diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzmgla" "b/data_all_eng_slimpj/shuffled/split2/finalzzmgla" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzmgla" @@ -0,0 +1,5 @@ +{"text":"\\section{INTRODUCTION}\n\nAn experimental study of relativistic heavy ion collisions augments\nour understanding of the QCD phase diagram~\\cite{cite:QCD_Diagram}.\nThe high energy density reached in such collisions at the Relativistic\nHeavy Ion Collider (RHIC) is believed to result in a novel state of\nhot and dense matter with properties strikingly different from that\nof a hadron gas or ordinary nuclear matter~\\cite{cite:STARWitePaper}. \n\nThe bulk properties of particle production are studied using\nidentified particle spectra at low momentum. Model-dependent\ninterpretations of the measured data provide insight into the\ncomplex dynamics of the collision and further explore the QCD\nphases through which the collision evolves. The dense system\nformed in the early stages of the collision continuously\nexpands and cools, until kinetic freeze-out, beyond which the\nparticles stream freely into the detector. Through\nmeasurements of species abundances and transverse momentum\ndistributions, information about the final stages of the\ncollision evolution at chemical and kinetic freeze-out can\nbe inferred.\n\nThe relative particle abundances and spectral shapes discussed\nhere were tested within the frameworks of statistical (chemical\nequilibrium)~\\cite{cite:StatModel} and\nBlast-wave~\\cite{cite:BlastWaveModel} models. In the chemical\nequilibrium model, particle abundances relative to the total system\nvolume (assumed to be the same for all particle species) are\ndescribed by system temperature at freeze-out, the baryon and\nstrangeness chemical potentials, and the strangeness\nsuppression factor. The Blast-wave model describes spectral shapes\nassuming a locally thermalized source with a common transverse\nflow velocity field. The results from Au+Au collisions at\n\\snn=200 and at\n62.4~GeV~\\cite{cite:200spectra,cite:62spectra,cite:LongPRC} have\nshown that the chemical freeze-out temperature, $T_{\\rm ch}$, has\nlittle dependence on centrality whereas the kinetic freeze-out\ntemperature, $T_{\\rm kin}$, decreases with increasing centrality of the\ncollision. Further, the radial flow velocity, $\\beta$, increases\nwith increasing centrality. The observed changes in $T_{\\rm kin}$ and\n$\\beta$ with centrality are consistent with higher energy and\npressure in the initial state for more central events. On the\nother hand, the centrality independence of the extracted chemical\nfreeze-out temperature indicates that even for different initial\nconditions, collisions always evolve to the same chemical\nfreeze-out. Moreover, the value for the chemical freeze-out\ntemperature in Au+Au is close to the critical temperature,\npredicted by some lattice QCD calculations~\\cite{Tcrit}. This\nsuggests that chemical freeze-out coincides with hadronization\nand therefore $T_{\\rm ch}$ provides a lower limit estimate\nfor the temperature of the prehadronic state~\\cite{cite:Olgaposter}.\nThe systematic behavior of the kinetic freeze-out properties with\ncharged hadron multiplicity appears to follow the same trend for\nall energies and systems at RHIC~\\cite{cite:200spectra,cite:62spectra}.\nIn this paper, the systematic studies of the QCD phase diagram\nfrom heavy ion collisions are enriched by the addition of new\nRHIC data from Cu+Cu collisions at \\snn=200 and 62.4~GeV. \n\n\\section{THE STAR EXPERIMENT}\nThe Cu+Cu data presented here were collected by the STAR experiment\nduring the RHIC 2005 run. Copper nuclei ($^{63}$Cu) were collided\nat \\snn=200 and 62.4~GeV. Data were recorded with a minimum bias\ntrigger obtained from the Beam-Beam Cherenkov counters~\\cite{cite:BBCs}\ncoupled with information from the Zero-Degree Calorimeters~\\cite{cite:ZDCs}.\nThis trigger is found to be sensitive to the top $\\sim$85\\% of\nthe inelastic cross-section. The data studied here correspond to\nthe top 60\\% of the inelastic cross-section (minimum bias) where\nlittle or no inefficiency of the triggering or vertex reconstruction\nis found. These 0-60\\% minimum bias events, with 24~M and 10~M\nevents recorded at 200 and 62.4~GeV, respectively, were divided\ninto six centrality bins each corresponding to a 10\\% interval of\nthe geometric cross section. Transverse mass distributions for\ncharged pions, charged kaons, protons and antiprotons, previously\nreported by STAR for \\snn=200 GeV $pp$ collisions and Au+Au\ncollisions at 62.4 and\n200~GeV~\\cite{cite:200spectra,cite:62spectra,cite:LongPRC}, are used\nfor comparison.\n\n\\begin{figure*}[!ht]\n\\includegraphics[width=0.475\\textwidth]{DrawDedxvsp}\n\\includegraphics[width=0.475\\textwidth]{DrawZ_pi}\n\\caption{\\label{fig:dedx} The left panel shows the truncated mean\nionization energy loss ($\\langle dE\/dx \\rangle$) in the TPC as a \nfunction of transverse momentum for positively charged tracks from\n200~GeV Cu+Cu collisions. The right panel shows $Z(\\pi)$, the\nlogarithm of the measured $\\langle dE\/dx \\rangle$ divided by the\ntheoretical expectation for energy loss of charged pions, for\n$0.40 < \\pT < 0.55$~GeV\/$c$. Also shown is an example four-Gaussian\nfit that is used to extract the raw yields for different species.}\n\\end{figure*}\n\n\\begin{figure}[!ht]\n\\includegraphics[width=0.475\\textwidth]{pBg_Central_200}\n\\vspace*{-0.3cm}\n\\caption{\\label{fig:pBack} Estimated fraction of background protons in\nthe raw proton sample as function of transverse momentum, for the most\ncentral \\snn=200~GeV Cu+Cu collisions. No strong centrality or energy\ndependence for this correction was observed for all Cu+Cu data available.}\n\\end{figure}\n\n\\begin{figure*}[!ht]\n\\includegraphics[width=0.95\\textwidth]{MinusSpectraCuCu200_and62_SplitPanels.eps}\n\\vspace{-0.5cm}\n\\caption{\\label{fig:SpectraMinus}The top row shows negatively charged pion\n(leftmost column), kaon (center) and anti-proton (right) spectra from Cu+Cu\ncollisions at \\snn=200~GeV. Six centrality classes are shown as dark\n(central 0-10\\%) to light (50-60\\%) shades. \nSpectra for 62.4~GeV Cu+Cu are shown on the bottom row. Statistical and\nsystematic errors (which do not exceed 7\\%) are smaller than the symbol size.}\n\\end{figure*}\n\n\\begin{figure*}[!ht]\n\\includegraphics[width=0.95\\textwidth]{PlusSpectraCuCu200_and62_SplitPanels.eps}\n\\vspace{-0.8cm}\n\\caption{\\label{fig:SpectraPlus}The top row shows positively charged pion\n(leftmost column), kaon (center) and proton (right) spectra from Cu+Cu\ncollisions at \\snn=200~GeV. Six centrality classes are shown as dark\n(central 0-10\\%) to light (50-60\\%) shades. \nSpectra for 62.4~GeV Cu+Cu are shown on the bottom row. Statistical and\nsystematic errors (which do not exceed 7\\%) are smaller than the symbol size.}\n\\end{figure*}\n \n\\begin{figure*}[!ht]\n\\includegraphics[width=0.98\\textwidth]{DrawDiffFits2_Plus200_SplitPanels.eps}\n\\caption{\\label{fig:SpecFits}\nComparison between the 10\\% most central Cu+Cu collision data (symbols)\nat \\snn=200~GeV and the corresponding Blast-wave model fit (dashed\nline) to $\\pi^{+}$ (left), K$^{+}$ (center) and proton (right) spectra\n-- note fit is performed simultaneously across species. The pion data\npoints below \\pT~=~0.5~GeV\/$c$ were not included in the Blast-wave fits\nto reduce the effect of resonance decays.\nA Bose-Einstein fit to the pion spectra over the entire\nfiducial range is also shown. The lower panels illustrate the quality\nof the fits by showing the difference between the measured points and\nthe fit expressed as the number of standard deviations.}\n\\end{figure*} \n\nThe STAR Time Projection Chamber (TPC)~\\cite{cite:STAR_TPC} tracks\nparticle trajectories over a wide range of momentum at mid-rapidity\n($|\\eta| < 1.8$). The particle identification at low-\\pT~uses\nmeasurements of truncated mean ionization energy loss,\n$\\langle dE\/dx \\rangle$, of the charged particles traversing the\nTPC. Particles of different mass show distinct patterns in the\n$\\langle dE\/dx \\rangle$ dependence, as shown in Fig.~\\ref{fig:dedx},\nleft panel. This allows statistical separation of pions and kaons\nin the momentum range $0.25 < \\pT < 0.80$~GeV\/{\\em c} at mid-rapidity\n($|y| < 0.1$), and of protons and anti-protons from other species in\nthe range $0.40 < \\pT < 1.20$~GeV\/{\\em c}.\n\nThe momentum measurement is given by the curvature of the particle\ntrajectories as they pass through the 0.5~T magnetic field of the\nSTAR detector. \nTo ensure optimal $dE\/dx$ resolution, only primary tracks, with {\\em dca}\n(distance of closest approach between the particle trajectory and the\nevent vertex) less than 3~cm, and at least 25 out of 45 possible fit\npoints are used in this analysis.\nParticle identification at mid-rapidity ($|y| < 0.1$)\nis achieved by fits to the $Z$-variable, defined as a logarithm of\n$\\langle dE\/dx \\rangle$ divided by the theoretically expected value\nfor each particle type, given by Bethe-Bloch formula~\\cite{cite:BB}.\nThis new variable is introduced to remove the\nstrong \\pT~dependence at low momenta. Such a normalized distribution\nis created for a given particle and centrality and is divided into\nnarrow transverse momentum slices (width $\\Delta$\\pT~=~50~MeV\/{\\em c}).\nThese momentum projections are fit with a combined four-Gaussian\nfunction, one for each of the particle species of a given charge:\n$\\pi$, K, $p$ and $e$. \nThe integral of each Gaussian provides the raw yield at each momentum.\nThis procedure is repeated for each particle species in order to assign\nthe correct rapidity for each track, using the mass of the particle.\nThus, fits to the auxiliary particles in each distribution (for example\nK, $p$ and $e$ for $\\pi$ analysis) are used only to estimate the\ncontamination when bands overlap. The right panel of Fig.~\\ref{fig:dedx}\nshows an example for pion yield extraction for one momentum slice. For\nmore details see Ref.~\\cite{cite:LongPRC}.\n\nThe raw yields extracted from each of the four-Gaussian fits are then\ncorrected for detector acceptance, single-track reconstruction\nefficiencies, and other effects as discussed below. To determine the\ncorrection factors, simulated tracks were embedded into real data on\nthe raw signal level and run through the standard reconstruction chain.\nThe estimated single-track reconstruction efficiency is about $80\\%$\nfor $\\pi^{\\pm}$ in Cu+Cu collisions and exhibits a small centrality\nand $p_T$ dependence. The \\pT-spectrum has also been corrected for the\nenergy loss\nby multiple scattering beyond that for pions which is calculated\nduring reconstruction. This affects the reconstructed momentum at\nlow values. The maximum value of this mass-dependent correction to\nthe measured \\pT~value for ${\\rm K}^{\\pm}$ and $p$($\\overline{p}$) was found\nto be $2\\%$ and $3\\%$, respectively, for the lowest measured \\pT~bin. \nAn additional correction for the background contamination in the proton\nsample is made. The background protons arise predominantly from\nsecondary interactions in the beam pipe and detector material\n(knock-out protons). It is estimated from data to be about 40\\% at\n\\pT~=~400~MeV\/{\\em c}, diminishing to near zero at \\pT~=~1~GeV\/{\\em c},\nas shown in Fig.~\\ref{fig:pBack}.\nTo estimate this correction factor we compare the distribution of proton\n{\\em dca}\nto that of the anti-protons (see\nRef.~\\cite{cite:LongPRC} for more details). In the measured {\\em dca}\nregion, the integral difference between protons and anti-protons (after\nnormalization by the anti-proton to proton ratio) is considered to be\nthe background contribution to the proton yield.\nPion yields are additionally corrected for feed-down contributions from\nweakly decaying particles, muon contamination, and background pions\nfrom detector material. This correction is found to decrease from about\n$15\\%$ at 0.3~GeV\/$c$ to about $5\\%$ at 1~GeV\/{\\em c}.\nThe (anti)protons presented in this paper are inclusive measurements\n(not corrected for weak decays). It has been found in previous studies\nthat the analysis cuts used for the low-\\pT~identified proton studies\n($dca < 3$~cm) reject only a negligible fraction of daughter protons from\nthe hyperon decays~\\cite{cite:200spectra}. Therefore, our sample \nreflects the total baryon production in the collision. Earlier Au+Au\nstudies~\\cite{cite:200spectra} and preliminary Lambda-hyperon spectra\nfrom Cu+Cu collisions~\\cite{cite:AntStrange} indicate that the\nfreeze-out spectral shapes are similar for $\\Lambda$s and protons,\nresulting in similar spectra shapes for primary and feed-down protons.\nThe fraction of the weak-decay feed-down protons is estimated to be\nabout $30\\%$~\\cite{cite:JuneWWND08}.\n\nThis analysis technique is used to obtain the low-\\pT~particle spectra\nfor all centrality bins at both 200 and 62.4~GeV center-of-mass\nenergies and for the Cu+Cu and Au+Au colliding systems. Additional\ntechnical details on the analysis and applied corrections can be\nfound in Refs.~\\cite{cite:200spectra,cite:62spectra} with a thorough\noverview in Ref.~\\cite{cite:LongPRC}.\n\n\n\\section{RESULTS}\n\nThe transverse momentum spectra are shown in\nFigs.~\\ref{fig:SpectraMinus}~and~\\ref{fig:SpectraPlus} for $\\pi^{\\pm}$\n(leftmost column), K$^{\\pm}$ (center) and (anti)protons (right) in\nCu+Cu collisions. The top row presents the data for \\snn=200~GeV, whilst \ndata for 62.4~GeV are shown in the bottom row. The symbol shades represent\ndifferent centrality bins. The particle and anti-particle spectral\nshapes are similar for all species in each centrality bin. At both\ncollision energies a mass dependence is observed in the slope of the\nparticle spectra. Due to the large number of events recorded and good\ntracking efficiency, the statistical errors are less than 1\\%. The\nsystematic uncertainties are similar to those determined in prior\nanalyses of low-\\pT~spectra in Au+Au collisions~\\cite{cite:LongPRC}.\nSystematic errors are divided into two classes: point-to-point and\nscale uncertainties. The overall scale uncertainty, mostly due to the\nembedding procedure for the single-track reconstruction efficiency, is\nestimated to be 5\\% for all particle species. Point-to-point\nuncertainties are determined for each \\pT~bin and particle species.\nFor pions and kaons, this error is evaluated to be less than 7\\% and\n13\\%, respectively. These maximal errors represent \\pT~bins where a\nsignificant $\\langle dE\/dx \\rangle$ overlap occurs between $\\pi^{\\pm}$,\nK$^{\\pm}$ or $e^{\\pm}$. For protons and anti-protons, the maximum\nerror is 5\\%. At low-\\pT, the proton uncertainty is greater than that\nfor anti-protons (4.0\\% versus 1.3\\%, respectively, at \n\\pT~=~400-450~MeV\/$c$) owing to the additional uncertainty from the\nproton's background. The uncertainty due to the background decreases\nrapidly from 3.7\\% at \\pT~=~400-450~MeV\/$c$ to 1.5\\% for $\\pT > 1$~GeV\/$c$.\n\nFor the anti-particle to particle yield ratios, systematic errors are\nmuch reduced due to a cancellation of the efficiency uncertainties\nand a partial cancellation of extrapolation uncertainties, as described\nabove. A systematic uncertainty of 2\\%, 3\\%, and 5\\% is assigned to\n$\\pi^{-}\/\\pi^{+}$, K$^{-}$\/K$^{+}$, $\\overline{p}\/p$, respectively.\n\nWe further fit the obtained \\pT~distributions to extract system\nproperties at different stages of the collision evolution. The first\nfit to the data probes collision properties at kinetic freeze-out.\nHere, a Blast-wave model~\\cite{cite:BlastWaveModel} is used to\nsimultaneously fit the $\\pi^{\\pm}$, K$^{\\pm}$ and (anti)proton spectra\nat a given centrality. This fit provides a good description of the\nspectra shapes, as illustrated in Fig.~\\ref{fig:SpecFits} with results\nfrom most central 200~GeV Cu+Cu data. The $\\pi^{\\pm}$ data points for\n$\\pT < 0.5$~GeV\/$c$ are excluded from the Blast-wave fits to reduce the\neffects of resonance decay contributions as done in previous\nworks~\\cite{cite:200spectra,cite:62spectra,cite:LongPRC}. Including\nthis low-\\pT~region in the fit leads to a poorer description of proton\nand kaon shapes, however the resultant modification of the extracted\nparameters remains well within their systematic uncertainty. The\nfreeze-out parameters obtained from this model are discussed later.\nAlso shown in Fig.~\\ref{fig:SpecFits} are Bose-Einstein\n($\\propto 1\/(\\exp{\\frac{\\mT}{T}}-1)$) fits to the $\\pi^{\\pm}$, which\nprovide a slightly better description of these data. For evaluation\nof the systematic uncertainties from extrapolation, $m_{T}$-exponential\n($\\propto 1\/\\exp{\\frac{\\mT}{T}}$) and Boltzmann\n($\\propto \\mT\/\\exp{\\frac{\\mT}{T}}$) fits are also used in the\nanalysis (for more details see Ref.~\\cite{cite:LongPRC}).\n\nThe particle mean-\\pT~and total particle yields at mid-rapidity\n($|y| < 0.1$) are shown in Fig.~\\ref{fig:MeanPt} and Fig.~\\ref{fig:dN\/dy},\nrespectively. The values presented for kaons and (anti)protons are\ndetermined from the measured spectra points, extrapolated outside the\nfiducial range using Blast-wave fits discussed above. Similarly, a\ncombination of the measured data-points and extrapolation from\nBose-Einstein fits is used for the pions. The measured fraction of\nthe total yield is found to be 62\\% for $\\pi^{\\pm}$, 58\\% for K$^{\\pm}$\nand 65\\% for (anti)protons for the most central 200 GeV data; these\nfractions are slightly higher in other centrality bins and at lower\nenergy~\\cite{cite:STAR9GeV}. The systematic uncertainty on $dN\/dy$ and\nmean-\\pT, shown in\nthe figures, includes the extrapolation uncertainty evaluated by means\nof various model fits mentioned earlier. Overall, these are estimated\nto be near 15\\% of the yields outside the fiducial range for pions and\nkaons and 15\\%-25\\% of the extrapolated yields for protons and\nanti-protons, depending on centrality.\n\nWe also determine the total charged hadron production per unit of\npseudo-rapidity, \\nch, at mid-rapidity. The total particle yield at\nmid-rapidity for each species, obtained by extrapolating the fits to\nthe measured spectrum in the momentum range outside our fiducial\ncoverage, was corrected for the Jacobian transformation\n\\nchy$\\rightarrow$\\nch. The sum of the total charged pion, kaon and\n(anti)proton yields was then corrected for the feed-down of weakly\ndecaying neutral strange particles, providing the estimate of\nprimordial charged hadron yield at mid-rapidity. A complementary\nmethod was also used, integrating over the charged hadron spectra\ncorrected for efficiency, feed-down and the Jacobian transformation,\nand yielded consistent results.\n\nThe mean-\\pT~of each particle species ($\\pi$, K, $p$) increases with\nthe number of charged hadrons at mid-rapidity \\nch, as shown in\nFig.~\\ref{fig:MeanPt}. Moreover, the mean-\\pT~for each particle\nspecies appears to scale with \\nch~at mid-rapidity, and to be\nindependent of the colliding system and the center-of-mass energy. \nThe particle yields show the same systematic scaling features with\n\\nch~as mean-\\pT~across system and collision energy, see\nFig.~\\ref{fig:dN\/dy}. In this logarithmic representation the particle\nyields for each species appear to increase linearly with multiplicity,\nwith Cu+Cu matching the Au+Au data at similar values of \\nch. When\nshown on a linear scale, the integrated yields exhibit a near-linear\ndependence with \\nch. The logarithmic scale for both axes used here\npreserves the apparent linear dependencies whilst better illustrating\nthe lower multiplicity Cu+Cu data. For the detailed features we\ninvestigate the relative particle production in the following.\n\n\\begin{figure}[!t]\n\\includegraphics[width=0.475\\textwidth]{DrawMeanPt_OnePanel.eps}\n\\caption{\\label{fig:MeanPt} Mean transverse momentum as a function of\ncharged hadron multiplicity at mid-rapidity for pions, kaons and\nanti-protons. For comparison, the mean-\\pT~values for Au+Au data are\nshown as bands. Open (closed) symbols\/bands depict data at\n\\snn=200~GeV (62.4~GeV). Error-bars represent statistical and systematic\nuncertainties added in quadrature. }\n\\end{figure}\n\n\\begin{figure}[!t]\n\\includegraphics[width=0.475\\textwidth]{DrawYield_ThreePanels_Log.eps}\n\\caption{\\label{fig:dN\/dy} Integrated yields at mid-rapidity for pions, kaons and\nanti-protons as a function of the charged particle density (\\nch),\nwhich is used as a measure of the centrality. For comparison Au+Au\ndata are shown as bands. Filled (open) points\/bands depict data at\n\\snn=200~GeV (62.4~GeV). Error-bars represent statistical and systematic\nuncertainties added in quadrature.} \n\\end{figure}\n\n\\begin{figure}[!t]\n\\includegraphics[width=0.475\\textwidth]{DrawKtoPiandPtoPiMultiPanel2.eps}\n\\caption{\\label{fig:Ratios} Integrated particle yield ratios at\n\\snn=200~GeV (closed symbols) and 62.4~GeV (open) for Cu+Cu (black)\nand Au+Au collisions (grey bands) versus \\nch~at mid-rapidity. Error-bars\nrepresent statistical and systematic uncertainties added in quadrature.}\n\\end{figure}\n\nThe relative abundances of particles provide an important insight\ninto the chemical properties of the system. The relative kaon yield\nreflects the strangeness production in the collision, whereas proton\nwith respect to pion production is dependent on the baryon production\nand transport. Figure~\\ref{fig:Ratios}($a$) shows the ratios for the\nnegatively charged particles, $\\overline{p}\/\\pi^{-}$ and ${\\rm K}^{-}\/\\pi^{-}$,\nas a function of \\nch, which exhibit similar \\nch-scaling behavior at\neach collision energy. \nThe slight decrease of the values for both ratios seen at the\nlower collision energy of\n 62.4~GeV is insignificant within\nexperimental uncertainties. Figure~\\ref{fig:Ratios}($b$) shows the\nratios for positively charged particles, $p\/\\pi^{+}$ and ${\\rm K}^{+}\/\\pi^{+}$,\nwhich also exhibit a \\nch-scaling behavior within the same collision\nenergy. The beam-energy effect is reversed here as compared to the ratio\nof negatively charged particles. Summing over the two charges\n(Fig.~\\ref{fig:Ratios}($c$)), the corresponding ratios exhibit a common\nscaling behavior with \\nch, independent of colliding system and collision\nenergy. The energy dependence of the positive and negative particle\nratios considered separately, points to the effects of baryon transport\nto mid-rapidity, which decreases with increasing energy (see also~\\cite{cite:STAR9GeV}).\n\nWe further explore the kaon production in Cu+Cu collisions to gain\nbetter insight into production of strange quarks. The ${\\rm K}^{-}\/\\pi^{-}$\nratio scales with \\nch~at both energies and there is no hint of an\nadditional strangeness enhancement of charged kaons in the smaller\nCu+Cu system compared to the larger Au+Au system. Early works from\nSPS energies reported such additional relative strangeness\nenhancement in the ${\\rm K}\/\\pi$ ratio for smaller systems, although no\nfinal confirmation of this observation is\navailable~\\cite{cite:SPS_NA49QM02,cite:SPS_NA49PRL}. The pion and\nkaon enhancement factors are compared in Fig.~\\ref{fig:Enhancement}.\nThis factor is defined as the yield per mean number of participating\nnucleons (estimated using a Glauber model), \\npart, in heavy-ion\ncollisions divided by the respective value in \\pp~collisions. A\nprogressive enhancement of kaon production with respect to pions as\na function of collision centrality is evident, as shown earlier by\nthe K\/$\\pi$ ratios (Fig.~\\ref{fig:Ratios}($a$) and\nFig.~\\ref{fig:Ratios}($b$)). A comparison of these enhancement\nfactors between Cu+Cu and Au+Au data is also shown. The enhancement\nfactors for kaons do not show universal scaling features with\nrespect to \\npart, and are indeed found to be higher in Cu+Cu\ncollisions compared to the Au+Au system. However, these features do\nnot appear to be unique to kaons. A similar trend is observed in\nFig.~\\ref{fig:Enhancement} in the\npion enhancement factors for the two systems. This suggests that the\nadditional enhancement, seen in the charged kaon yields, is not\nrelated to strangeness production, but other physics mechanisms, for\nexample, additional entropy production. It should be noted that while\ncomparing more spherical central Cu+Cu collisions with semi-peripheral\nAu+Au collisions, the initial conditions may not be reflected by\n\\npart~alone. \n\n\\begin{figure*}[t]\n\\includegraphics[width=0.85\\textwidth]{DrawYield_per_pp.eps}\n\\caption{\\label{fig:Enhancement} Enhancement factors for negatively\ncharged pions (left), kaons (center), and anti-protons (right) as\nfunction of \\npart~in \\snn=200 Cu+Cu and Au+Au collisions.\nError-bars represent statistical and systematic uncertainties on the\nA+A measurements added in quadrature. The shaded bands depict model\nuncertainties on number of participants calculation. The bands on the\nleft show uncertainties from the $pp$ measurements that are correlated\nfor all data points. }\n\\end{figure*}\n\n\nIn contrast to pions and kaons, protons show minimal evolution with\ncentrality and no difference between Cu+Cu and Au+Au systems is\nobserved. Figure~\\ref{fig:Ratios}($d$) illustrates the difference in\nanti-proton and proton production across energies. As observed in\nother collision systems, the ratio is found to increase and becomes\ncloser to unity for higher energy\ncollisions~\\cite{cite:PHOBOS_ppPartRatios}. The anti-proton to\nproton ratio gives information on the amount of baryon transport.\nIn line with the earlier STAR results, our measurements indicate\nthat while a finite excess of baryons over anti-baryons is still\npresent at RHIC energies, $p-\\bar{p}$ pair production becomes an\nimportant factor. Little or no change due to an increase in the\nsystem size (centrality) is apparent in the Cu+Cu data at 200~GeV,\nwhile 62.4~GeV data show a decreasing trend with increasing centrality\nfor this ratio for both Cu+Cu and Au+Au data.\n\n\\section{FREEZE-OUT PROPERTIES}\n\nThe particle yields and their ratios provide further information on the\nthermal properties of the system at kinetic and chemical freeze-out.\n\n\\subsection{Kinetic Properties}\nThe completion of all elastic scattering marks the final stage of\ncollision evolution and could be interpreted as a kinetic freeze-out,\nwhere the particle momentum spectra are fixed. To quantify this\nstage, fits are made simultaneously to the spectra of all particle\nspecies, but independently for each centrality class (see\nFig.~\\ref{fig:SpecFits} for example). The fits used here are based\non the previously discussed Blast-wave model~\\cite{cite:BlastWaveModel},\nwhich assumes a radially boosted thermal source. These\nhydrodynamically-motivated fits describe the mass dependence of\nparticle spectral shapes in terms of the radial flow velocity ($\\beta$),\nthe kinetic freeze-out temperature ($T_{\\rm kin}$) and the flow velocity\nprofile exponent ($n$) at the final freeze-out. The extracted value for\n$n$ is not used to derive any physics interpretation. The effects from\nresonance contributions to the pion spectral shape are reduced by\nexcluding the low-\\pT~data points (below 0.5~GeV\/{\\em c}). To enable a\ncomparison with earlier results on \\pp~and Au+Au\ncollisions~\\cite{cite:200spectra,cite:LongPRC}, the same model and the\nsame procedures for the fits are adopted, thereby avoiding any possible\nsystematic bias.\n\n\\begin{figure*}[t]\n\\includegraphics[width=0.95\\textwidth]{DrawKineticFreezeOut_3Panel.eps}\n\\caption{\\label{fig:KineticFO} Comparison of kinetic freeze-out\nproperties obtained from fits to Cu+Cu (symbols) and Au+Au (bands)\ncollision data at \\snn=200 (closed symbols\/bands) and 62.4~GeV (open).\nThe kinetic freeze-out temperature, $T_{\\rm kin}$, is shown versus flow\nvelocity, $\\beta$, and multiplicity in panels (a) and (b) respectively\n(more central collisions are to the right side of each plot).\nPanel (b) also shows the multiplicity dependence of the chemical\nfreeze-out temperature, $T_{\\rm ch}$, (square symbols). Panel (c) shows\nthe multiplicity dependence of the average radial flow velocity.}\n\\end{figure*}\n\n\nThe Blast-wave fit results for the temperature of freeze-out are shown in\nFig.~\\ref{fig:KineticFO}. $T_{\\rm kin}$ and $\\beta$ show similar\ndependences as a function of \\nch~in both Cu+Cu and Au+Au collisions,\nevolving smoothly from the lowest to the highest multiplicity, from \\pp~to\ncentral Au+Au. $T_{\\rm kin}$ decreases smoothly with centrality implying\nthat freeze-out occurs at a lower temperature in more central collisions. \nThe similarity of kinetic freeze-out parameters in the events with similar\nmultiplicity from different colliding species is confirmed by the data alone.\nAs noted earlier, the particle mean-\\pT~increases with increasing \\nch,\nwhich is consistent with an increase of radial flow with centrality. \nWe note, however, that other physics mechanisms, for example, hard and\nsemi-hard scatterings, can contribute to higher mean-\\pT~values observed\nfor kaon and proton spectra~\\cite{cite:Trainor}. Direct spectral shape\ncomparisons of Cu+Cu and Au+Au events from similar multiplicity bins, shown\nin Fig.~\\ref{fig:CompareSpectra}, show the same \\pT-dependencies between\npion spectra from the two systems. The same is seen to hold for the\nrespective kaon and proton spectra. The middle panel of Fig.~\\ref{fig:KineticFO}\nshows in addition the chemical\nfreeze-out temperatures for different\ncolliding systems at different energies. Both the chemical freeze-out\nand the kinetic freeze-out temperature show similar scaling features,\nreflecting the common trends in mean-\\pT~and the ratios of $p\/\\pi$ and\n${\\rm K}\/\\pi$, discussed earlier. Similarly, on the left panel of\nFig.~\\ref{fig:KineticFO} we observe a\ncommon \\nch-dependence for the average radial flow velocity at kinetic\nfreeze-out. \n\nA more important observation is that the obtained kinetic freeze-out\nparameters for pions, kaons and (anti)protons follow the same trends\nwith \\nch, independent of collision energy, even though the production\ncross-sections of the underlying spectra are different. This observation\nsuggests that the kinetic freeze-out properties are determined by the\ninitial state. Furthermore, a model-dependent connection between the\nnumbers of produced charged particles and the initial gluon density of\nthe colliding system~\\cite{cite:CGC} can be used to deduce that the\nfreeze-out properties are most probably determined at the initial\nstages of the collision and are driven by the initial energy density.\n\n\\begin{figure*}[t]\n\\includegraphics[width=0.95\\textwidth]{MinusSpectraComp_SplitPanelsLin.eps}\n\\caption{\\label{fig:CompareSpectra}\nComparison of the spectral shape between Cu+Cu and Au+Au data at\n\\snn=200~GeV. Centrality classes are chosen with a similar average\ncharged hadron multiplicity at mid-rapidity. Pion (left), kaon\n(center) and anti-proton (right) spectra are shown for 10-20\\% central\n(40-50\\%) Cu+Cu (symbols) compared to 40-50\\% mid-peripheral (60-70\\%)\nAu+Au (lines).}\n\\end{figure*}\n\n\\subsection{Chemical Properties}\n\n\\begin{figure}[!ht]\n\\includegraphics[width=0.475\\textwidth]{ratiosChem2}\n\\caption{\\label{fig:ChemicalFit200}\nThe upper panel shows statistical model fit predictions (grey lines)\nfor the measured particle ratios (circles) from central 200~GeV Cu+Cu\ncollisions. The lower panel illustrates the fit quality by showing\nthe difference between the measured data and the model prediction in\nterms of the number of standard deviations ($N_{\\sigma}$) determined\nby systematic (data) uncertainty.}\n\\end{figure}\n\n\\begin{figure}[ht]\n\\vspace{0.5cm}\n\\includegraphics[width=0.475\\textwidth]{DrawChemicalFreezeOutMuBNch.eps}\n\\caption{\\label{fig:ChemicalFO2} Baryon and strangeness chemical potentials, $\\mu_{\\rm B}$ and $\\mu_{\\rm S}$,\n as a function of \\nch~for 200\nand 62.4~GeV in Cu+Cu (symbols) and Au+Au collisions (bands). }\n\\end{figure}\n\n\\begin{figure*}[t]\n\\includegraphics[width=0.95\\textwidth]{DrawChemicalFreezeOut.eps}\n\\caption{\\label{fig:ChemicalFO} Left: chemical freeze-out temperature,\n$T_{\\rm ch}$, as function of the baryon-chemical potential, $\\mu_{\\rm B}$,\nderived for central Au+Au (0-5\\% for 200 and 62.4 GeV~\\cite{cite:LongPRC} \nand 0-10\\% for 9.2 GeV~\\cite{cite:STAR9GeV}) and Cu+Cu (0-10\\%) collisions.\nFor comparison, results for minimum bias \\pp~collisions at 200~GeV are also\nshown along with additional heavy-ion data points compiled for lower\ncollision energies~\\cite{cite:Claymans}. The dashed line represents\na common fit to all\navailable heavy-ion data described in the text. Right: strangeness\nsuppression factor, $\\gamma_{\\rm S}$, as a function of \\nch~for 200\nand 62.4~GeV in Cu+Cu (symbols) and Au+Au collisions (bands). }\n\\end{figure*}\n\n\n\nChemical freeze-out occurs at the stage of the collision when all\ninelastic interactions cease and the produced particle composition in\nterms of yields is fixed. Valuable information for this collision\nstage can be obtained directly from the experimental results by\nforming particle ratios and comparing them across different collision\nsystems and energies.\n\nThe ratios of the different particle yields in Cu+Cu collisions\nare further analyzed within the framework of the statistical\nmodel~\\cite{cite:StatModel}. This model describes the chemical\nfreeze-out of the colliding system by several fit\nparameters: the {\\em temperature} at which freeze-out occurs\n($T_{\\rm ch}$), the {\\em cost} of producing matter in terms of\nbaryon and strangeness chemical potentials ($\\mu_{\\rm B}$,\n$\\mu_{\\rm S}$), and an additional {\\it ad-hoc} parameter, known as\nthe strangeness suppression factor, ($\\gamma_{\\rm s}$), to reconcile the\nlower yield of strange hadrons in collisions\ninvolving smaller species (for example \\pp~and d+Au).\n\nThese statistical fits are performed on the relative particle\nabundances from $\\pi^{\\pm}$, K$^{\\pm}$ and $p$($\\overline{p}$) alone.\nFigure~\\ref{fig:ChemicalFit200} shows an example of the resultant fit\nto the identified hadron ratios from central \\snn=200~GeV Cu+Cu\ncollisions. The lower panel of this figure illustrates the fit\nquality. We note, that the successful description of the ratios by the\nmodel could not prove the attainment of chemical equilibrium, but\nsuggests the statistical nature of particle production in these\ncollisions~\\cite{cite:prc126}. The results obtained for the\nfreeze-out parameters are shown in\nFigs.~\\ref{fig:KineticFO}, \\ref{fig:ChemicalFO2}~and~\\ref{fig:ChemicalFO}. \n\n\nStatistical model fits to a wider variety of hadron yields were also\nattempted using preliminary results for the $\\Lambda$, $\\Xi$ and\n$\\phi$ particles and anti-particles from 200~GeV Cu+Cu data\nfrom~\\cite{cite:AntStrange}. Including more particles into the model\nfits reduces the systematic uncertainty on the extracted parameters\nand resulted in parameter values consistent with those obtained from\nfits using $\\pi^{\\pm}$, K$^{\\pm}$ and $p$($\\overline{p}$) alone\nreported here. In general, the observed systematic trends in the\nfreeze-out parameters as a function of the collision centrality are\npreserved ~\\cite{cite:LongPRC, cite:STARWitePaper}.\n\nFigure~\\ref{fig:ChemicalFO} (left panel) shows the evolution of the chemical\nfreeze-out temperature versus baryon chemical potential in central\nheavy-ion collisions from the very low energy SIS data through AGS\nand SPS to RHIC (STAR data points only). The overall evolution of\n$T_{\\rm ch}$ can be reproduced by the phenomenological model\nfit~\\cite{cite:Claymans} applied here to all the data points shown\n(dashed line). As the collision energy increases,\n the temperature at freeze-out is found to increase up to\n SPS energies. This\nis followed by a plateau at RHIC energies at a value close to that\nof the hadronization temperature expected from lattice QCD\ncalculations. At RHIC, for all systems and center-of-mass energies,\n$T_{\\rm ch}$ appears to be universal, as shown in\nFig.~\\ref{fig:KineticFO} (middle panel).\n\nThe value of the baryon-chemical potential at a given center-of-mass\nenergy is found to be slightly higher for the larger system, with Au+Au\nand Cu+Cu measurements showing common trends with charged hadron\nmultiplicity (Fig.~\\ref{fig:ChemicalFO2}). We note that, presented in\nthe same figure, values of strangeness chemical potential are close to\nzero with no obvious systematic trends for all energies and colliding\nsystems studied at RHIC.\nWithin a given system, $\\mu_{\\rm B}$ reflects the decrease in net-baryon\ndensity with increasing collision energy from \\snn=62.4 to 200~GeV.\nThis behavior can be observed directly from the particle ratios,\nwhere $\\overline{p}\/p$ increases as a function of energy\n(Fig.~\\ref{fig:Ratios}). For the most central Cu+Cu events we measure\n$\\overline{p}\/p = 0.80\\pm0.04$ at 200~GeV, and $0.55\\pm0.03$ at 62.4~GeV.\nThe lack of centrality dependence in the baryon to meson ratios in\nCu+Cu and Au+Au data, points to similar freeze-out temperatures for\nthe studied systems. The constant values of $T_{\\rm ch}$ at RHIC\nenergies for collisions with different initial conditions, energy and\nnet-baryon density, points to a common hadronization temperature of\nthe systems. \n\nAnother parameter extracted from the fit, which is related to\nstrangeness production, is the strangeness suppression factor,\n$\\gamma_{\\rm s}$, shown versus \\nch~ in Fig.~\\ref{fig:ChemicalFO}.\nThe suppression of strange hadron yields is observed\nin smaller systems, such as \\pp~and peripheral collisions. Within\nstatistical models this can be explained by a reduced production\nvolume~\\cite{cite:StrangenessSupp}. At low beam energies, where\nequilibration of $s$ quarks with respect to $u$ and $d$ is not\nexpected, the suppression is also seen. We find that within the\nsystematic errors on the fit parameters the strangeness suppression\nfactor in Cu+Cu is consistent with that for Au+Au for the same number\nof charged particles, \\nch. As only charged kaon yields were included\nin the fit, this observation is directly related to an absence of any\nadditional enhancement in ${\\rm K}\/\\pi$ at the same \\nch~in the smaller Cu+Cu \nsystem with respect to the larger Au+Au system as discussed previously.\n \nThe $\\gamma_{\\rm s}$ parameter shows a similar increase with centrality\nfor both systems and energies. The value of $\\gamma_{\\rm s}$ approaching\nunity for the central Au+Au collisions in the context of thermal model\nwould imply that the produced strangeness is close to equilibrium. \n \n\n\n\\section{Summary}\nWe have presented measurements of identified charged hadron spectra\nin Cu+Cu collisions for two center-of-mass energies, 200 and 62.4~GeV.\nThese new results of $\\pi^{\\pm}$, K$^{\\pm}$ and $p$($\\overline{p}$)\nhave further enriched the variety of low-\\pT~spectra at RHIC. The\ndata have been studied within the statistical hadronization and\nBlast-wave model frameworks in order to characterize the properties\nof the final hadronic state of the colliding system as a function of\nsystem size, collision energy and centrality.\n\nThese multidimensional systematic studies reveal remarkable\nsimilarities between the different colliding systems. No additional\nenhancement of kaon yields with respect to pions is observed for the\nsmaller Cu+Cu system compared to Au+Au. The obtained particle ratios,\nmean-\\pT~and the freeze-out parameters, including the strangeness\nsuppression factor, $\\gamma_{\\rm s}$, are found to exhibit a smooth\nevolution with \\nch, and similar properties at the same number of\nproduced charged hadrons are observed for all collision systems and\ncenter-of-mass energies. The bulk properties studied \nhave a strong correspondence with the total particle yield. Within\nthermal models this reflects a relation between the energy per\nparticle at freeze-out and the entropy derived from particle yields,\nwhich reflects the initial state properties for adiabatic expansion.\nThe baryon chemical potential could in addition be influenced by\nthe initial valence quark distribution and by baryon transport\nduring expansion, leading to a more complicated dependence. The\nscaling features of freeze-out properties are not presented at the\nsame \\npart~for lighter and heavier ions as scaling is badly\nbroken when data measured at different energies are compared. This\nsuggests that \\npart~does not reflect the initial state of the system\naccurately.\n\nWe thank the RHIC Operations Group and RCF at BNL, the NERSC Center at LBNL and the Open Science Grid consortium for providing resources and support. This work was supported in part by the Offices of NP and HEP within the U.S. DOE Office of Science, the U.S. NSF, the Sloan Foundation, the DFG cluster of excellence `Origin and Structure of the Universe' of Germany, CNRS\/IN2P3, STFC and EPSRC of the United Kingdom, FAPESP CNPq of Brazil, Ministry of Ed. and Sci. of the Russian Federation, NNSFC, CAS, MoST, and MoE of China, GA and MSMT of the Czech Republic, FOM and NWO of the Netherlands, DAE, DST, and CSIR of India, Polish Ministry of Sci. and Higher Ed., Korea Research Foundation, Ministry of Sci., Ed. and Sports of the Rep. Of Croatia, Russian Ministry of Sci. and Tech, and RosAtom of Russia.\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section*{Setup}\n\n\\subsection*{Fundamentals of viscous streaming}\nViscous streaming \\cite{holtsmark1954boundary,bertelsen1973nonlinear}, an inertial phenomenon, arises when a small feature (typically a solid inclusion or a microbubble) of characteristic length $a$, is immersed in a liquid of kinematic viscosity $\\nu$, that oscillates with angular frequency $\\omega$ and amplitude $A\\ll a$. Inertial rectification, facilitated by stress concentration at the boundaries of the immersed body, results in steady flows that can be effectively leveraged to manipulate fluid, suspended particles, and chemicals \\cite{lutz2003microfluidics, lutz2006hydrodynamic, lutz2005microscopic, parthasarathy2019streaming, bhosale2021multi}. For simple geometries of uniform curvature (cylinders, spheres), viscous streaming flows have been theoretically, computationally, and experimentally well characterized as function of the Stokes layer thickness $\\delta_\\text{AC}\\sim\\mathcal{O}(\\sqrt{\\nu\/\\omega})$ and length-scale $a$ \\cite{holtsmark1954boundary,riley2001steady,lutz2005microscopic,lane1955acoustical,bhosale_parthasarathy_gazzola_2020}. More recently, the use of multi-curvature streaming bodies has revealed rich flow repertoires, furthering opportunities in transport, separation, and assembly \\cite{parthasarathy2019streaming,bhosale_parthasarathy_gazzola_2020,chan2021three,bhosale2021multi,bhosale2022soft}.\n\n\\subsection*{Biostreamer fabrication and flow imaging integration}\n\nWe aim to build a millimeter-scale, 3D muscle tissue to harvest contractile motions for the generation of oscillatory flows. All the while, the geometry of the tissue itself is selected and leveraged to modulate identifiable streaming fields, facilitate biofabrication, $\\mu$PIV imaging, and flow analysis.\n\nHere, skeletal muscle myoblasts are chosen for their ability to form 3D tissue and fuse into muscle fibers (myotubes) that align spontaneously along stress fields, thus providing directed contractions \\cite{neal2014formation}. This can be achieved by allowing the tissue to anchor on a pair of posts \\cite{aydin2019neuromuscular,aydin2020development}, providing axial tension thus longitudinal alignment and contraction, or by letting the tissue wrap around a single cylindrical pillar \\cite{uzel2014microfabrication}, generating tangential stresses and tangentially aligned fibers. These, upon contraction, cause the circumference of the tissue to shorten, resulting in overall radial actuation \\cite{li2019biohybrid,li2022adaptive}.\n\nThe latter approach is particularly suited to our investigation (\\cref{fig:1}a). Indeed, the cylindrical pillar minimizes uneven stress distributions, favoring regular fiber organization, tissue shape, and uniform radial contractions. This, in turn, is conducive of reproducible flow responses. Further, geometric axisymmetry implies the existence of a midplane where the flow becomes two-dimensional, simplifying flow imaging and analysis. Finally, a ring shape can be conveniently suspended in bulk liquid by simply sliding it onto a thin submerged rod (\\cref{fig:1}b).\n\nTo fabricate our muscle ring (\\cref{fig:1}a), a polydimethylsiloxane (PDMS) mold is first created by pouring liquid precursor into a machined negative mold and allowing it to polymerize. A mixture of C2C12 mouse skeletal muscle myoblasts, ECM, and type-I collagen is then cast into the PDMS mold, where the myoblast-laden gel compacts into rings \\cite{bell1979production}. They are subsequently transferred onto a hydrogel tube submerged in culture medium, to induce differentiation, during which contractile myotubes form and align. Because of tube compliance and muscle internal tension, the diameter of the torus shrinks of typically 1--2~mm, before stabilizing.\n\n\n\\begin{figure*}[ht!]\n\\centering\n\\includegraphics[width=\\textwidth]{images\/fig2.png}\n\\caption{Computational design of the biostreamer and autonomous streaming. \n(a) Modeling of the muscle ring and its radial contractions. The computational muscle ring and surrounding liquid environment are characterized, consistent with experiments, by the non-dimensional Stokes layer thickness $\\delta_{AC} \/ a = 0.11$, where the length scale $a$ is the depth of the muscle ring.\n(b-c) Numerical simulation of the streaming flow around the muscle ring. The visualization of the flow field is presented using two different methods: (b) Stokes stream function and (c) dynamical system representation.\n(d-f) Observed critical points from simulation and illustrations of their corresponding local flow patterns.\n(g) Expected flow structure at the midplane of the muscle ring from simulation. Critical points are marked by circles with colors corresponding to (d-f). Colour contours on the presented plane indicate regions of clockwise (blue) and counter-clockwise (orange) recirculating fluid. Regions near the critical points where potential particle manipulations can be achieved are encircled in bold lines. \n(h) Confocal fluorescent $z$-stack image of a section of a muscle ring showing the circumferentially aligned myotubes (green) and cell nuclei (blue). \n(i) Bright-field image showing the muscle ring hanging on the thin rod. The ring has measured inner diameter of 3.5~mm, outer diameter 4.7~mm, and depth 1.4~mm. Orange curves illustrate the radial muscle contractions.\n(j) Kymograph of the top and bottom edge of the muscle ring showing its periodic contractions over 2 seconds.\n(k) Plots of the top and bottom edge displacement due to muscle actuation over 4 seconds.\n(l) Time-averaged streamlines of the upper half of the muscle ring from experiment juxtaposed with streamlines of the lower half of the muscle ring from numerical simulation, showing the matching centers and recirculating regions. \n(m) Dominant first POD mode extracted using $x$ velocity components ($u$) from experiments (upper half) juxtaposed with those from simulations (lower half), demonstrating coherent regions from comparatively high correlation zones (flow locations of the same color indicate that local fluid particles move coherently in a correlated fashion; orange\/blue colors indicate positive and negative values; the change in sign indicates an opposite sense of motion). \nThe location used for obtaining the experimental flow spectrum is marked. \nSimilar results are observed when analyzing at other flow locations (Fig. S3b) and when considering the $y$ velocity component $v$ (Fig. S3d). \n(n) Frequency spectrum of the experimental flow ($u$), simulated flow, and muscle actuation. The normalized amplitude is presented on a logarithmic scale. \n}\n\\label{fig:2}\n\\end{figure*}\n\n\nFor flow measurement and analysis, a $\\mu$PIV system is realized (\\cref{fig:1}b). A rectangular aluminum enclosure is machine fabricated and bonded on a glass slide to hold culture medium laden with tracer particles of diameter 6--8~$\\mu$m. Besides serving as a container, the aluminum box also shields the tracers from external electric fields, minimizing drift caused by their slight negative charge. Two of the box opposing walls present cutouts at their centers to hold a thin aluminum rod onto which the muscle ring is slid. The box has inner dimensions of 1~cm$\\times$1~cm$\\times$1~cm. This choice is dictated by the need to minimize convection, which is observed to be significant (compared to expected streaming velocities) for larger volumes. A direct consequence of using a relatively small box is that flow boundary effects come into play. Nonetheless, we note that streaming is robust to moderate confinement, as demonstrated in periodic lattices both experimentally and numerically \\cite{bhosale2021multi}. We thus expect that an eventual streaming field generated by the muscle ring will be topologically unaffected by the presence of the walls, although we do anticipate a geometric compression of the main flow structures towards the center of the enclosure. This, in turn, may facilitate flow visualization. Indeed, by bringing characteristic flow features closer together, the area to be imaged is reduced, allowing improved resolution and $\\mu$PIV reconstruction accuracy.\n\n\nThe box and glass slide are mounted on the $x$-$y$ translation stage of an Olympus IX81 inverted microscope equipped with a $2\\times$ objective. The differentiated biostreamer is then transferred into the $\\mu$PIV system, and we adjust the stage such that the muscle ring is at the center of the field of view. We focus the microscope to the midplane of the toroidal tissue. A Hamamatsu digital camera connected to the microscope is utilized to capture images at 50 Hz, which are paired with a 5-frame step size and integrated with spatial cross correlation to produce velocity field data \\cite{hart1998elimination,keane1992theory}.\nThe biostreamer is thus allowed to contract spontaneously, or upon stimulation by an external optical fiber setup, generating an oscillatory flow in its surrounding. In the case of optical stimulation, tissue biofabrication entails the use of C2C12 cells with a mutated variant of the blue light-sensitive ion channel, Channelrhodopsin-2 (ChR2). Myotubes differentiated from these myoblasts selectively respond to precise light wavelengths, in our case 470~nm \\cite{raman2016optogenetic}. To reach the illumination intensity required for muscle contraction \\cite{sakar2012formation,bryson2014optical,klapoetke2014independent}, we connect the fiber to a laser diode source producing light pulses of intensity 1.6~mW\/mm$^2$. The specific combination of light wavelength and intensity prevents the $\\mu$PIV illumination system from interfering with the biostreamer activity.\n\n\n\n\n\n\n\n\\section*{Results}\n\n\\subsection*{Computational design of the biostreamer} Streaming literature has almost exclusively examined the use of bodies without holes (genus-0), unlike our muscle ring (genus-1). While in one recent instance \\cite{chan2021three} a rigid torus was computationally investigated for externally driven flow oscillations (perpendicular to the body axis of symmetry), radial self-contractions have not been considered so far. Further complicating the design specifics of our system, 3D skeletal muscle constructs have been demonstrated in a range of dimensions and actuation conditions. Viable muscle rings have been indeed reported for inner diameters ranging from 0.5~mm to 14~mm \\cite{okano1997tissue,gwyther2011engineered,li2019biohybrid,li2022adaptive,pagan2018simulation}, with wall thicknesses between 0.5~mm and 4~mm \\cite{gwyther2011engineered,pagan2018simulation}. Spontaneous contractions have been found to occur within the frequency range of 1~Hz to 3~Hz, and oscillation amplitudes observed to vary between 20~$\\mu$m and 500~$\\mu$m. \\cite{li2019biohybrid,li2022adaptive,pagan2018simulation}. Additionally, optogenetic C2C12 muscle tissues have been shown to produce consistent force responses to optical stimulation across the 1~Hz to 4~Hz range \\cite{raman2016optogenetic}.\n\n\n\nTo provide insight into the behavior of our coupled flow-muscle system, and in search of design solutions able to robustly generate streaming, we simulate different biostreamer instances within the above outlined, realizable parameter space. We model the muscle ring as a solid undergoing prescribed radial oscillations (while conserving volume, \\cref{fig:2}a), immersed in an incompressible viscous fluid of unbounded domain. The coupled system is numerically discretized and simulated using a remeshed vortex method coupled with Brinkmann penalization (Methods), an approach \\cite{gazzola2011simulations} demonstrated in a variety of settings \\cite{bhosale2021remeshed,gazzola2011simulations,gazzola2012c,gazzola2012flow,gazzola2014reinforcement,gazzola2016learning}, including streaming \\cite{parthasarathy2019streaming,bhosale_parthasarathy_gazzola_2020,chan2021three}.\n\n\nIn simulations, the investigated biostreamers are observed to produce qualitatively similar streaming flows, although quantitative details vary depending on muscle geometry and actuation properties. A representative instance is reported in \\cref{fig:2}b-g. On the left (\\cref{fig:2}b), isosurfaces of the time-averaged Stokes stream function (blue and orange representing clockwise and counter-clockwise flow rotations) highlight the presence of four recirculating annular regions (two on each side of the muscle ring). Two large regions are found to envelope the biostreamer, while a pair of smaller, nested ones approximately lie within the muscle itself. On the right (\\cref{fig:2}c) is the corresponding dynamical representation, whereby critical points (characterized by zero velocity) are extracted from the flow and classified based on their stability properties (through the Jacobian of the local velocity field).\n\n\n\nThe flow is found to be organized around four types of critical points: centers (pink -- \\cref{fig:2}d), saddles (yellow -- \\cref{fig:2}e) and attractive (blue) \/ repulsive (red) node-saddle-saddle (NSS -- \\cref{fig:2}f). In our axisymmetric case, both centers and saddles are 2D degenerate \\cite{strogatz2018nonlinear}, since the surrounding local flow has no out-of-plane (i.e. azimuthal) component. These degenerate points, when mapped to 3D space, form four continuous center-rings (pink) and four continuous saddle-rings (yellow), as illustrated in \\cref{fig:2}c. They collectively define the four annular recirculating regions described above.\n\nThe extracted flow skeleton underscores the structures that are most likely to be captured experimentally. A streamfunction slice through the midplane (\\cref{fig:2}g) mimics experimental visualization conditions and provides intuition. The locations at which the rings cross the midplane are identified by dots of corresponding colors, elucidating the role of centers and saddles in shaping the flow while highlighting their practical utility. Specifically, centers initiate recirculation and can be employed to attract\/retain particles \\cite{chong2013inertial,chong2016transport} and mix fluids, while saddles (and connecting streamlines) partition the flow, enabling particle separation and sorting \\cite{lutz2003microfluidics,thameem2016particle} or spatially controlled chemistry \\cite{lutz2003microfluidics,lutz2006characterizing}. Based on \\cref{fig:2}g, we anticipate that inner rings, being tucked within the muscle, will be difficult to image experimentally. Further, repulsive (red) and attractive (blue) saddles, that lie along the axis of symmetry in unperturbed simulations, may not be found on the experimentally imaged midplane, as a consequence of implementation imperfections and perturbations induced by the aluminum rod. Thus, the two external center-rings are likely the most robustly detectable structures. We then expect to observe, projected on the $\\mu$PIV imaging plane, four centers, located around the muscle ring and associated with recirculatory flows.\n\nOverall, we numerically find that muscle rings of approximately $\\sim$4~mm inner diameter and $\\sim$5~mm outer diameter may be ideal candidates for experimental realization. Indeed, at typical oscillation amplitudes ($\\sim 100$~$\\mu$m) and frequencies (1~Hz to 4~Hz), such rings are observed to generate streaming velocities ($\\sim 10~\\mu\\text{m\/s}$) significantly larger than background disturbances ($\\sim 1~\\mu\\text{m\/s}$). Further, key flow structures are predicted to closely surround the muscle, thus falling within a compact field of view, while remaining sufficiently separated for experimental detection. \n\n\n\n\n\n\n\n\n\\subsection*{Autonomous streaming}\nWe grow the biostreamer targeting the computationally identified specifications. A PDMS toroidal mold of 5~mm inner diameter, 12~mm outer diameter, and 3~mm depth is fabricated and the mixture of muscle cells and ECM is seeded (\\cref{fig:1}a). Mold dimensions are empirically determined to account for compaction, differentiation, and associated shrinking, leading to muscles of the approximately desired geometry ($\\sim$4~mm inner, $\\sim$5~mm outer diameters). Confocal fluorescent imaging also confirms myotubes circumferential alignment, for radial contraction (\\cref{fig:2}h).\n\nUpon transfer to the $\\mu$PIV system, the rings' specific dimensions and spontaneous contraction amplitudes\/frequencies are determined by tracking and averaging top and bottom edge displacements across 1,000 camera images acquired at 50~Hz. For the representative muscle of \\cref{fig:2}h-n, we measure a 3.5~mm inner diameter, 4.7~mm outer diameter, 1.4~mm tissue depth, 3.1~Hz dominant contraction frequency, and 72~$\\mu$m\/117~$\\mu$m average peak-to-peak top\/bottom amplitudes. Obtained amplitudes and frequencies are reflected in the kymographs of \\cref{fig:2}j, over 2 seconds, and in the time-varying edge displacements of \\cref{fig:2}k, over 4 seconds.\n\nTo facilitate comparison, we input these muscle-specific parameters into our simulations and analyze both experimentally and computationally obtained flows. We first consider experimental steady streamlines, depicted in \\cref{fig:2}l, top half. These are obtained by averaging recorded instantaneous velocities in the midplane, over 8 contraction cycles. This procedure removes (approximately) the flow's oscillatory components, revealing the underlying rectified, steady streaming field. As anticipated, we observe two distinct centers and corresponding flow recirculation regions. These features are also visible in the mirrored streaming field obtained from simulations, revealing close agreement (\\cref{fig:2}l, bottom half). We note the slight asymmetry that characterizes experiments, on account of ring geometry and actuation imperfections as well as due to the presence of the aluminum rod. These perturbations are also likely responsible for the missing NSS along the ring axis of symmetry, features that we previously identified as non-robustly observable. We further highlight how in experiments, centers are found closer to the biostreamer than in simulations. This is consistent with the intuition that wall effects (enclosure) would result in flow geometric compression. Finally, we note that the overall streaming response is found to be repeatable, with associated features seen across cyclic muscle contractions and samples (Fig. S1, S3).\n\nWe proceed with identifying additional hallmarks of streaming, through a combination of Proper Orthogonal Decomposition (POD) \\cite{sirovich1991analysis} and flow spectral analyses. POD considers space-time correlations in the velocity field and extracts energetic coherent structures lingering in the flow. We can employ this information to determine flow locations highly representative, in terms of their spectral fingerprint, of the underlying flow. We thus consider experimental and simulated midplane velocities over 20 seconds, extract the corresponding dominant POD mode (\\cref{fig:2}m, 90\\% of total energy, orange\/blue intensity correlates to local flow energy), and sample it at different coordinates (\\cref{fig:2}m and S3) at which spectral analysis of raw velocity data is performed.\n\nIn frequency domain, streaming flows present a well-known signature: a driving-frequency main peak capturing first order $\\mathcal{O}(\\epsilon)$ oscillatory effects, a zero-frequency peak (about ten times smaller) capturing second order $\\mathcal{O}(\\epsilon^2)$ streaming rectification, and a series of peaks of exponentially decaying magnitude at multiples of the driving frequency, corresponding to increasingly higher-order effects \\cite{holtsmark1954boundary}. \nOur data are found to quantitatively recapitulate this structure, as illustrated in \\cref{fig:2}n where we report the spectra of measured (black) and simulated (blue) velocities at the representative location marked as \\textcircled{$\\times$} in \\cref{fig:2}m. \nAdditionally, for reference, we plot the spectrum of the muscle top edge displacement (orange) of \\cref{fig:2}k. As can be noticed, the measured\/simulated flows indeed respond to the input muscle frequency (peak at 3.1~Hz) and produce rectification (peak at 0~Hz). The relative magnitude of these two peaks is also in line with theory, with rectification about one order of magnitude weaker, on account of being a second order effect.\n\nWe also note the presence of the characteristic peaks at multiples of the driving frequency. In simulations (blue bars), as expected from theory, the peak corresponding to the first frequency doubling ($\\sim$6~Hz) presents a magnitude comparable to the rectification peak, consistent with the fact that this is a second order effect too. Remaining frequency-multiples peaks are also detected from simulated flows (Fig. S2), although they are not visible in \\cref{fig:2}n given their negligible normalized amplitudes (that fall below the plotted lower bound), on account of their exponential decay \\cite{holtsmark1954boundary}.\n\nThe experimental flow response (black bars) does present a similar structure, whereby driving, rectification and cascading peaks are all found at the theoretically and computationally expected frequencies. However, the relative magnitude of the peaks at frequencies multiples of the driving one, is significantly stronger than predicted. This is not necessarily surprising for the following complementary reasons. First, the muscle is soft and can deform in response to the flow, unlike in theory where the body is considered to be rigid or in simulations where the ring's deformations are imposed. Disregarding body compliance leads to underestimating the strength of the streaming response, as recently demonstrated in \\cite{bhosale2022soft} whereby flow-induced elastic deformations are shown to be an additional source of streaming. Second, frequency multiples might originate from the tissue itself on account of its non-linear mechanical response to a driving oscillatory forcing (in this case cyclic contractions), through a mechanism mathematically similar to fluids \\cite{holtsmark1954boundary}. This effect, pertaining to the solid phase, might also couple with the flow, strengthening both responses. That a combination of these factors may be at play is suggested by the muscle spectrum (orange bars) of \\cref{fig:2}n, whereby deformations are found to precisely sync with the flow, presenting a set of highly structured harmonics that are unlikely to be the result of spontaneous, higher-frequency tissue contractions. While an explanation for the observed peak enhancement is provided, dissecting the exact mechanisms at play is beyond the scope of this work and not strictly necessary to confirm the presence of streaming or demonstrate the utility of our bio-hybrid platform.\n\nOverall, given the observed qualitative and quantitative agreement between experiments, simulations and theory, we conclude that our tissue-engineered biostreamer is indeed capable of generating streaming autonomously.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\begin{figure*}[t!]\n\\centering\n\\includegraphics[width=\\textwidth]{images\/fig3.png}\n\\caption{Optically controlled streaming. \n(a) Confocal fluorescent $z$-stack images of a section of a muscle ring showing the myotubes (MF 20, green) and the ChR2 marker (tdTomato, red) at the same locations. \n(b) Muscle edge displacement of the top half of the muscle ring over 3 seconds (pink) and corresponding light stimulation pattern made of pulses of 10~ms duration (blue) at different frequencies. (top) 2~Hz stimulation results in an average peak-to-peak amplitude of 81~$\\mu$m, (middle) 3~Hz stimulation and 59~$\\mu$m peak-to-peak amplitude, and (bottom) 4~Hz stimulation and 61~$\\mu$m peak-to-peak amplitude. (c) Frequency spectrum of the experimental flow ($u$), simulated flow, and muscle actuation at optical driving frequencies of (top) 2~Hz, (middle) 3~Hz, and (bottom) 4~Hz. The normalized amplitude is presented on a logarithmic scale. \n}\n\\label{fig:3}\n\\end{figure*}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\subsection*{Optically controlled streaming}\nNow, we demonstrate controllability of the native function of the muscle ring by exploiting muscle contractions in response to an external light stimulus. \nTo control the actuation frequency using optical stimulation, we fabricate our biostreamer following the same process of the autonomous streamer, however, we now utilize optogenetic C2C12 myoblasts.\nAfter fabrication, we use confocal fluorescence microscopy to confirm the expression of ChR2 after differentiation.\nIndeed, we observe the persistent expression of ChR2 within the tissue (\\cref{fig:3}a) and, therefore, expect the muscle ring to respond to an external light stimulation of wavelength 470~nm. \nAn optical fiber is fixed on the microscope stage and is used to uniformly illuminate the muscle ring within the enclosure. \nThe fiber delivers 10~ms pulses from the connected laser diode source at driving frequencies of 2, 3, and 4~Hz. \nThese driving frequencies are less than, equal to, and greater than the autonomous biostreamer actuation frequency ($\\sim 3.1$ Hz).\n\n\n\n\nUpon light stimulation, the representative muscle ring of \\cref{fig:3} is observed to contract robustly for all three driving frequencies, as quantified by tracking its edge displacement \\cref{fig:3}b.\nAt the lowest driving frequency (2~Hz), the muscle actuation response is characterized by rapid contractions ($\\sim 0.1$~s) followed by slower relaxations ($\\sim 0.4$~s) to its original configuration (\\cref{fig:3}b, top).\nAt higher driving frequencies (3 and 4~Hz), the muscle responds to the stimulus consistently, with contractions over $\\sim 0.1$ s (\\cref{fig:3}b, middle\/bottom).\nHowever, the relaxation time ($\\leq 0.2$~s) is no longer sufficient for returning to the original configuration and smaller edge displacement amplitudes are realized. \nFor the driving frequencies of 2, 3, and 4 Hz, the average peak-to-peak edge displacements are indeed measured to be $81 \\ \\mu$m, $59 \\ \\mu$m, and $61 \\ \\mu$m, respectively. \n\n\n\n\n\nFinally, using the same spectral analysis approach of \\cref{fig:2}, we characterize the oscillatory and rectified flow fields generated by the optically driven biostreamer, both in simulations and experiments.\nFor all driving frequencies, we find that the experimental flow (black) responds with hallmarks similar to the spontaneous case, as the dominant mode shifts to align with the optical driving frequency (\\cref{fig:3}c), underscoring the system controllability.\nFurther, and importantly, the zero-frequency peak, which corresponds to the rectified streaming flow, is observed for all three driving frequencies. \nIn good agreement with theory and simulations (blue), the magnitude of the zero-frequency peak is an order of magnitude smaller than the dominant mode. \nMoreover, the presence of the cascading peaks at multiples of the driving frequency are evident for all three cases. \nSpecifically, for a driving frequency of 2 Hz, peaks of decaying magnitude are observed at 4, 6, 8, and 10 Hz. \nSimilarly, for driving frequencies of 3 and 4 Hz, peaks of decaying magnitudes are observed at 6 and 9 Hz, and at 8 and 12 Hz, respectively. \nAs for the autonomous biostreamer, the magnitude of the cascading peaks is larger in experiments than simulations, for the same reasons previously described. \n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section*{Conclusion}\nIn this work, we present a bio-hybrid setup that allows us to investigate viscous streaming in biologically-powered, untethered, three-dimensional and millimeter-scale systems, approximating conditions experienced by a host of small aquatic creatures, long speculated (but not rigorously, experimentally confirmed) to autonomously rectify surrounding flows.\n\nBy combining tissue engineering, micro-particle image velocimetry, numerical simulations and flow analysis, we design and demonstrate a $\\sim$3.5 millimeter muscle ring that, once freely suspended in bulk liquid, generates viscous streaming, autonomously or in a remotely controlled fashion, via light stimulation. This experimentally verifies the speculated ability of millimeter-scale organism to produce and sustain streaming, while paving the way for potential applications in microfluidics and robotics. Further, this work sets the stage for exploring novel streaming flows and dynamics through design and actuation strategies afforded by tissue-engineering approaches \\cite{filippi2022will}, broadening the scope of bio-hybrid technology to both fundamental and applied fluid mechanics.\n\n\n\n\\matmethods{\n\n\\subsection*{PDMS Mold}\nPDMS base, Sylgard 184, and curing agent (Dow Corning) are mixed with 10:1 ratio by weight and degassed in a vacuum desiccator. \nThe mixture is then poured into the machine fabricated aluminum negative mold and cured at 60 $^{\\circ}$C overnight. \nThe PDMS molds are peeled off from the negative structures the next day and sterilized by autoclave. The PDMS mold has an inner diameter of 5~mm, an outer diameter of 12~mm, and an depth of 3~mm.\nPrior to seeding the cell-gel solution, PDMS molds are treated overnight with 1\\% (w\/v) pluronic F-127 (Sigma) dissolved in phosphate-buffered saline to reduce adhesion of cells and proteins.\n\n\\subsection*{Cell culture and muscle ring formation}\nC2C12 murine myoblasts and those that are transfected with pLenti2-EF1$\\alpha$-ChR2[H134R]-tdTomato-WPRE plasmid to express Channelrhodopsin-2 (ChR2) are seeded in cell culture flasks with growth medium consisted of 10\\% fetal bovine serum (Gibco), 1\\% L-glutamine (Gibco) and 1\\% penicillin-streptomycin (Lonza) in Dulbecco's modified eagle medium (Gibco) and grown to $70-80\\%$ confluency. \nFor muscle tissue formation, cells are collected and mixed with Matrigel basement membrane matrix and type-I collagen diluted by growth medium. \nThe cell-gel mixture is then injected into the PDMS mold and allowed to polymerize for 1 hour at 37$^{\\circ}$C, then the sample is inundated in growth medium and incubated for 2 days. \nAfter 2 days, they are transferred onto hydrogel tubes immersed in differentiation medium, which replaces the fetal bovine serum with horse serum (Gibco) in the growth medium, to induce the formation of myotubes.\nThe muscle rings are kept on the tubes for 10 days, with the differentiation medium refreshed every 2 days, before being used for the experiment. \n\n\n\\subsection*{Experiment setup}\nThe aluminum enclosure for the experiment is machined and glued to a microscope glass slide using Kwik-Sil silicone adhesive (World Precision Instrument). \nThe box also serves to shield electric field, minimizing tracking particles motion caused by their slight negative charge. \nBefore experiment, one muscle ring is transferred from the hydrogel tube to a thin aluminum rod of 1~mm diameter, which is then transferred to the middle of the aluminum enclosure filled with culture media mixed with tracking particles. \nThe aluminum enclosure apparatus is then secured on the microscope stage. \nA cover slip is added on top of the box to reduce convection, and the particles are allowed to settle for five minutes before recording. \nRecording is done using a digital complementary metal\u2013oxide semiconductor camera (Hamamatsu) with $2\\times$ objective focused to center plane of muscle based on sharpness of the muscle ring edge. \nEach recording is 20 seconds long with 50~Hz capturing frequency, 20~ms exposure time and $2 \\times 2$ binning. \nA control experiment, which only have culture media, is conducted with same temperature and carbon-dioxide control as streaming experiment to verify the particles motion without muscle ring is insignificant. \nOptical stimulation is carried out by an optical fiber of diameter 2.6~mm connected to a laser diode (Doric), fixed on the microscope stage next to the aluminum enclosure, producing light pulses at 2, 3 and 4~Hz. \nEach light pulse is 10~ms long with intensity 1.6~mW\/mm$^2$. \n\n\\subsection*{Simulation method and numerical implementation}\nWe briefly recap the governing equations and numerical solution technique. \nWe consider incompressible viscous flows in an unbounded domain $\\Sigma$. \nIn this fluid domain, immersed solid bodies perform simple harmonic oscillations. \nThe bodies are density matched and have support $\\Omega$ and boundary $\\partial\\Omega$, respectively. \nThe flow can then be described using the incompressible Navier-Stokes equations \\cref{eq:1}.\n\n\\begin{equation}\\label{eq:1}\n\\nabla \\cdot \\boldsymbol{u}=0; \\quad\\frac{\\partial u}{\\partial t}+(\\boldsymbol{u} \\cdot \\boldsymbol{\\nabla}) \\boldsymbol{u}=-\\frac{\\boldsymbol{\\nabla} P}{\\rho}+\\nu \\nabla^{2} u, \\quad x \\in \\Sigma \\backslash \\Omega\n\\end{equation}\n\nwhere $\\rho$, $P$, $\\boldsymbol{u}$ and $\\nu$ are the fluid density, pressure, velocity and kinematic viscosity, respectively. \nThe dynamics of the fluid-solid system is coupled via the no-slip boundary condition $\\boldsymbol{u}=\\boldsymbol{u}_{s}$, where $\\boldsymbol{u}_{s}$ is the solid body velocity.\nThe muscle ring contraction is prescribed via a time-varying core radius of a torus $R(t) = R_0 + A\\sin(\\omega t)$, where $R_0$ is the core radius in rest configuration, and $A$ and $\\omega$ are the contraction amplitude and frequency, respectively.\nIn order to conserve volume, we vary the tube radius of the torus accordingly through $r(t) = r_0\\sqrt{R_0 \/ R(t)}$, where $r_0$ is the tube radius in rest configuration.\nHence, the overall muscle contraction in our simulation is described as a torus with body velocity $dR \/ dt = A \\omega \\cos(\\omega t)$ accounting for time-varying core radius due to muscle contraction, and body deformation velocity $dr \/ dt = -0.5 A \\omega r_0 R(t)^{-1.5} \\sqrt{R_0} \\cos(\\omega t)$ accounting for tube radius adjustment to conserve volume.\nThe 3D system of equations is then solved using a velocity-vorticity formulation with a combination of remeshed vortex methods and Brinkmann penalization \\cite{gazzola2011simulations}, in the axisymmetric coordinate system. This method has been validated across a range of flow\u2013structure interaction problems, from flow past bluff bodies to biological swimming \\cite{gazzola2011simulations,gazzola2012flow,gazzola2012c,gazzola2014reinforcement,gazzola2016learning,bhosale2021remeshed,bhosale2022soft}. Recently, the accuracy of this method has been demonstrated in rectified flow contexts as well, capturing steady streaming responses from arbitrary rigid shapes in 2D and 3D \\cite{parthasarathy2019streaming,bhosale_parthasarathy_gazzola_2020,chan2021three,bhosale2021multi}. \n\n\\subsection*{Particle image velocimetry}\nVelocity field measurement of streaming is carried out by micro-particle image velocimetry ($\\mu$PIV). \nThe culture medium is seeded with polystyrene spheroids of diameter 6-8~$\\mu$m and specific gravity 1.05. \nThe aluminum enclosure minimizes particles motion caused by their slight negative charge. \nA control experiment, in which only the muscle ring is absent while all other conditions are maintained, is first conducted to verify the particle motion without muscle ring is insignificant. \nAfter the muscle ring is transferred into the system, the microscope is focused to center plane of muscle ring based on sharpness of the its edges. \n1000 images is then collected at a frequency of 50~Hz at 1~MP ($1024 \\times 1024$ pixels) resolution. \nThe images are then paired with 5 frames step size and then integrated with spatial cross correlation method \\cite{hart1998elimination,keane1992theory}. \nThe final integration window resulted in $10 \\times 10$ pixels with 50\\% overlap and a final grid spacing of $\\Delta X = \\Delta Y = 32.5 ~\\mu$m.\n\n\\subsection*{Proper orthogonal decomposition}\nDominant coherent structures are obtained with snapshot proper orthogonal decomposition (POD) following Sirovich \\cite{sirovich1987turbulence}. \nVelocity fluctuations $\\boldsymbol{u}^{\\prime}(\\boldsymbol{x}, t)$ are decomposed into a deterministic spatially-correlated part $\\phi^n(\\boldsymbol{x})$ and time-dependent coefficients $a^n(t)$ according to \\cref{eq:2}\n\\begin{equation}\\label{eq:2}\n\\boldsymbol{u}^{\\prime}(\\boldsymbol{x}, t)=\\sum_{n=1}^{N} a^{n}(t) \\boldsymbol{\\phi}^{n}(\\boldsymbol{x})\n\\end{equation}\nwhere $N\\approx 1000$ is the number of snapshot. \nAccording to corresponding ratio of the eigenvalues to summation of $N$ eigenvalues, i.e. $E_{n}=\\lambda_{n} \/ \\sum_{m=1}^{N} \\lambda_{m}$, each mode is sorted by its contribution to the turbulence kinetic energy. \n\n\\subsection*{Spectral analysis}\nFor the muscle ring, the displacement of the top and bottom edges is tracked using the image analysis software Tracker(physlets.org\/tracker). An FFT is then performed on the displacement data to produce the spectrum of muscle actuation. For the experimental flow, one point within the recirculation region is selected and FFT is performed on its time-dependent POD coefficients. \n\n\\subsection*{Muscle ring staining}\nThe muscle ring is rinsed with PBS and fixed in 4\\%v\/v of paraformaldehyde for 30 minutes. It is then washed three times for 5 minutes to permeabilize the tissue sample and incubated with 0.2\\%v\/v Triton X-100 (Sigma) diluted in PBS for 15 min. The muscle ring is then blocked and stored in Image-iT FX Signal Enhancer blocking solution (Invitrogen) at 4$^\\circ$C overnight.\nThe primary antibody, mouse antimyosin heavy chain (MF-20) is used to stain for myosin heavy chain with a 1:400 dilution ratio. The sample is incubated overnight at 4$^\\circ$C and then washed three times for 5 minutes before staining with secondary antibodies on the next day. The secondary antibody, AlexaFluor-488 anti-mouse (Invitrogen) is used to stain the MF-20 antibody overnight. The sample is incubated with DAPI (Invitrogen) for 15 minutes to stain nuclei at 4$^\\circ$C. After washing with PBS three times, the LSM 700 is used for the confocal fluorescent imaging.\n\n}\n\n\\showmatmethods{}\n\n\n\\acknow{The authors thank S. Hilgenfeldt for helpful discussions over the course of this work. The authors acknowledge support by the National Science Foundation under NSF CAREER Award \\#CBET\u20131846752 (MG) and the NSF Expedition 'Mind in Vitro' Award \\#IIS\u20132123781 (MG, TS).}\n\n\\showacknow{}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{section:introduction}\n\n\n\n\\begin{figure}\n\\vskip -0.01in\n\\centering\n\\includegraphics[width=7.5cm]{noise_v8.pdf}\n\\caption{Illustration of how interleaved connections contaminate the choice of neural network connections. The red, purple, and green links denote three candidate connections, among which the red one is the best choice (according to the individual evaluation). However, when interleaved connections (the dashed links) are added for the search sampling, the judgment of the candidates gradually becomes ridiculous. The bottom figures indicate the weight change of the three candidates throughout the search process (warm-up for 20 epochs), while $(a)$ is under the interleaving-free setting and $1$--$3$ interleaved connections are added for $(b)$--$(d)$. Specifically, the interleaved connection $(3,7)$ is added for $(b)$, the connections $(1,4)$, $(3,7)$ added for $(c)$, and $(1,4)$, $(2,6)$, $(3,7)$ added for $(d)$.}\n\\label{fig:introduction}\n\\end{figure}\n\nNeural architecture search (NAS) is a research field that aims to automatically design deep neural networks~\\cite{NASNet,AmoebaNet,metaQNN}. There are two important factors that define a NAS algorithm, namely, the search space that determines what kinds of architectures can appear, and the search strategy that explores the search space efficiently. Despite the rapid development of search algorithms which have become faster and more effective, the search space design is still in a preliminary status. In particular, for the most popular search spaces used in the community, either MobileNet-v3~\\cite{mobilenetv3} or DARTS~\\cite{liu2018darts}, the macro structure (\\emph{i.e.}, how the network blocks are connected) is not allowed to change. Such a conservative strategy is good for search stability (\\textit{e.g.}, one can guarantee to achieve good performance even with methods that are slightly above random search), but it reduces the flexibility of NAS, impeding the exploration of more complicated (and possibly more effective) neural architectures.\n\nThe goal of this paper is to break through the limitation of existing search spaces. For this purpose, we first note that the MobileNet-v3 and DARTS allow a cell to be connected to 1 and 2 precursors, respectively, resulting in relatively simple macro structures. In opposite, we propose a variant that each cell is connected to $L$ precursors ($L$ is 4, 6, or 8), and each connection can be either present or absent. We evaluate three differentiable NAS algorithms, namely DARTS~\\cite{liu2018darts}, PC-DARTS~\\cite{xu2020pcdarts}, and GOLD-NAS~\\cite{GoldNAS} in the designed $L$-chain search space, and all of them run into degraded results. We perform diagnosis in the failure cases and the devil turns out to be the so-called \\textbf{interleaved connections}, which refers to a pair of connections $(a,b)$ and $(c,d)$ that satisfies $aL)$, there are $2^{2L}-1$ input possibilities because each of the $2L$ operators can be on or off. As a result, there are $(2^2-1)\\times(2^4-1)\\times\\cdot\\cdot\\cdot\\times(2^{2(L-1)}-1)\\times(2^{2L}-1)^{N-L}$ combinations if there are $N$ nodes in one stage. There are 3 stages to be searched in our space, Fig.~\\ref{fig:space}, with the node number of 18, 20 and 18 respectively. Therefore, if $L$ is 4, 6 and 8, there are about $1.8\\times10^{116}$, $7.5\\times10^{163}$ and $1.6\\times10^{204}$ possible architectures respectively. The complexity comparison of popular spaces are shown in Tab.~\\ref{tab.space_complexity}. \n\n\n\n\n\n\n\\begin{figure*}[t]\n\\begin{center}\n\\includegraphics[width=.90\\linewidth]{pipeline_V6.pdf}\n\\end{center}\n\\caption{The flowchart of IF-NAS. The key is to avoid interleaved connections in \\textit{any} time. For this purpose, we repeat a loop with a length of $L$, and each time part of the connections remain active and all others are removed. The layers of the same color are always activated and inactivated together. This figure is best viewed in color.}\n\\label{fig:pipeline}\n\\end{figure*}\n\n\\subsection{Failure Cases and Interleaved Connections}\n\nWe first evaluate DARTS~\\cite{liu2018darts} and GOLD-NAS~\\cite{GoldNAS} in the search space of $\\mathbb{S}^{(4)}$. \nOn the ImageNet-1k dataset, trained for 250 epochs for each network, DARTS and GOLD-NAS report top-1 accuracy of $74.7\\%$ and $75.3\\%$, respectively, and the FLOPs of both networks are close to $600\\mathrm{M}$ (\\textit{i.e.}, the mobile setting), Tab.~\\ref{tab.main_results}. \nIn comparison, the simple chain-styled architecture (each node is only connected to its direct precursor) achieves $74.9\\%$ with merely $520\\mathrm{M}$ FLOPs.\nMore interestingly, if we only preserve one input for each node, DARTS and GOLD-NAS report completely failed results of $70.2\\%$ and $71.2\\%$ (trained for 100 epochs), which are even much worse than preserve one input randomly, Tab.~\\ref{tab.ablation_keep1}.\n\nWe investigate the searched architectures, and find that both DARTS and GOLD-NAS tend to find long-distance connections. This decreases the depth of the searched architectures, however, empirically, deeper networks often lead to better performance.\nThis drives us to rethink the reason that the algorithm gets confused. For this purpose, we randomly choose an intermediate layer from the super-network and the algorithm needs to determine the preference among its precursors.\n\nWe start from the simplest situation that all network connections are frozen (the architectural weights are fixed) besides the candidate connections. We show the trend of three (pre-$1$, pre-$3$, pre-$6$) connections in Figure~\\ref{fig:introduction}. One can observe that the pre-$1$ connection overwhelms other two connections, aligning with our expectation that a deeper network performs better. However, when we insert one connection that lies between these candidates, we observe that the advantage of the pre-$1$ connection largely shrinks, and a long training procedure is required for it to take the lead. The situation continues deteriorating if we insert more connections into this region. When two or more connections are added, the algorithm cannot guarantee to promote the pre-$1$ connection and, sometimes, the ranking of the three connections is totally reversed.\n\nThe above results inspire us that two connections are easily interfered by each other if the covering regions (\\textit{i.e.}, the interval between both ends) overlap. Hereafter, we name such pair of connections \\textbf{interleaved connections}. Mathematically, denote a connection that links the a-th and b-th nodes as $(a, b)$, where $b-a\\ge 2$. Overlook the candidate connections of the activated nodes, two connections, $(a,b)$ and $(a',b')$, interleave if and only if there exists at least one integer $d$ that satisfies $ak_1^{\\text{fp}8}$ while stable 1JO2TP (purple branch) are present between $k_1^{\\text{fp}11}