k}\\right)\\,dx+\n\\int_{\\mathbb{R}^n} \\varphi(x)\\, B(\\tau_k, v)(x)\\,dx\\right|\\\\\n\\le\\;& \\lim_{k\\to+\\infty}\\left[ \\frac{C}{k^{2s}}\n\\int_{B_1} |v(x)|\\,|\\varphi(x)|\\,dx+\n\\int_{\\mathbb{R}^n} |\\varphi(x)|\\, |B(\\tau_k, v)(x)|\\,dx\\right]\n\\\\ =\\;&\\lim_{k\\to+\\infty}\\int_{B_1} |\\varphi(x)|\\, |B(\\tau_k, v)(x)|\\,dx,\n\\end{split}\\end{equation}\nwhere\n$$ B(f,g):= c\\, \\int_{\\mathbb{R}^n} \\frac{(f(x)-f(y))(g(x)-g(y))}{|x-y|^{n+2s}}\\,dy,$$\nfor some~$c>0$, see e.g. page~636 in~\\cite{MR3211862}\nfor such bilinear form.\n\nNotice now that\n\\begin{eqnarray*}\n|B(\\tau_k, v)(x)|&\\le&\nC\\int_{\\mathbb{R}^n} \\frac{\\min\\{ 1,\\frac{|x-y|}{k}\\}\\,\n\\min\\{ 1,|x-y|\\}}{|x-y|^{n+2s}}\\,dy\\\\\n&\\le&\n\\frac{C}{k} \\int_{B_1(x)} \\frac{|x-y|^2}{|x-y|^{n+2s}}\\,dy\n+\n\\frac{C}{k}\\int_{B_k(x)\\setminus B_1(x)} \\frac{|x-y|}{|x-y|^{n+2s}}\\,dy\n+\nC\\int_{\\mathbb{R}^n\\setminus B_k(x)} \\frac{dy}{|x-y|^{n+2s}}\\\\ &\\le&\\frac{C}{k}+\\frac{C}{k^{2s}}.\n\\end{eqnarray*}\nWe plug this information into~\\eqref{udofjvw9etgierufgh} and we obtain~\\eqref{PAal203948},\nas desired.\n\nNow, we insert~\\eqref{PAal203948} into~\\eqref{PAal203948:PP} and we conclude that\n\\begin{eqnarray*}&& \\lim_{R\\to+\\infty} \\inf_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n\\\\ &&\\qquad\\le \\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_\\varphi(x,z)|^2 +2\\nabla E_v(x,z)\\cdot\\nabla\nE_\\varphi(x,z)\\big)\\,dx\\,dz\\\\\n&&\\qquad=\n\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2 -|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz.\n\\end{eqnarray*}\nHence, to prove~\\eqref{0OAPQO182}, it remains to show that\n\\begin{equation}\\label{0OAPQO182:BIS}\n\\begin{split}\n&\\lim_{R\\to+\\infty}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n\\\\\n\\le \\;& \\lim_{R\\to+\\infty} \\inf_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\n.\\end{split}\\end{equation}\nFor this, we fix~$\\Psi\\in\nC^\\infty_0({\\mathcal{B}}_R)$ with~$\\Psi(x,0)=\\varphi(x)$ and we use again~\\eqref{PAal203948}\nto see that\n\\begin{equation} \\label{00:10023a:023948}\n\\begin{split}\n& \\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Psi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\\\\\n=\\;&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi(x,z)|^2+2z^a\\nabla\\Psi(x,z)\\cdot\\nabla E_v(x,z)\\big)\\,dx\\,dz\n\\\\\n=\\;&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi(x,z)|^2+2z^a\\nabla E_\\varphi(x,z)\\cdot\\nabla E_v(x,z)\\big)\\,dx\\,dz\n\\\\&\\qquad+2\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Psi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\n\\\\ \\ge\\;&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi(x,z)|^2+2z^a\\nabla E_\\varphi(x,z)\\cdot\\nabla E_v(x,z)\\big)\\,dx\\,dz\n\\\\&\\qquad-2\\sup_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\left|\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla E_v(x,z)\\cdot\\nabla\\big(\\Phi(x,z)-E_\\varphi(x,z)\\big)\\,dx\\,dz\\right|\n\\\\ =\\;&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi(x,z)|^2+2z^a\\nabla E_\\varphi(x,z)\\cdot\\nabla E_v(x,z)\\big)\\,dx\\,dz\n-\\mu(R),\n\\end{split}\\end{equation}\nwith~$\\mu(R)$ independent of~$\\Psi$ and infinitesimal as~$R\\to+\\infty$.\n\nNow we take~$\\Psi_*\\in C^\\infty_0({\\mathcal{B}}_R)$ such that~$\\Psi_*(x,0)=\\varphi(x)$\nthat minimizes the functional~$\\Psi\\mapsto\\int_{ {\\mathcal{B}}_R^+ }z^a|\\nabla\\Psi(x,z)|^2\\,dx\\,dz$\nin such class. Then, using the variational equation for minimizers\ninside~${\\mathcal{B}}_R^+$ and the fact that~$\\Psi_*(x,0)=E_\\varphi(x,0)$, we see that\n\\begin{eqnarray*}\n&& \\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi(x,z)|^2-|\\nabla E_\\varphi(x,z)|^2\\big)\\,dx\\,dz\\\\\n&\\ge&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi_*(x,z)|^2-|\\nabla E_\\varphi(x,z)|^2\\big)\\,dx\\,dz\\\\\n&=&\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\nabla\\big(\\Psi_*-E_\\varphi\\big)(x,z)\\cdot\n\\nabla\\big(\\Psi_*+E_\\varphi\\big)(x,z)\n\\,dx\\,dz\\\\\n&=& \\int_{ {\\mathcal{B}}_R^+ }{\\rm div}\\,\\Big(\nz^a\\big(\\Psi_*-E_\\varphi\\big)(x,z)\n\\nabla\\big(\\Psi_*+E_\\varphi\\big)(x,z)\\Big)\n\\,dx\\,dz\\\\\n&=& \\int_{ L_R\\cup U_R}\nz^a\\big(\\Psi_*-E_\\varphi\\big)(x,z)\n\\partial_\\nu\\big(\\Psi_*+E_\\varphi\\big)(x,z)\\,d{\\mathcal{H}}^n(x,z)\\\\\n&=& -\\int_{ L_R\\cup U_R}\nz^a E_\\varphi(x,z)\n\\partial_\\nu E_\\varphi(x,z)\\,d{\\mathcal{H}}^n(x,z)\n.\n\\end{eqnarray*}\nTherefore, recalling~\\eqref{HENCE1} and~\\eqref{HENCE2}, we conclude that\n\\[\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla \\Psi(x,z)|^2-|\\nabla E_\\varphi(x,z)|^2\\big)\\,dx\\,dz\\ge -\\int_{ L_R\\cup U_R}\n\\frac{C}{R^{n+2s}}\n\\,d{\\mathcal{H}}^n(x,z)\\ge- \\frac{C}{R^{2s}}.\n\\]\n{F}rom this and~\\eqref{00:10023a:023948} we thus obtain that\n\\begin{eqnarray*}\n&& \\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Psi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\\\\\n&\\ge& \n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_\\varphi(x,z)|^2+2z^a\\nabla E_\\varphi(x,z)\\cdot\\nabla E_v(x,z)\\big)\\,dx\\,dz\n-\\mu(R)- \\frac{C}{R^{2s}}\\\\\n&=&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{\\varphi+v}(x,z)|^2\n-|\\nabla E_{v}(x,z)|^2\n\\big)\\,dx\\,dz-\\mu(R)- \\frac{C}{R^{2s}}.\n\\end{eqnarray*}\nSince this is valid for any~$\\Psi\\in C^\\infty_0({\\mathcal{B}}_R)$ with~$\\Psi(x,0)=\\varphi(x)$,\nwe conclude that\n\\begin{eqnarray*}\n&&\\inf_{ {\\Phi\\in\nC^\\infty_0({\\mathcal{B}}_R)}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla (E_{v}+\\Phi)(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz\\\\\n&\\ge&\n\\int_{ {\\mathcal{B}}_R^+ }\nz^a\\big(|\\nabla E_{v+\\varphi}(x,z)|^2\n-|\\nabla E_v(x,z)|^2\\big)\\,dx\\,dz-\\mu(R)- \\frac{C}{R^{2s}}\n.\\end{eqnarray*}\nBy taking the limit as~$R\\to+\\infty$,\nwe obtain~\\eqref{0OAPQO182:BIS}, as desired,\nand so we have\ncompleted the proof of Proposition~\\ref{0OAPQO182:P}.\n\\end{proof}\n\nIn view of Proposition~\\ref{0OAPQO182:P}, \nwe can now\nrelate the original and the extended energy functionals,\naccording to the following result:\n\n\\begin{corollary}\\label{CPOR:EQ}\nLet~$n\\ge2$ and~$s\\in(0,1)$.\nFor any~$v\\in W^{2,\\infty}(\\mathbb{R}^n)$ it holds that\n\\begin{equation}\\label{0o0oP}\n\\inf_{{R>0}\\atop{\\Phi\\in C^\\infty_0(\\mathcal{B}_R)}}\n{\\mathcal{E}}_{\\mathcal{B}_R}(E_v+\\Phi)-{\\mathcal{E}}_{\\mathcal{B}_R}(E_v)\\;\n=\\;\n\\inf_{{R>0}\\atop{\\varphi\\in C^\\infty_0(B_R)}}\n{\\mathcal{F}}_{B_R}(v+\\varphi)-\n{\\mathcal{F}}_{B_R}(v).\n\\end{equation}\n\\end{corollary}\n\n\\begin{proof}\n Notice that, given~$\\varphi\\in C^\\infty_0(\\mathbb{R}^n)$,\n\\begin{equation}\\label{IDEPHI}\n\\begin{split}\n&\\inf_{ {{R>0}\\atop{\\Phi\\in C^\\infty_0(\\mathcal{B}_R)}}\\atop{\\Phi(x,0)=\\varphi(x)}}\n{\\mathcal{E}}_{\\mathcal{B}_R}(E_v+\\Phi)-{\\mathcal{E}}_{\\mathcal{B}_R}(E_v)\\;\n\\\\=\\;& \n\\inf_{ {{R>0}\\atop{\\Phi\\in C^\\infty_0(\\mathcal{B}_R)}}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{\\mathbb{R}^{n+1}_+} z^a\\Big( |\\nabla (E_v+\\Phi)(x,z)|^2-|\\nabla E_v(x,z)|^2\n\\Big)\\,dx\\,dz\\\\ &\\qquad\\qquad\\qquad+\n\\int_{\\mathbb{R}^n}\\Big( F(v(x)+\\varphi(x))-F(v(x))\\Big)\\,dx\n\\\\ =\\;&\\lim_{R\\to+\\infty}\n\\inf_{ {{\\Phi\\in C^\\infty_0(\\mathcal{B}_R)}}\\atop{\\Phi(x,0)=\\varphi(x)}}\n\\int_{\\mathbb{R}^{n+1}_+} z^a\\Big( |\\nabla (E_v+\\Phi)(x,z)|^2-|\\nabla E_v(x,z)|^2\\Big)\n\\,dx\\,dz\\\\&\\qquad\\qquad\\qquad+\n\\int_{\\mathbb{R}^n}\\Big( F(v(x)+\\varphi(x))-F(v(x))\\Big)\\,dx\n\\\\=\\;&\n\\iint_{\\mathbb{R}^{2n}}\n\\frac{|(v+\\varphi)(x)-(v+\\varphi)(y)|^2-\n|v(x)-v(y)|^2\n}{|x-y|^{n+2s}}\\,dx\\,dy\\\\&\\qquad\\qquad\\qquad\n+\n\\int_{\\mathbb{R}^n}\\Big( F(v(x)+\\varphi(x))-F(v(x))\\Big)\\,dx\\\\\n=\\;&\n{\\mathcal{F}}_{B_R}(v+\\varphi)-\n{\\mathcal{F}}_{B_R}(v),\n\\end{split}\\end{equation}\nthanks to Proposition~\\ref{0OAPQO182:P}.\nSince the identity in~\\eqref{IDEPHI}\nis valid for any~$\\varphi\\in C^\\infty_0(\\mathbb{R}^n)$\n(i.e. for any~$R>0$ and any~$\\varphi\\in C^\\infty_0(B_R)$), taking the infimum in such\nclass we obtain~\\eqref{0o0oP}, as desired.\n\\end{proof}\n\nDue to Corollary~\\ref{CPOR:EQ},\nthe following equivalence result for minimizers holds true:\n\n\\begin{proposition}\\label{EQUI DL}\n$E_v$ is an extended local minimizer according to Definition~\\ref{D:L:2}\nif and only if~$v$ is a local minimizer according to Definition~\\ref{D:L}.\n\\end{proposition}\n\n\\begin{proof} We observe that~$E_v$ is an extended local minimizer according to Definition~\\ref{D:L:2}\nif and only if the first term in~\\eqref{0o0oP} is nonnegative;\non the other hand, $v$ is a local minimizer according to Definition~\\ref{D:L}\nif and only if the last term in~\\eqref{0o0oP} is nonnegative; since\nthe two terms in~\\eqref{0o0oP} are equal, the desired result is established.\n\\end{proof}\n\nIn view of Proposition~\\ref{EQUI DL} (see also Lemma 6.1 in~\\cite{cabre-TAMS}),\nit is natural to say that~$u$ is a stable solution of~\\eqref{ALLEN-GEN}\nif the second derivative of the associated energy functional is nonnegative,\naccording to the following setting:\n\n\\begin{definition}\nLet~$u$ be a solution of~\\eqref{ALLEN-GEN} in~$\\mathbb{R}^n$. We say that~$u$ is stable if\n$$ \\int_{\\mathbb{R}^{n+1}_+} z^a |\\nabla \\zeta(x,z)|^2\\,dx\\,dz+\\int_{\\mathbb{R}^n} F''(u(x))\\,\\zeta^2(x,0)\\,dx\\ge0$$\nfor any~$\\zeta\\in C^\\infty_0(\\mathbb{R}^{n+1})$.\n\\end{definition}\n\n\\section{Variational classification of $1$D solutions}\\label{SEC:2}\n\nThe goal of this section is to establish the following result:\n\n\\begin{lemma}\\label{CL:MO}\nLet~$s\\in(0,1)$ and\n$v\\in C^2(\\mathbb{R},[-1,1])$ be a stable solution of~$(-\\Delta)^s v=f(v)$ in~$\\mathbb{R}$.\nAssume also that~$\\dot v\\ge0$. Then~$v$ is a local minimizer.\n\\end{lemma}\n\n\\begin{proof} The monotonicity of~$v$ implies that the following limits exist:\n$$ \\underline\\ell:= \\lim_{t\\to-\\infty}v(t)\\le\\lim_{t\\to+\\infty}v(t) =:\\overline\\ell.$$\nWe also consider the sequence of functions~$v_k(t):=v(t+k)$.\nBy the Theorem of Ascoli, up to a subsequence we know that~$v_k$ converges to~$\\overline\\ell$\nin~$C^2_{\\rm loc}(\\mathbb{R})$, and so, passing the equation to the limit, we conclude that~$f(\\overline\\ell)=0$.\nSimilarly, one sees that~$f(\\underline\\ell)=0$.\nAs a consequence,\n\\begin{equation}\\label{889:p}\n\\underline\\ell,\\,\\overline\\ell\\in\\{-1,0,1\\}.\\end{equation}\nNow, we claim that\n\\begin{equation}\\label{NOT Z}\n{\\mbox{$v$ is not identically zero.}}\n\\end{equation}\nThe proof is by contradiction: if\n$v$ is identically zero, we take~$\\psi\\in C^\\infty_0(\\mathbb{R}^2)$\nand, for~$\\varepsilon>0$, we let~$\\psi_\\varepsilon(x,z):=\\psi(\\varepsilon x,\\varepsilon z)$. \nThe stability inequality for~$\\psi_\\varepsilon$ gives that\n\\begin{eqnarray*}\n0 &\\le& \\int_{\\mathbb{R}^2_+} z^a |\\nabla\\psi_\\varepsilon(x,z)|^2\\,dx\\,dz+\n\\int_{\\mathbb{R}} F''(v(x))\\,\\psi_\\varepsilon^2(x,0)\\,dx\\\\ &=&\n\\varepsilon^2 \\int_{\\mathbb{R}^2_+} z^a |\\nabla\\psi(\\varepsilon x,\\varepsilon z)|^2\\,dx\\,dz+\n\\int_{\\mathbb{R}} F''(v(x))\\,\\psi^2(\\varepsilon x,0)\\,dx\n\\\\ &=&\n\\varepsilon^{-a} \\int_{\\mathbb{R}^2_+} Z^a |\\nabla\\psi(X,Z)|^2\\,dX\\,dZ-\\varepsilon^{-1}\n\\int_{\\mathbb{R}} f'(0)\\,\\psi^2(X,0)\\,dX\n\\\\ &=& C_1\\varepsilon^{2s-1}-C_2\\varepsilon^{-1} f'(0),\n\\end{eqnarray*}\nfor some~$C_1$, $C_2>0$. {F}rom this, one obtains that\n$$ f'(0)\\le \\lim_{\\varepsilon\\searrow0} \\frac{C_1\\varepsilon^{2s}}{C_2}=0.$$\nThis is a contradiction, since~$f'(0)>0$ and thus~\\eqref{NOT Z} is proved.\n\nTo complete the proof of Lemma~\\ref{CL:MO},\nwe now distinguish two cases, either~$v$ is constant or not.\nIf~$v$ is constant, then it is either identically~$-1$ or identically~$1$,\ndue to~\\eqref{NOT Z}, and this implies the desired result.\n\nSo, we can now focus on the case in which $v$ is not constant.\nThen, $\\underline\\ell<\\overline\\ell$. So, from~\\eqref{889:p},\nwe have that~$v$ is a transition layer connecting:\n\\begin{enumerate}\n\\item either $-1$ to~$0$,\n\\item or $0$ to $1$,\n\\item or $-1$ to $1$.\n\\end{enumerate}\nIn view of Theorem 2.2(i) in~\\cite{cabre-sire-AIHP}, the first two cases cannot occur\nand therefore\n\\begin{equation}\\label{78s:10234}\n\\lim_{t\\to\\pm\\infty} v(t)=\\pm1.\n\\end{equation}\nSince the proof of this fact relies on the theory of layer solutions,\nwe provide the details of the argument that we used.\nWe argue for a contradiction and we suppose that\n\\begin{eqnarray}\n&& \\label{C:caso1}\n{\\mbox{ either }} \\lim_{t\\to-\\infty} v(t)=-1 {\\mbox{ and }} \\lim_{t\\to+\\infty} v(t)=0\\\\&&\\label{C:caso2}\n{\\mbox{ or }} \\lim_{t\\to-\\infty} v(t)=0 {\\mbox{ and }} \\lim_{t\\to+\\infty} v(t)=1.\n\\end{eqnarray}\nBy maximum principle, we know that~$\\dot v>0$.\nThen, if we set either~$\\tilde v(t):= 2v(t)+1$ (if the case in~\\eqref{C:caso1} holds true)\nor~$\\tilde v(t):= 2v(t)-1$ (if~\\eqref{C:caso2} holds true), we have that\nthe derivative of~$\\tilde v$ is strictly positive and\n\\begin{equation}\\label{78s:10234:BIS} \\lim_{t\\to\\pm\\infty}\\tilde v(t)=\\pm1.\\end{equation}\nIn addition,\n$$ (-\\Delta)^s \\tilde v(t)= 2(-\\Delta)^s v(t)= 2f(v(t))=2 \nf\\left( \\frac{\\tilde v(t)\\mp1}{2}\\right)=\n-G'(\\tilde v(t)),$$\nwith\n$$ G(r):= -2\\int_{0}^r f\\left( \\frac{\\tau\\mp1}{2}\\right)\\,d\\tau\n=\n-4\\int_{\\mp1\/2}^{(r\\mp1)\/2} f(\\sigma)\\,d\\sigma.$$\nThis and~\\eqref{78s:10234:BIS} give that we are in the setting of\nTheorem 2.2(i) in~\\cite{cabre-sire-AIHP}. In particular, from formula~(2.8)\nin~\\cite{cabre-sire-AIHP} we know that\n$$ 0=G(-1)-G(1)=\n4\\int_{\\mp1\/2}^{(1\\mp1)\/2} f(\\sigma)\\,d\\sigma\n-4\\int_{\\mp1\/2}^{(-1\\mp1)\/2} f(\\sigma)\\,d\\sigma=\n4\\int_{(-1\\mp1)\/2}^{(1\\mp1)\/2} f(\\sigma)\\,d\\sigma\n,$$\nhence\n$$ {\\mbox{ either }}\\quad\n\\int_{-1}^{0} f(\\sigma)\\,d\\sigma=0 \\quad{\\mbox{ or }}\\quad\n\\int_{0}^{1} f(\\sigma)\\,d\\sigma=0.$$\nThis is in contradiction with~\\eqref{BISTABLE}\nand so it proves~\\eqref{78s:10234}.\n\nHence, necessarily~$v$ is a transition layer connecting~$-1$ to~$1$\nand so it is minimal due to the sliding method (see e.g. the proof of Lemma~9.1\nin~\\cite{VSS}).\n\\end{proof}\n\n\\section{Classification of the profiles at infinity}\\label{CLASS:A}\n\nIn this section, we consider the two profiles \nof a given solution at infinity. Namely, if~$s\\in(0,1)$ and~$u\\in C^2(\\mathbb{R}^n,[-1,1])$\nis a solution of~$(-\\Delta)^s u=f(u)$ in~$\\mathbb{R}^n$, \nwith~$\\partial_{x_n}u>0$ in~$\\mathbb{R}^n$,\nwe set\n$$ \\underline u(x'):=\\lim_{x_n\\to-\\infty} u(x',x_n)\\quad{\\mbox{ and }}\\quad\n\\overline u(x'):=\\lim_{x_n\\to+\\infty} u(x',x_n).$$\nIn this setting, we have:\n\n\\begin{lemma} \\label{02738}\nAssume that~$n=3$. Then,\nboth~$\\underline u$ and~$\\overline u$ are $1$D and local minimizers.\n\\end{lemma}\n\n\\begin{proof} By passing the equation to the limit, we have that\n\\begin{equation}\\label{sta000}\n{\\mbox{both~$\\underline u$ and~$\\overline u$ are \nstable solutions in~$\\mathbb{R}^{n-1}=\\mathbb{R}^2$.}}\\end{equation}\nThe proof of~\\eqref{sta000} is based on a general\nargument (see e.g.~\\cite{MR2483642}), given in details\nhere for the sake of completeness.\nLet~$\\xi\\in C^\\infty_0(\\mathbb{R}^{n})$ and~$\\eta\\in C^\\infty_0((-1,1))$, with~$\\eta(0)=1$.\nGiven~$\\varepsilon>0$, we set~$\\xi_\\varepsilon(x,z)=\\xi_\\varepsilon(x',x_n,z):=\n\\xi(x',z)\\eta(\\varepsilon x_n)$. Notice that~$\\xi_\\varepsilon\\in C^\\infty_0(\\mathbb{R}^{n+1})$,\ntherefore by the stability of~$u$ and the translation invariance we\nhave that\n\\begin{eqnarray*}\n0 &\\le& \\lim_{t\\to+\\infty}\nC \\int_{\\mathbb{R}^{n+1}_+} z^a |\\nabla \\xi_\\varepsilon(x,z)|^2\\,dx\\,dz+\n\\int_{\\mathbb{R}^n} F''( u(x',x_n+t))\\,\\xi_\\varepsilon(x,0)\\,dx\\\\&=&\nC \\int_{\\mathbb{R}^{n+1}_+} z^a |\\nabla \\xi_\\varepsilon(x,z)|^2\\,dx\\,dz+\n\\int_{\\mathbb{R}^n} F''( \\overline u(x'))\\,\\xi_\\varepsilon^2(x,0)\\,dx\\\\\n&=&\nC \\int_{\\mathbb{R}^{n+1}_+} z^a \\Big(\n|\\nabla \\xi(x',z)|^2 \\eta^2(\\varepsilon x_n)+\n\\varepsilon^2 |\\nabla \\eta(\\varepsilon x_n)|^2 \\xi^2(x',z)+2\\varepsilon\n\\eta(\\varepsilon x_n)\\xi(x',z)\\nabla\\eta(\\varepsilon x_n)\\cdot\\nabla \\xi(x',z)\n\\Big)\\,dx\\,dz\\\\&&\\quad+\n\\int_{\\mathbb{R}^n} F''( \\overline u(x'))\\,\\xi^2(x',0)\\eta^2(\\varepsilon x_n)\\,dx.\n\\end{eqnarray*}\nHence, taking the limit as~$\\varepsilon\\to0$,\n$$ 0\\le\nC \\int_{\\mathbb{R}^{n+1}_+} z^a \n|\\nabla \\xi(x',z)|^2 \\,dx\\,dz+\n\\int_{\\mathbb{R}^n} F''( \\overline u(x'))\\,\\xi^2(x',0)\\,dx,$$\nand so~$\\overline u$ is stable in~$\\mathbb{R}^{n-1}$.\nThis proves~\\eqref{sta000} for $\\overline u$ (the case of~$\\underline u$\nis similar).\n\nAs a consequence of \\eqref{sta000} and of the classification results in the plane\n(see in particular Theorem 2.12 in~\\cite{cabre-TAMS}, or~\\cite{SV09}), we conclude that~$\n\\underline u$ and~$\\overline u$ are~$1$D and monotone.\nThen, the local minimality is a consequence of Lemma~\\ref{CL:MO}.\n\\end{proof}\n\n\\section{Local minimization by range constraint}\\label{CLASS:B}\n\nIn this section, we point out that perturbations which do not pointwise exceed\nthe limit profiles necessarily increase the energy. For the classical case,\nthis property has been exploited in Theorem~4.5 of~\\cite{AAC},\nTheorem~10.4 of~\\cite{danielli} and \nLemma~2.2 of~\\cite{FV11}. In the framework of this paper, the result that we need is the following:\n\n\\begin{lemma}\\label{INTRAP}\nLet~$s\\in(0,1)$ and~$u\\in C^2(\\mathbb{R}^n,[-1,1])$\nbe a solution of~$(-\\Delta)^s u=f(u)$ in~$\\mathbb{R}^n$, \nwith~$\\partial_{x_n}u>0$ in~$\\mathbb{R}^n$.\nLet\n$$ \\underline u(x'):=\\lim_{x_n\\to-\\infty} u(x',x_n)\\quad{\\mbox{ and }}\\quad\n\\overline u(x'):=\\lim_{x_n\\to+\\infty} u(x',x_n).$$\nLet also~$R>0$ and~$\\Phi\\in C^\\infty_0({\\mathcal{B}}_R)$ and suppose that\n$$ (E_u+\\Phi)(X)\\in \\big[ E_{\\underline u}(X),\\,E_{\\overline u}(X)\\big]\\qquad{\\mbox{for \nany }}X\\in\\mathbb{R}^{n+1}_+.$$\nThen~${\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_u)\\le{\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_u+\\Phi)$.\n\\end{lemma}\n\n\\begin{proof} The argument is by contradiction. \nWe suppose that there exist~$R>0$\nand a perturbation~$\\Phi\\in C^\\infty_0({\\mathcal{B}}_R)$ with\n$$ (E_u+\\Phi)(X)\\in \\big[ E_{\\underline u}(X),\\,E_{\\overline u}(X)\\big]\\qquad{\\mbox{for \nany }}X\\in\\mathbb{R}^{n+1}_+$$\nand such that\n$$ {\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_u+\\Phi) < {\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_u).$$\nThat is, letting~$\\varphi(x):=\\Phi(x,0)$, \nby Proposition~\\ref{0OAPQO182:P},\n\\begin{eqnarray*}\n0 &>& \n\\frac{1}{2}\\int_{\\mathbb{R}^{n+1}_+}z^a \\big(|\\nabla (E_u+\\Phi)(x,z)|^2-\n|\\nabla E_u(x,z)|^2\\big)\\,dx\\,dz\n+\\int_{B_R} \\big( F((u+\\varphi)(x))-F(u(x))\\big)\\,dx\n\\\\ &=&\n\\frac{1}{2}\\iint_{\\mathbb{R}^{2n}}\\frac{|(u+\\varphi)(x)-(u+\\varphi)(y)|^2\n-|u(x)-u(y)|^2}{|x-y|^{n+2s}}\\,dx\\,dy\n+\\int_{B_R} \\big( F((u+\\varphi)(x))-F(u(x))\\big)\\,dx\\\\&=&\n{\\mathcal{F}}_{B_R}(u+\\varphi)-{\\mathcal{F}}_{B_R}(u).\n\\end{eqnarray*}\nTherefore, there exists a perturbation~$w_o:=u+\\varphi_o$ of~$u$, with~$\\varphi_o\\in C^\\infty_0(B_R)$,\nsuch that\n\\begin{equation}\\label{TRA:1}\nw_o(x)\\in \\big[ {\\underline u}(x'),\\,{\\overline u}(x')\\big]\\qquad{\\mbox{for \nany }}x\\in\\mathbb{R}^n\\end{equation}\nand\n\\begin{equation}\\label{GHL:ocon} {\\mathcal{F}}_{B_R}(w_o)=\\inf_{{\\varphi\\in C^\\infty_0(B_R)}\\atop{\n(u+\\varphi)(x)\n\\in [ {\\underline u}(x'),\\,{\\overline u}(x')]}}\n{\\mathcal{F}}_{B_R}(u+\\varphi)<{\\mathcal{F}}_{B_R}(u).\\end{equation}\nAs a consequence, by taking energy perturbations,\nwe see that~$(-\\Delta)^s w_o=f'(w_o)$ inside~$\\{\nx\\in\\mathbb{R}^n {\\mbox{ s.t. }} w_o(x)\\in [ {\\underline u}(x'),\\,{\\overline u}(x')]\\}$.\nIn addition, if~$w(x_o)={\\overline u}(x_o)$, then~$(-\\Delta)^s w_o(x_o)\\le f'(w_o(x_o))$.\nSimilarly,\nif~$w(x_o)={\\underline u}(x_o)$, then~$(-\\Delta)^s w_o(x_o)\\ge f'(w_o(x_o))$.\n\nNow we claim that strict inequalities hold in~\\eqref{TRA:1}, namely\n\\begin{equation}\\label{TRA:2}\nw_o(x)\\in \\big( {\\underline u}(x'),\\,{\\overline u}(x')\\big)\\qquad{\\mbox{for \nany }}x\\in\\mathbb{R}^n.\\end{equation}\nTo check this, suppose by contradiction, for instance,\nthat there exists~$x_o\\in\\mathbb{R}^n$\nsuch that~$w_o(x_o)={\\overline u}(x_o')$. Then, the function~$\\zeta(x):=\n{\\overline u}(x')-w_o(x)$ has a minimum at~$x_o$. Accordingly,\n$$ 0\\ge(-\\Delta)^s \\zeta(x_o) \\ge f({\\overline u}(x_o'))- f'(w_o(x_o)) =0.$$\nThis gives that~$\\zeta$ must vanish identically, and thus that~$w_o$ is identically equal to~$\n{\\overline u}$. Then, taking~$\\bar x\\in\\mathbb{R}^n\\setminus B_R$, we have that\n$$ {\\overline u}(\\bar x')= w_o(\\bar x)=u(\\bar x)<\n{\\overline u}(\\bar x').$$\nThis is a contradiction, and therefore~\\eqref{TRA:2} is established.\n\n{F}rom~\\eqref{TRA:2}, it follows that\n$$ \\max_{\\overline{B_R}}( w_o-{\\overline u})<0.$$\nNow, we let~$w_k(x):=u(x+ke_n)$. We claim that there exists~$\\bar k\\in\\mathbb{N}$ such that\n\\begin{equation} \\label{df:29}\nw_{\\bar k}(x)>w_o(x) {\\mbox{ for any }} x\\in\\mathbb{R}^n.\\end{equation}\nTo prove this, we argue by contradiction and suppose that for any~$k\\in\\mathbb{N}$\nthere exists~$x_k\\in\\mathbb{R}^n$ such that~$\nw_{k}(x_k)\\le w_o(x_k)$.\nThen, $x_k\\in B_R$\n(otherwise~$w_o(x_k)=u(x_k)w_o\n\\}$. We claim that\n\\begin{equation}\\label{k0}\nk_\\star=0.\n\\end{equation}\nOnce again, suppose not. Then, the function~$v:=w^{k_\\star}-w_o$\nwould satisfy~$v\\ge0$ in~$\\mathbb{R}^n$, with~$v(\\tilde x)=0$, for some~$\\tilde x\\in\\overline{B_R}$.\nAs a consequence,\n$$ 0\\ge (-\\Delta)^s v(\\tilde x) = f(w^{k_\\star}(\\tilde x))-f(w(\\tilde x))=0,$$\nwhich implies that~$v$ vanishes identically. In particular, fixing~$x_\\star$\noutside~$B_R$, we would have that\n$$ 0=v(x_\\star)=\nw^{k_\\star}(x_\\star)-w_o(x_\\star)=\nu(x_\\star+k_\\star e_n)-u(x_\\star)>0.$$\nThis contradiction completes the proof of~\\eqref{k0}.\n\nNow, in view of~\\eqref{k0}, we obtain that, for any~$x\\in\\mathbb{R}^n$,\n$$ w_o(x)\\le\\lim_{k\\searrow0} w_k(x)=\n\\lim_{k\\searrow0} u(x+ke_n)=u(x).$$\nSimilarly, one can prove that~$w_o(x)\\ge u(x)$.\nTherefore, $w_o$ and~$u$ must coincide\nand so~${\\mathcal{F}}_{B_R}(w_o)={\\mathcal{F}}_{B_R}(u)$.\nBut this fact is in contradiction with~\\eqref{GHL:ocon}\nand so we have completed the proof of Lemma~\\ref{INTRAP}.\n\\end{proof}\n\n\\section{Local minimization properties inherited from those of the profiles at infinity}\\label{CLASS:C}\n\nIn this section, we show that if the profiles at infinity are local minimizers,\nthen so is the original solution. In the classical case of the Laplacian, this property\nwas discussed, for instance, in Proposition~2.3 of~\\cite{FV11}.\nIn our setting, the result that\nwe need is the following (and it uses\nthe pivotal definition of extended local minimizer in Definition~\\ref{D:L:2}):\n\n\\begin{lemma}\\label{0tyg27427412}\nLet~$s\\in(0,1)$ and~$u\\in C^2(\\mathbb{R}^n,[-1,1])$\nbe a solution of~$(-\\Delta)^s u=f(u)$ in~$\\mathbb{R}^n$, \nwith~$\\partial_{x_n}u>0$ in~$\\mathbb{R}^n$.\nLet\n$$ \\underline u(x'):=\\lim_{x_n\\to-\\infty} u(x',x_n)\\quad{\\mbox{ and }}\\quad\n\\overline u(x'):=\\lim_{x_n\\to+\\infty} u(x',x_n)$$\nand suppose that~$\\underline u$ and~$\\overline u$ are local minimizers (in~$\\mathbb{R}^{n-1}$).\nThen, $E_u$ is an extended local minimizer (in~$\\mathbb{R}^{n+1}_+$)\nand~$u$ is a local minimizer (in~$\\mathbb{R}^{n}$).\n\\end{lemma}\n\n\\begin{proof} Our goal is to show that $E_u$ is an extended local minimizer \nin the sense of Definition~\\ref{D:L:2} (from this, we also obtain that~$u$ is a local minimizer,\nthanks to Proposition~\\ref{EQUI DL}). To this aim, fixed~$x_n\\in\\mathbb{R}$,\nwe use the following ``slicing notation'' for a domain~$\\Omega\\subset\\mathbb{R}^{n+1}$:\nwe let\n$$ \\Omega^{x_n} := \\{ (x',z)\\in\\mathbb{R}^{n-1}\\times\\mathbb{R} {\\mbox{ s.t. }}\n(x',x_n,z)\\in\\Omega\\}.$$\nAlso, the function~$\\overline u:\\mathbb{R}^{n-1}\\to\\mathbb{R}$ can be seen as a function on~$\\mathbb{R}^n$,\nby defining~$\\overline u_\\star(x',x_n):=\\overline u(x')$ and so we can consider its $a$-harmonic\nextension~$E_{\\overline u_\\star}$.\nGiven~$R>0$ and~$\\Phi\\in C^\\infty_0({\\mathcal{B}}_R)$, with~$\\varphi(x):=\\Phi(x,0)$,\nwe also define~$\\Phi_{x_n}(x',z):=\\Phi(x',x_n,z)$ and~$\\varphi_{x_n}(x'):=\\varphi(x',x_n)$.\nHence\nwe have that\n\\begin{equation}\\label{8920tr}\n\\begin{split}\n{\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_{\\overline u_\\star}+\\Phi)\\;\n&= \n\\frac{1}{2}\n\\int_{{\\mathcal{B}}_R^+}z^a |\\nabla (E_{\\overline u_\\star}+\\Phi)(x,z)|^2\\,dx\\,dz+\n\\int_{B_R} F(\\overline u(x')+\\varphi(x))\\,dx\n\\\\ &\\ge \\int_\\mathbb{R}\\left[\n\\frac{1}{2}\n\\int_{(B_R)_{x_n}\\times(0,R)}z^a |\\nabla_{(x',z)} E_{\\overline u}(x',z)+\\nabla_{(x',z)}\n\\Phi_{x_n} (x',z)|^2\\,dx'\\,dz\\right.\\\\ &\\qquad+\\left.\n\\int_{(B_R)_{x_n}} F(\\overline u(x')+\\varphi_{x_n}(x))\\,dx'\n\\right]\\,dx_n\n\\end{split}\\end{equation}\nAlso, since~$\\overline u$ is a local minimizer in~$\\mathbb{R}^{n-1}$, we have\nthat~$E_{\\overline u}$ is an extended local minimizer in~$\\mathbb{R}^{n-1}\\times(0,+\\infty)$,\nthanks to\nProposition~\\ref{EQUI DL}, and therefore\n\\begin{eqnarray*}&&\n\\frac{1}{2}\n\\int_{(B_R)_{x_n}\\times(0,R)}z^a |\\nabla_{(x',z)} E_{\\overline u}(x',z)+\\nabla_{(x',z)}\n\\Phi_{x_n} (x',z)|^2\\,dx'\\,dz+\n\\int_{(B_R)_{x_n}} F(\\overline u(x')+\\varphi_{x_n}(x))\\,dx'\n\\\\ &\\ge&\n\\frac{1}{2}\n\\int_{(B_R)_{x_n}\\times(0,R)}z^a |\\nabla_{(x',z)} E_{\\overline u}(x',z)|^2\\,dx'\\,dz\n+\\int_{(B_R)_{x_n}} F(\\overline u(x'))\\,dx'.\n\\end{eqnarray*}\nBy inserting this inequality into~\\eqref{8920tr}, we obtain that~$\n{\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_{\\overline u_\\star}+\\Phi)\\ge\n{\\mathcal{E}}_{ {\\mathcal{B}}_R }(E_{\\overline u_\\star})$.\nThat is, \n\\begin{equation}\\label{OVER}\n{\\mbox{$E_{\\overline u_\\star}$ is an extended local minimizer in~$\\mathbb{R}^{n+1}_+$.\n}}\\end{equation}\nSimilarly,\none can define~$\\underline u_\\star(x',x_n):=\\underline u(x')$\nand conclude that\n\\begin{equation}\\label{UNDER}\n{\\mbox{$E_{\\underline u_\\star}$ is an extended local minimizer in~$\\mathbb{R}^{n+1}_+$.\n}}\\end{equation}\nNow,\ngiven~$R>0$ and~$\\Psi\\in C^\\infty_0({\\mathcal{B}}_R)$, with~$\\psi(x):=\\Psi(x,0)$,\nwe consider the perturbation~$E_u+\\Psi$.\nWe define\n\\begin{eqnarray*}\n&&\\alpha(X):=\\left\\{\n\\begin{matrix}\nE_{\\overline u_\\star}(X) & {\\mbox{ if }} (E_u+\\Psi)(X)\\le E_{\\overline u_\\star}(X),\\\\\n(E_u+\\Psi)(X) & {\\mbox{ if }} (E_u+\\Psi)(X)> E_{\\overline u_\\star}(X),\n\\end{matrix}\n\\right. \\\\\n&&\\beta(X):=\\left\\{\n\\begin{matrix}\nE_{\\overline u_\\star}(X) & {\\mbox{ if }} (E_u+\\Psi)(X)> E_{\\overline u_\\star}(X),\\\\\nE_{\\underline u_\\star}(X) & {\\mbox{ if }} (E_u+\\Psi)(X)< E_{\\underline u_\\star}(X),\\\\\n(E_u+\\Psi)(X) & {\\mbox{ if }} (E_u+\\Psi)(X)\\in\\big[ E_{\\underline u_\\star}(X),\n\\,E_{\\overline u_\\star}(X)\\big],\n\\end{matrix}\n\\right. \\\\\n{\\mbox{and }}&&\\gamma(X):=\\left\\{\n\\begin{matrix}\nE_{\\underline u_\\star}(X) & {\\mbox{ if }} (E_u+\\Psi)(X)\\ge E_{\\underline u_\\star}(X),\\\\\n(E_u+\\Psi)(X) & {\\mbox{ if }} (E_u+\\Psi)(X)< E_{\\underline u_\\star}(X).\n\\end{matrix}\n\\right. \n\\end{eqnarray*}\nBy~\\eqref{OVER}, we have that\n\\begin{equation}\\label{P:X1}\n{\\mathcal{E}}_{ \\{ E_u+\\Psi>E_{\\overline u_\\star}\\} }(E_{\\overline u_\\star}) \\le \n{\\mathcal{E}}_{ \\{ E_u+\\Psi>E_{\\overline u_\\star}\\} }(\\alpha).\n\\end{equation}\nSimilarly,\nby~\\eqref{UNDER}, we have that\n\\begin{equation}\\label{P:X2}\n{\\mathcal{E}}_{ \\{ E_u+\\PsiE_{\\overline u_\\star}\\}$ and~$\n\\{ E_u+\\PsiE_{\\overline u_\\star}\\} }(E_u+\\Psi)+\n{\\mathcal{E}}_{\\{ E_u+\\PsiE_{\\overline u_\\star}\\} }(\\alpha)+\n{\\mathcal{E}}_{\\{ E_u+\\PsiE_{\\overline u_\\star}\\} }(E_{\\overline u_\\star})+\n{\\mathcal{E}}_{\\{ E_u+\\PsiE_{\\overline u_\\star}\\} }(\\beta)+\n{\\mathcal{E}}_{\\{ E_u+\\Psi0$ in~$\\mathbb{R}^n$.\n\nLet \n$$ \\underline u(x'):=\\lim_{x_n\\to-\\infty} u(x',x_n)\\quad{\\mbox{ and }}\\quad\n\\overline u(x'):=\\lim_{x_n\\to+\\infty} u(x',x_n).$$\nAssume that (possibly after a rotation) \n$\\underline u$ and~$\\overline u$ depend on at most two Euclidean variables\n(not necessarily the same).\n\nThen $u$ is a local minimizer.\n\nMoreover, if $n\\le8$, there exists~$s(n)\\in \\left[0,\\frac12\\right)$ such that if~$\ns\\in\\left(s(n),\\frac12\\right)$ then~$u$ is $1$D.\n\\end{theorem}\n\nThe proof of Theorem~\\ref{MAIN:VAR} shares\nthe point of view taken in~\\cite{MR2483642} and~\\cite{FV11},\nand follows the same lines\nas that of Theorem~\\ref{MAIN}, with the modifications listed here below:\n\\begin{itemize}\n\\item By\nfollowing verbatim the proof of Lemma~\\ref{02738}, and\nexploiting that~$\\underline u$ and~$\\overline u$ are stable two-dimensional solutions, one obtains that \nthey are local minimizers and~$1$D;\n\\item {F}rom this and Lemma~\\ref{0tyg27427412}, one deduces\nthe first claim in Theorem~\\ref{MAIN:VAR};\n\\item The second claim in Theorem~\\ref{MAIN:VAR} follows\nfrom the first claim and the argument in Section~\\ref{CLASS:D}\n(in this framework, for $s$ large enough,\none can exploit\nTheorem~1.6 of~\\cite{DSVarxiv}\nin place of Theorem~1.4 of~\\cite{DSVarxiv}).\n\\end{itemize}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{sec:int}\n\\setcounter{equation}{0}\n\nBy an original observation of Jucys~\\cite{j:yo},\nall primitive idempotents\nof the symmetric group $\\Sym_n$ can be obtained by taking certain\nlimit values of the rational function\n\\beql{phiu}\n\\Phi(u_1,\\dots,u_n)= \\prod_{1\\leqslant i1,\\\\\ns_is_{i+1}s_i&=s_{i+1}s_is_{i+1},\\qquad\n\\ee_i\\ee_{i+1}\\ee_i=\\ee_i,\\qquad \\ee_{i+1}\\ee_i\\ee_{i+1}=\\ee_{i+1},\\\\\n\\su_i\\ee_{i+1}\\ee_i&=\\su_{i+1}\\ee_i,\\qquad \\ee_{i+1}\\ee_i\n\\su_{i+1}=\\ee_{i+1}\\su_i,\\qquad\ni=1,\\dots,n-2.\n\\end{aligned}\n\\non\n\\end{equation}\n\\epr\n\nThe generators $s_i$ and $\\ee_i$ correspond to the following diagrams\nrespectively:\n\n\\begin{center}\n\\begin{picture}(400,60)\n\\thinlines\n\n\\put(10,20){\\circle*{3}}\n\\put(30,20){\\circle*{3}}\n\\put(70,20){\\circle*{3}}\n\\put(90,20){\\circle*{3}}\n\\put(130,20){\\circle*{3}}\n\\put(150,20){\\circle*{3}}\n\n\\put(10,40){\\circle*{3}}\n\\put(30,40){\\circle*{3}}\n\\put(70,40){\\circle*{3}}\n\\put(90,40){\\circle*{3}}\n\\put(130,40){\\circle*{3}}\n\\put(150,40){\\circle*{3}}\n\n\\put(10,20){\\line(0,1){20}}\n\\put(30,20){\\line(0,1){20}}\n\\put(70,20){\\line(1,1){20}}\n\\put(90,20){\\line(-1,1){20}}\n\\put(130,20){\\line(0,1){20}}\n\\put(150,20){\\line(0,1){20}}\n\n\\put(45,25){$\\cdots$}\n\\put(105,25){$\\cdots$}\n\n\\put(8,5){\\scriptsize $1$ }\n\\put(28,5){\\scriptsize $2$ }\n\\put(68,5){\\scriptsize $i$ }\n\\put(86,5){\\scriptsize $i+1$ }\n\\put(122,5){\\scriptsize $n-1$ }\n\\put(150,5){\\scriptsize $n$ }\n\n\\put(190,25){\\text{and}}\n\n\\put(250,20){\\circle*{3}}\n\\put(270,20){\\circle*{3}}\n\\put(310,20){\\circle*{3}}\n\\put(330,20){\\circle*{3}}\n\\put(370,20){\\circle*{3}}\n\\put(390,20){\\circle*{3}}\n\n\\put(250,40){\\circle*{3}}\n\\put(270,40){\\circle*{3}}\n\\put(310,40){\\circle*{3}}\n\\put(330,40){\\circle*{3}}\n\\put(370,40){\\circle*{3}}\n\\put(390,40){\\circle*{3}}\n\n\\put(250,20){\\line(0,1){20}}\n\\put(270,20){\\line(0,1){20}}\n\\put(320,20){\\oval(20,12)[t]}\n\\put(320,40){\\oval(20,12)[b]}\n\\put(370,20){\\line(0,1){20}}\n\\put(390,20){\\line(0,1){20}}\n\n\\put(285,25){$\\cdots$}\n\\put(345,25){$\\cdots$}\n\n\\put(248,5){\\scriptsize $1$ }\n\\put(268,5){\\scriptsize $2$ }\n\\put(308,5){\\scriptsize $i$ }\n\\put(326,5){\\scriptsize $i+1$ }\n\\put(362,5){\\scriptsize $n-1$ }\n\\put(390,5){\\scriptsize $n$ }\n\n\\end{picture}\n\\end{center}\n\nThe subalgebra of $\\Bc_n(\\om)$ generated over $\\CC$\nby $s_1,\\dots,s_{n-1}$\nis isomorphic to the group algebra $\\CC[\\Sym_n]$ so that $s_i$\ncan be identified with the transposition $(i,i+1)$.\nThen for any $1\\leqslant i