Carbon Chain Molecules: Production and Spectroscopic Detection Inaugural-Dissertation zur Erlangung des Doktorgrades der Mathematisch-Naturwissenschaftlichen Fakultät der Universität zu Köln vorgelegt von Guido W. Fuchs aus Polch UNIVERSITÄT ZU KÖLN Köln 2002 Berichterstatter: Prof. Dr. G. Winnewisser Prof. Dr. J. Jolie Tag der mündlichen Prüfung: 18.2.2003 Abstract In this work the production and detection of carbon chain molecules in laboratory and interstellar space are presented. The work is divided into three parts. Part I, the production of reactive molecules. The availability of efficient molecular sources are of great importance for absorption and emission experiments. Hence, their characterization and optimization is indispensable for the success of these kinds of experiments. Molecular sources can be very specialized concerning the species produced. The excimer laser ablation source used in Cologne is highly efficient in the production of pure carbon molecules, i.e. carbon clusters. However, the carbon cluster yield in the range from C10 to C60 is still not satisfactory. For an improvement of the production rates new methods have to be tested. In the course of this thesis a excimer laser ablation is compared with Nd:YAG laser ablation. For that purpose a quadrupole mass spectrometer has been used to characterize the Nd:YAG laser ablation source. Investigations on a slit nozzle discharge source have been performed. This type of molecule source is able to produce pure carbon clusters but was originately developed for the production of hydro-carbon molecules. Both kinds of molecule sources, i.e. laser ablation as well as slit nozzle discharge sources, produce a plasma which causes significant problems when recording mass spectra. Therefore, a mass spectrometer specially designed for plasma applications in combination with the discharge slit nozzle was tested in Lichtenstein/Balzer. Cations as well as anions could be detected but no signal of discharge related neutrals were found. In addition to the mass spectroscopic studies also infrared (IR) absorption experiments have been performed. In the course of this thesis the Cologne carbon cluster experiment has been rebuild. In particular, the preceding IR diode laser spectrometer has been replaced by a new one which has been largely improved by using a liquid nitrogen dewar for the laser diode, new detectors for the 2000 cm−1 frequency region, stable optical setup and the development of new data acquisition and calibration software. First measurements are presented. Part II, measurements of Cn N radicals. At the Harvard Laboratory Astrochemistry Group measurements have been performed on mono-substituted C3 N isotopomers (cyanoethynyl) in a supersonic molecular beam using Fourier transform microwave spectroscopy. A detailed spectroscopic characterization of 13 CCCN, C13 CCN, CC13 CN and CCC15 N including their hyperfine spectra is given in this work. The rotational and leading centrifugal distortion constants were determined to high accuracy by using microwave data between 9.5 - 38.4 GHz and previously measured millimeter data. The Fermi contact b(14 N), dipole-dipole c(14 N), and the nitrogen quadrupole hyperfine coupling constants for 13 CCCN, C13 CCN, and CC13 CN have been determined and the previously published b(13 C) and c(13 C) values were stated more precisely [McCarthy et al., J.Chem.Phys. 103, 7820 (1995)]. The magnetic hyperfine coupling constants of the presented 13 C isotopic species of C3 N differ from those of the isoelectronic chain C4 H, but are fairly close to those of the isoelectronic C2 H, indicating a rather pure 2 Σ electronic ground state. The CCC15 N b and c magnetic hyperfine constants follow the expected ii values derived from the 14 N species. In addition two new cyano radicals, linear C4 N and C6 N were analyzed. C4 N, C6 N are linear chains with 2 Π electronic ground states and both have resolvable hyperfine structure and Λ-type doubling. At least four transitions in the lowest-energy fine structure component Ω =1/2 were measured between 7 and 22 GHz and, at most, 9 spectroscopic constants were required to reproduce their spectra to a few parts in 107 . Although the strongest lines of C6 N are more than five times less intense than those of C5 N, owing to large differences in the ground state dipole moments, both new chains are more abundant than C5 N. Searches for C7 N have so far been unsuccessful. The absence of lines at the predicted frequencies requires that the product of the dipole moment times the abundance (µ · Na ) to be more than 60 times smaller for C7 N than for C5 N, suggesting that the ground state of C7 N may be 2 Π, for which the dipole moment is calculated to be small. Part III, Cn N radicals in the interstellar space. Astrophysical investigations of C3 N isotopomers are compared with laboratory data presented in this work. Possible misassignments of 13 CCCN lines towards IRC+10216 are investigated. Finally, a search for C2 N has been performed towards the envelope of the late-type star IRC+10216 using the IRAM 30m telescope at Pico Veleta, Spain. Three lines in the frequency bands at 154, 224 and 248 GHz have been detected at transition frequencies of C2 N and preliminary assigned to C2 N as a carrier. Thus, a rotational temperature as well as a column density of C2 N could be estimated. The results are in agreement with estimations deduced by previous observations (Guèlin & Cernicharo [74] ). Further astronomical measurements are necessary to confirm this tentative detection. Kurzzusammenfassung In der vorliegenden Arbeit werden die Produktion und die Messung von Radikalen im Labor sowie im Weltraum an ausgesuchten Beispielen vorgestellt. Die Arbeit ist in drei Teile gegliedert. Teil 1 befaßt sich mit der Charakterisierung von Molekülquellen. Die in Köln verwendete Excimer-Laserablationsquelle ist hoch effizient in der Erzeugung von reinen Kohlenstoffmolekülen, sog. Kohlenstoff Clustern. Zunächst wird eine ExcimerLaserablation mit einer Nd:YAG- Laserablation verglichen. Dabei wurde ein QuadrupolMassenspektrometer zur Charakterisierung der Nd:YAG-Ablationquelle eingesetzt. Desweiteren wurde eine Schlitzdüsen-Entladungsquelle untersucht die neben der Produktion von Kohlenwasserstoffen und anderer kohlenstoff-basierter Moleküle auch reine Kohlenstoffcluster erzeugen kann. In beiden Molekülquellenarten entsteht ein Plasma, daß zu erheblichen Schwierigkeiten bei der Aufnahme von Massenspektren führt. Ein speziell für Plasmen vorgesehenes Massenspektrometer wurde in Lichtenstein/Balzer mit Hilfe der Schlitzdüsen-Entladungsquelle getestet. Erste Ergebnisse werden vorgestellt. Zusätzlich wurde am Kölner Kohlenstoff Cluster Experiment das vorhandene IR-Dioden Spektrometer erneuert. Wesentliche Verbesserungen wurden erreicht durch den Einsatz eines Flüssigstickstoff-Dewars für die Kühlung der Laserdioden, nachweisempfindlichere Detektoren für den Frequenzbereich um 2000 cm−1 , einen stabileren optischen Aufbau, iii sowie die Entwicklung neuer Meß -und Kalibrationssoftware. Erste Meßungen werden vorgestellt. In Teil 2 dieser Arbeit werden Messungen an einfach-substituierten C3 N Isotopomeren sowie Untersuchungen an C4 N und C6 N vorgestellt. Die Messungen an 13 CCCN, C13 CCN, CC13 CN und CCC15 N führten zur detailierten spektroskopischen Charakterisierung der Radikale und wurden an einem Fourier Transform Mikrowellen Spektrometer der Harvard Laboratory Astrochemistry Group vorgenommen. Die linearen, mit 13 C und 15 N substituierten C3 N Moleküle wurden mittels einer elektrischen Entladungsquelle mit anschließender adiabatischen Expansion hergestellt. Mit den gemessenen Mikrowellendaten zwischen 9.5 und 38.4 GHz und den zuvor bekannten Millimeterwellen-Daten konnten die Rotations- sowie die führenden Zentrifugalverzerrungsterme sehr genau ermittelt, die Fermikontakt- sowie die Dipol-Dipol Wechselwirkung der 13 C-Isotope präzisiert und die magnetische Wechselwirkung der 14 N bzw. 15 N-Isotope erstmals ermittelt werden. Zusätzlich wurden zwei neue Cyan-Radikale, lineares C4 N und C6 N, untersucht. Basierend auf Messngen der Ω=1/2 Zustände zwischen 7 und 22 GHz [121], wurden die Molekülparameter der sich im 2 Π elektronischen Grundzustand befindenden Radikale ermittelt. Beide Spezies zeigen eine Hyperfeinstrukturaufspaltung und Λ-Verdopplung. Die in dieser Arbeit bestimmten neun Molekülparameter je Radikal ermöglichen eine Reproduktion der Spektren bis auf wenige kHz Genauigkeit. In Teil 3 werden astrophysikalische Untersuchungen an linearen C3 N Isotopomeren mit den in dieser Arbeit gewonnenen Labordaten verglichen. Eigene Arbeiten umfassen die Suche nach C2 N in der Sternenhülle von IRC+10216 mit Hilfe des IRAM 30m Teleskops am Pico Veleta, Spanien. Es wurden drei Linien in den Frequenzbändern um 154, 224 und 248 GHz beobachtet, die mit Rotationsübergängen von C2 N übereinstimmen und eine vorläufige Zuordnung dieser Linien zu C2 N erlauben. Weitere astrophysikalische Messungen sind jedoch notwendig um eine eindeutige Detektion von C2 N in IRC+10216 sicherzustellen. iv ”Jedermann sieht die Grenzen seiner eigenen Vision als die Grenzen der Welt an.” by Arthur Schopenhauer (1788 - 1860) For my parents Agnes and Werner Fuchs Contents Abstract / Kurzzusammenfassung i Zusammenfassung ix 1 Introduction 1 I 7 Characterization of Radical Sources 2 Molecule Source Characterization 2.1 Ablation Technique . . . . . . . 2.1.1 Excimer Laser Ablation 2.1.2 Nd:YAG Laser Ablation 2.2 Discharge Plasma . . . . . . . . 2.3 Conclusions and Prospects . . . 3 The 3.1 3.2 3.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 13 14 17 25 30 Cologne Carbon Cluster Experiment: The New Setup 33 The Cologne Carbon Cluster experiment . . . . . . . . . . . . . . . . . . 34 The New Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Conclusions and Prospects . . . . . . . . . . . . . . . . . . . . . . . . . . 37 II C3 N Isotopomers, C4 N, and C6 N 39 4 Experimental Setup 4.1 The Production of Cn N Radicals 4.1.1 The Precursor Gases . . . 4.1.2 The Discharge Nozzle . . . 4.2 Adiabatic Expansion . . . . . . . 4.3 The Fourier Transform Microwave 41 44 44 47 52 63 5 Linear Cn N, Cyanide Radicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spectrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 vi Contents 6 Theoretical Considerations 6.1 Pure Rotation of Linear Molecules . . . . . . . . . . . 6.1.1 Selection Rules . . . . . . . . . . . . . . . . . 6.2 Fine Structure . . . . . . . . . . . . . . . . . . . . . . 6.2.1 Hund’s Coupling Cases a) and b) . . . . . . . 6.2.2 Λ-type Doubling, and l-type Doubling . . . . 6.3 Hyperfine Structure . . . . . . . . . . . . . . . . . . . 6.3.1 Magnetic Hyperfine Structure . . . . . . . . . 6.3.2 The Electric Quadrupole Interaction . . . . . 6.4 Matrix Representation of the Hamiltonian . . . . . . 6.4.1 The Matrix Representation of the 2 Π-Radicals 7 Measurements and Analysis 7.1 The C3 N Mono-Substituted Isotopomers 7.1.1 CCC15 N . . . . . . . . . . . . . . 7.1.2 13 CCCN, C13 CCN and CC13 CN . 7.2 C4 N and C6 N . . . . . . . . . . . . . . . 7.3 The Search for C7 N . . . . . . . . . . . . 7.4 Conclusions and Prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 79 81 82 83 86 87 88 91 93 93 . . . . . . 97 98 98 103 112 123 123 III Linear Cn N Chains in Space 8 Cn N Chains in Space 8.1 The Search for Interstellar C2 N . 8.1.1 Observation . . . . . . . . 8.1.2 Data Analysis . . . . . . . 8.1.3 Discussion . . . . . . . . . 8.1.4 Conclusions and Prospects 125 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 134 134 137 140 141 IV Appendix 143 A Linear Cn H 145 B The HQ Matrix Elements 147 C Molecular Constants of C13 CCN and CC13 CN 151 D Tables: Interstellar C3 N ,C5 N, and C3 N Isotopomers 153 Bibliography 157 List of Figures 173 Contents vii List of Tables 175 Acknowledgments 177 Beglaubigung 179 Publication List 179 Curriculum Vitae 181 viii Contents Zusammenfassung Interstellares Gas, sowie von kalten Sternen ausgestoßenes Gas, zeigt eine große Vielfalt an beobachteten Molekülen, in der Kohlenstoff als viert häufigstes Element im Universum eine zentrale Rolle spielt. In der vorliegenden Arbeit werden die Produktion und die Messung von Radikalen im Labor sowie im Weltraum an ausgesuchten Beispielen vorgestellt. Insbesondere wird die Produktion von kohlenstoffhaltigen Radikalen wie reine Kohlenstoffcluster Cn oder Moleküle der Form Cn N erläutert. Die Anwendung der hier vorgestellten Techniken kann aber auch zur Produktion anderer Moleküle eingesetzt werden, z.B. von Sin Cm , NCn N, HCn H, HCn N, etc. . Effiziente Molekülquellen sind von zentraler Bedeutung für die Entdeckung und Strukturbestimmung neuer Moleküle mit Hilfe der Emissions- und Absorptionsspektroskopie. Die Charakterisierung und Optimierung von Molekülquellen ist daher wichtig in Hinblick auf zukünftige Erfolge auf diesem Gebiet der Forschung. Die in Köln verwendete Excimer-Laserablationsquelle ist hoch effizient in der Erzeugung von reinen Kohlenstoffmolekülen, sog. Kohlenstoff Clustern, und ermöglichte die Entdeckung von linearem C8 und C10 [61, 14]. Die Detektion von größeren Clustern erscheint jedoch zunehmend schwieriger, sodaß einer Erschließung neuer Produktionstechniken eine wachsende Bedeutung zukommt. In dieser Arbeit wurde zunächst eine Excimer-Laserablation mit einer Nd:YAG- Laserablation verglichen. Der seperate Aufbau einer Testapparatur erlaubte den Einsatz eines Quadrupol-Massenspektrometers zur Charakterisierung der Nd:YAGAblationquelle. Desweiteren wurde eine Schlitzdüsen-Entladungsquelle untersucht die neben der Produktion von Kohlenwasserstoffen und anderer kohlenstoff-basierter Molekülen auch reine Kohlenstoffcluster erzeugen kann. Bei beiden Quellen, d.h. bei Entladungs- sowie bei Ablationsquelle, entsteht ein Plasma, das zu erheblichen Schwierigkeiten bei der Aufnahme von Massenspektren führt. Diese lassen sich jedoch durch Verwendung geeigneter Energiefilter beheben, wie Testmessungen an einem Plasmamonitor der Firma Inficon AG gezeigt haben. In dem Entladungsplasma konnten dann Kationen sowie Anionen nachgewiesen werden. Neutralteilchen sind allerdings wesentlich schwieriger nachzuweisen. Neben der Optimierung der Quellen ist eine gesteigerte Nachweisempfindlichkeit der verwendeten Spektrometer essentiell. Wesentliche Verbesserungen des vorhandenen IRDioden Spektrometers wurden durch den Einsatz eines Flüssigstickstoff-Dewars zur Kühlung der Laserdioden erreicht. Die Verwendung von InSb-Detektoren statt der bisher verwendeten HgCaTe Detektoren führt im Frequenzbereich um 2000 cm−1 ebenfalls zu einem Gewinn im Signal-Rausch-Verhältnis. Ein stabilerer optischer Aufbau, sowie die Entwicklung neuer Meß -und Kalibrationssoftware führten zu einer Verbesserung der x Zusammenfassung Systemstabilität sowie zur präzisen Frequenzzuordnung der Mess-Signale. Erste Messungen belegen dies auf eindrucksvolle Weise. Desweiteren werden spektroskopische Untersuchungen an Kohlenstoffkettenmolekülen mit einer Nitrilgruppe Cn N im cm-Wellenlängenbereich vorgestellt. Im einzelnen sind dies vier C3 N Isotopomere sowie die Kettenmoleküle C4 N und C6 N. Die Messungen an 13 CCCN, C13 CCN, CC13 CN und CCC15 N führten zur detailierten spektroskopischen Charakterisierung der Radikale und wurden an einem Fourier Transform Mikrowellen Spektrometer der Harvard Laboratory Astrochemistry Group vorgenommen. Die linearen, mit 13 C und 15 N substituierten C3 N Moleküle wurden mittels einer elektrischen Entladungsquelle mit anschließender adiabatischen Expansion hergestellt. Mit den gemessenen Mikrowellendaten zwischen 9.5 und 38.4 GHz und den zuvor bekannten Millimeterwellen-Daten konnten die Rotations- sowie die führenden Zentrifugalverzerrungsterme sehr genau ermittelt, die Fermikontakt- sowie die Dipol-Dipol Wechselwirkung der 13 C-Isotope präzisiert und die magnetische Wechselwirkung der 14 N bzw. 15 N-Isotope erstmals ermittelt werden. Die magnetischen Kopplungskonstanten der 13 C enthaltenden C3 N Radikale unterscheiden sich von denen der isoelektronischen C4 H Ketten, liegen aber nahe an denen von C2 H bekannten Werten und lassen somit auf einen fast reinen 2 Σ Grundzustand schließen. Die CCC15 N magnetischen Hyperfeinkonstanten folgen den von den 14 N-Radikalen theoretisch abgeleiteten Werten. Zusätzlich wurden zwei neue Cyan-Radikale, lineares C4 N und C6 N, untersucht. Basierend auf Messungen an jeweils vier Rotationsübergängen im unteren Ω=1/2 Zustand zwischen 7 und 22 GHz [121], wurden die Molekülparameter der sich im 2 Π elektronischen Grundzustand befindenden Radikale ermittelt. Beide Spezies zeigen eine Hyperfeinstrukturaufspaltung und Λ-Verdopplung. Die in dieser Arbeit bestimmten 9 Molekülparameter je Radikal ermöglichen eine Reproduktion der Spektren bis auf eine Genauigkeit von ca. 107 . Obwohl die stärksten Linien von C6 N etwa 5 mal schwächer sind als die entsprechenden C5 N Linien, was auf einen großen Unterschied im Grundzustandsdipolmoment zurückzuführen ist, liegen beide neuen Kettenmoleküle in einer größeren Häufigkeit als C5 N vor. Es wurde im Verlaufe dieser Arbeit auch versucht C7 N spektroskopisch nachzuweisen. Trotz guter Vorhersagen konnte jedoch keine der in den beobachteten Spektren enthaltenen Linien auf C7 N zurückgeführt werden. Das Fehlen der erwarteten Linien setzt vorraus, daß das Produkt aus Dipolmoment und Häufigkeit (µ · Na ) mehr als 60 mal kleiner für C7 N als für C5 N ist, sodaß C7 N wahrscheinlich nicht im 2 Σ sondern im 2 Π Grundzustand vorliegt. Berechnungen ergeben, daß das zum 2 Π Grundzustand gehörige Dipolmoment sehr klein ist. Cn N Radikale sind auch im Weltraum schon nachgewiesen. In Teil 3 werden astrophysikalische Untersuchungen an linearen C3 N Isotopomeren mit den in dieser Arbeit gewonnenen Labordaten verglichen. Eigene Arbeiten umfassen die Suche nach C2 N in der Sternhülle von IRC+10216 mit Hilfe des IRAM 30m Teleskops am Pico Veleta, Spanien. Es wurden drei Linien beobachtet, die mit Rotationsübergängen von C2 N übereinstimmen und eine vorläufige Zuordnung dieser Linien zu C2 N erlauben. Die Säulendichte konnten abgeschätzt werden und steht in Einklang mit theoretischen Vor- xi hersagen von Millar & Herbst [132]. Weitere astrophysikalische Messungen sind jedoch notwendig um eine eindeutige Detektion von C2 N in IRC+10216 sicherzustellen. xii Zusammenfassung 1 Introduction “ I ask you to look both ways. For the road to a knowledge of the stars leads through the atom; and important knowledge of the atom has been reached through the stars. ” Sir Arthur Eddington (1882 - 1944), Stars and Atoms Interstellar gas and gas ejected into space by cool stars are known to contain a rich collection of molecules [123]. After Hydrogen, Helium and Oxygen, Carbon, the fourth abundant element in space with its plurality of possible bondings unclose a whole zoo of different carbon bearing molecules, ions and radicals. An important feature of the carbon atom is that it easily builds long carbon chain molecules. Many of the organic compounds are familiar to the terrestrial chemist and can be found in a standard chemical stockroom like formaldehyde, ethanol, and methyl formate but a large part of the known interstellar molecules, as seen in Tab. 1.1, are entirely new species. A closer look on these species reveals that while two-thirds of the diatomic molecules and one-third of the triatomics are inorganic there exists no inorganic molecule detected in space with more than 5 atoms indicating that future discoveries of large molecules are likely to exclusively comprise organic compounds.The particularity of the carbon chains found in space is that most of them are highly unsaturated and therefore represent a form of matter which is explosively unstable through polymerization even at moderate densities and difficult to study on earth. It is for this reason that laboratory detection lagged behind the astronomical discoveries of new radicals for a long time. There are two different types of carbon chains: those which carry a permanent electric dipole moment and those that are nonpolar. Most of the molecules found in space are polar compounds which can be detected with radio- or microwave techniques. Today, modern laboratory spectrometers like FTMW spectrometers or advanced mm-wave to THz-spectrometers can achieve a high sensitivity and frequency accuracy of 2 kHz in the 10 GHz region up to 50 kHz in the 1 THz region. Application of these powerful spectrometers on supersonic molecular beams in the mid 1990’s resulted in an avalanche of new detected molecules most of which are candidates of future astronomical discovery like long carbon chains, chains attached to rings, silicon-carbon rings, and protonated molecular ions. In the gas 2 1 Introduction phase, rotational transitions and low lying bending vibrations can be observed and yield in an unambiguous identification of molecules. Large telescopes with a high signal-tonoise ratio can achieve frequency accuracies of one part in 107 [179]. This connection between laboratory spectroscopy and radioastronomy yielded already results in the early 1960’s with the detection of the hydroxyl radical [192] in ’63, followed by ammonia [39] and water [40] 5 years later. In principal an unambiguous identification of molecules in space is also possible in the optical or UV frequency region like it happened for CH, CH+ and CN in the early 1940’s [195]. The majority of discoveries, however, have been made in the frequency region between 100 and 300 GHz (1-3 mm region). Nonpolar species which are equally important for astrochemistry than polar molecules have no pure rotational spectrum and can therefore not be detected by radio astronomy. This shortcoming has been partially overcome by the onset of infrared astronomy. Molecules such as CO2 , C2 H2 , CH3 , C3 or even C5 have been identified in this frequency region by ro-vibrational absorption or by emission from hot gas as for example in the case of H2 [50]. Beside asymmetric stretching modes in the mid-infrared region many carbon chain molecules exhibit infrared active, low-energy bending vibrations. All pure carbon chain molecules Cn with n > 2 are expected to display bending vibrations in the range 30 150 cm−1 , i.e. in the far-infrared region. C3 for example has already been detected in the interstellar space by KAO (Kuiper Airborne Observatory) near 2 THz [59]. Further spectroscopic data in these frequency region will be essential for future missions, such as SOFIA (Stratospheric Observatory For Infrared Astronomy) and Herschel, a 3 m telescope space mission. Because of the success in detecting more than 120 molecules so far and the prospect of finding many more, the interest in carbon bearing molecules is stronger than ever. Considering a spiral galaxy like our Milky Way, the interstellar medium consists about 75% of hydrogen and 24% of helium which leaves only 1% of the total interstellar mass for all other chemical elements [179]. ”It is one of the paradoxes of interstellar chemistry that unsaturated carbon should be so conspicuous in a regime where hydrogen is the dominant chemically active element by more than three orders of magnitude” ([123], p.178). A hydrogen-helium chemistry by its own is very poor, however, 80% of the remaining gas is consisting of C, N, and O which have a reach chemistry. Following this argument it comes as no surprise that molecules containing C, H and N like HCn N, Cn H or Cn N play a key role in the chemistry of molecular clouds or circumstellar envelopes, as can be seen in Tab. 1.1. 73% of the detected interstellar molecules contain carbon, 66% hydrogen, 34% nitrogen and 28% oxygen. It is a remarkable fact that in certain sources in space a large number of reactive organic compounds often exist in comparable abundances with stable compounds of the same size. Although big progress has been made in theoretical astrochemistry, this science is still heavily dependent on pure fact gathering. New chemical models like those of Millar & Herbst [131] or Doty et al. [48] include as much as 407 molecules, ions and radicals connected by 3851 reactions [132], and most of these species are still not detected in space. It can be seen as disadvantage of modern chemical network models, though, that they de- 3 pend on such large numbers of parameters and reaction coefficients which are not always known to high precision and which in turn allow for a certain flexibility in predictions. By new astronomical measurements, these parameters can be refined and the models can be tested. Radicals like Cn or Cn N which play an important role in the production or depletion of cyanopolyynes are a sensitive probe for such models. To make spectroscopic work contribute optimally to astrophysics, it is advantageous to examine molecules of small or medium size which contain C, H, or N. Therefore in this thesis radicals like linear isotopic C3 N, C4 N and C6 N, have been investigated and first measurements on C7 N and C8 N have been started but did not result in any detection so far. The main task of laboratory spectroscopy concerning its contribution to astrophysics is to make accurate frequency predictions available for species yet undetected. Under this considerations a spectroscopic characterization is said to be complete when it provides either measured transitions to high precision or calculated transitions from the derived spectroscopic constants to an accuracy similar to measurements. For open shell molecules as presented in this work not only the rotational and leading centrifugal distortion constants have to be determined but also several constants which characterize the hyperfine structure as well as the Λ-doubling caused by the unpaired electron. However, the first detection of linear C3 N in the gas phase has not been done in laboratory but with a radio telescope by Guélin and Thaddeus [78, 79] in 1977. Today, the study of the HC3 N-C3 N pair is of great importance to test and refine photochemical models e.g. of carbon-rich stars like IRC+10◦ 216 [38]. C5 N has first been measured in laboratory by Kasai et al. [99] and then in the dark cloud TMC-1 by Guélin et al. [77]. A detailed study of the electronic structure of C3 N is also worth undertaking because this radical is isoelectronic with C4 H, a well-studied molecule in which an extremely low-lying A2 Π electronic state strongly interacts with the X 2 Σ+ ground state [121]. Owing to large zero-order mixing between the two states, the 13 C hyperfine coupling constants of C4 H at each substituted position along the chain [37] differ significantly from those of the closely related chain CCH [120]. Determination of the analogous constants for isotopic C3 N should provide an unambiguous comparison of the electronic structure and chemical bonding for these two isoelectronic radicals. Even-membered Cn N chain radicals may play important roles in interstellar chemistry and soot formation [48, 81]. So far, CCN is the only member of this group that has been detected in the gas-phase. Like CN, the electronic spectrum of CCN was observed more than twenty-five years [128] before its pure rotational spectrum was measured [142]. Laboratory detection of longer C2n N radicals has proven to be difficult because they, like CCN, are expected to have 2 Π electronic ground states and small dipole moments [144]. Non of the Cn N molecules with n even have yet been detected in space. Recently, the feasibility of C2 N detection in space was proposed by Mebel & Kaiser [127]. Therefore, in the course of this work a search for linear C2 N towards the late-type star IRC+10216 has been performed. If it is said that the distribution of more complex and reactive molecules in space remains largely unknown and that many have been observed in only a single or few sources, the same is true for the terrestrial ”sources” of these species. For a long time, production of 4 1 Introduction highly reactive molecules in laboratory was a main obstacle in their detection. So far, only application of sensitive spectrometers in combination with supersonic molecular beams seem to be particular effective for the study of reactive carbon chains. Here the rotational spectra of the radicals are greatly simplified at low rotational temperatures of a few degree Kelvin which can easily be achieved in supersonic expansions. In this work, different kinds of production methods have been investigated (Chapter 2). Excimer laser ablation sources achieve high production rates of pure carbon molecules Cn or silicon-carbon molecules Sin Cm . The Cologne carbon cluster experiment consists of a highly sensitive IR tunable diode laser spectrometer (Chapter 3). Radicals like linear C8 and C10 have been observed with this spectrometer for the first time [61, 14]. Another method to produce radicals is the usage of discharge nozzles. In this work, two kinds of discharge nozzles have been applied, a slit nozzle and a pinhole nozzle. The first one has been used to examine a discharge plasma by the usage of a plasma monitor and the latter one was needed for the production of C4 N, C6 N and C3 N isotopomers. The complete production method as well as the FTMW spectrometer used for the detection of the Cn N radicals is decribed in Chapter 4. Chapter 5 provides a brief introduction of spectroscopic properties of Cn N radicals known so far. Main aspects of the theory of radical spectroscopy are summarized in Chapter 6. Measurements and analysis of the C3 N isotopomers, C4 N and C6 N radicals are discussed in Chapter 7. The role of Cn N chains in the interstellar medium and their detections so far as well as the search for C2 N towards IRC+10216 performed in the course of this thesis is described in Chapter 8. 5 Table 1.1: Known Interstellar and Circumstellar Molecules (Dec Number of Atoms (2-7) 2 3 4 5 6 H2 C3 c-C3 H C5 C5 H AlF C2 H l-C3 H C4 H l-H2 C4 AlCl C2 O C3 N C4 Si C 2 H4 C2 C2 S C3 O l-C3 H2 CH3 CN CH CH2 C3 S c-C3 H2 CH3 NC + CH HCN C2 H2 CH2 CN CH3 OH + CN HCO CH2 D ? CH4 CH3 SH CO HCO+ HCCN HC3 N HC3 NH+ CO+ HCS+ HCNH+ HC2 NC HC2 CHO + CP HOC HNCO HCOOH NH2 CHO CSi H2 O HNCS H2 CHN C5 N + HCl H2 S HOCO H2 C 2 O l-HC4 H KCl HNC H2 CO H2 NCN NH HNO H2 CN HNC3 NO MgCN H2 CS SiH4 + NS MgNC H3 O H2 COH+ NaCl N 2 H+ NH3 OH N2 O c-SiC3 PN NaCN CH3 SO OCS SO+ SO2 SiN c-SiC2 SiO CO2 SiS NH2 CS H+ 3 HF H2 D SH SiCN HD AlNC FeO Number of Atoms (8-13) 8 9 10 11 12 CH3 C3 N CH3 C4 H CH3 C5 N HC9 N C 6 H6 HCOOCH3 CH3 CH2 CN (CH3 )2 CO CH3 COOH? (CH3 )2 O C7 H CH3 CH2 OH H2 C 6 HC7 N CH2 OHCHO C8 H l-HC6 H 2002), [31] 7 C6 H CH2 CHCN CH3 C2 H HC5 N HCOCH3 NH2 CH3 c-C2 H4 O CH2 CHOH 13 HC11 N 6 1 Introduction Part I Characterization of Radical Sources 2 Molecule Source Characterization “ ...one or two atoms can convert a fuel to a poison, change a color, render an inedible substance edible, or replace a pungent odor with a fragrant one. That changing a single atom can have such consequences is the wonder of the chemical world.“ P. W. Atkins, ”Molecules” In this thesis, it will be shown how radicals can be produced and detected by different high resolution spectroscopic tools like IR spectroscopy and FTMW spectroscopy. The aim is to determine as many molecular properties as possible like moment of inertia, electronic structure, vibrational motions, hyperfine interactions, etc. Well known production methods in combination with spectrometers of extreme high sensitivity allow for the detection of molecules and radicals up to linear C13 , HC13 H or HC17 N [62, 123]. However, detection of more complex radicals or molecules of even larger sizes seem to be impaired by the effectiveness of the production methods. For example, for the detection of C13 with a vibrational dipole moment of ∼2 Debye, a particle density of ≈ 1010 cm−3 is required to record the rotationally resolved IR spectrum of an asymmetric stretching mode. In future, molecular sources have to produce even higher yields to allow for the detection of molecules with smaller dipole moments or unfavorable partition functions. For the same reason, it is advantageous to produce molecules with low rotational temperatures because the spectra simplify and gain intensity of low lying transitions. Another important requirement of molecular sources is the stability of their production yield. When spectrometers work at their limit of sensitivity the signal has to be integrated many times in order to increase the signal-to-noise ratio. Therefore, it is apparent that conditions have to be stable over the whole integration time. Furthermore, the characterization of available molecule sources is an important step towards the development of new production techniques. Out of many techniques to produce radicals, two particularly important methods have been chosen which are the laser ablation and the discharge nozzles technique. Both are applicable to the Cologne IR-Carbon Cluster experiment but due to advantage of laser ablation sources to only produce pure carbon molecules Cn this technique is predominantly used. It has resulted already in a number of first detections of linear carbon 2 Molecule Source Characterization relative Scale 10 Frequency [nm] Figure 2.1: Spectra of C2 at 516 nm measured in Basel. A laser ablation source (top) and and discharge nozzle (bottom) was used to produce the C2 radicals. clusters like C8 and C10 [61]. The detection of C13 [62] has shown that even higher members of linear Cn chains can be detected with this type of ablation source. This molecule source is particularly valuable for the production of iso-atomic molecules like Cn , Sin or Men with n=2,3,4,... but also for other species like Cn Sim when using appropriate target materials. The Cologne IR experiment will be described in Chapter 3. The used ablation source is in principle similar to the Smalley type ablation source [169, 41]. However, important improvements have been done to adapt this technique to laser spectroscopical needs which led to a tremendous success in the production of radicals. The improved design has been described in [186, 58, 14]. Details will be discussed further down. So far an excimer laser with 25-50 ns pulses and typical pulse energies of 270 mJ was used as a strong laser light source at repetition rates of 50 Hz. It is evident that longer pulses with higher energies increase the amount of ablated material. However, the main question is whether the ablated material is available in a molecular form or whether it immediately builds grains and soot. For molecular spectroscopic reasons it is desirable to have a broad mass distribution of ablated material in a molecular form with high abundances of particles containing between 3 and 30 atoms. As an alternative to the excimer laser, a Nd:YAG laser was used with pulse energies up to several Joules and pulse lengths of 0.1 to 1 ms. A new experimental setup was built to characterize the ablated material by means of quadrupole mass spectroscopy (QMS). 11 Figure 2.2: C60 optimized mass spectrum. a) mainly small carbon clusters are produced. By changing the flow rate of the buffer gas larger clusters were produced (b). For certain laser and source conditions the production of C60 can drastically be enhanced (c). [80] Laser ablation is not unique in producing radicals or iso-atomic molecules. There are several other techniques like sputtering, electric arcs, the use of ovens, etc. which could in principle as well be used but which have proved in praxis to be less efficient. Considering these alternatives the discharge nozzle technique appears to be quiet comparable to the laser ablation technique and was therefore also taken into account in the experiment. The applied discharge slit nozzle was developed by the Basel group [111, 135]. The conceptual design of a discharge nozzle will be described in more detail in Section 2.2 as well as in Chapter 4.1.2. Between each of the described production methods, there are principle differences concerning the conditions under which the molecules can be detected, i.e. pressure, temperature, chemical composition, jet boundaries and layers. Properties of molecular beams, i.e. supersonic jets, will be discussed in Chapter 4.2. As an example, two spectra of C2 are shown in Fig. 2.1 which were recorded using two different molecular sources. Both sources are able to produce C2 in sufficient amounts for cavity ringdown spectroscopy in the visible region but the rotational temperatures 12 2 Molecule Source Characterization of the molecules are different for both techniques. The laser ablation spectrum has a rotational temperature of about 10-20 K whereas for the discharge spectrum, Trot is in the order of 100 K. For an extended spectroscopic analysis of molecules it is important to produce molecules under different physical conditions, e.g. different temperatures which is possible by using the appropriate molecule source. A detailed understanding of different types of production methods is thus indispensable. For this purpose, a diagnostic tool is needed that focus on yield and relative mass distribution of molecules or radicals produced by different types of molecule sources rather than their structure or other intrinsic properties. Mass spectrometers, for example, have an extremely high sensitivity compared to other spectrometers but do not allow for any information beyond pure mass-per-charge distribution. Beside this restriction, they can provide valuable information on chemical reaction mechanisms. Broad scans covering molecules of nearly every size allows for source controlling which can support the production of certain species. A famous example is the discovery of the Buckminster fullerene C60 in 1985 by Smalley, Curl, and Kroto [108, 41, 169] 1 . For this, a laser ablation source was monitored by a mass spectrometer so that experimental conditions could successfully be varied in order to yield C60 in huge amounts (see Fig. 2.2). 1 In 1996, Kroto, Smalley, and Curl have won the Nobel prize in Chemistry for their discovery of C60 . 2.1 Ablation Technique 13 Liquid 100 Liquid−Vapor Triple Point Pressure (ATM) 10 Solid 1 Vapor 0.1 3400 3600 3800 4000 4200 4400 4600 4800 Temperature (K) Figure 2.3: Pressure-temperature diagram of graphite [106, 155]. 2.1 Ablation Technique Laser ablation can be described as removal of material by applying an intense light pulse of high energy onto a target in order to vaporize solid or liquid materials. Under atmospheric pressure graphite is solid and has a transition to the gas phase at temperatures higher than 4000 K, see Fig. 2.3. Laser ablation can achieve high temperatures and high gas phase carbon densities and is therefore an ideal tool for spectroscopy on carbon clusters and other carbon bearing radicals. The effect of intense laser light on solid material such as graphite results in lattice vibrations, electronic excitations or direct ionization which causes bonds to break. Depending on the laser power irradiated, the material vaporizes or liquefies. With beam intensities higher than 108 W/cm2 , more material is vaporized while consequently less is liquefied [86]. The produced vapor does not condense immediately on the surface but flows off thus perturbing the conditions of a local thermodynamic equilibrium. The number of ablated carbon atoms depends on the irradiated laser wavelength, laser power and the evaporation heat of graphite. Only a small fraction of the radiation is directly absorbed. If n denotes the real and κ the imaginary part of the complex refraction index, than 2 +κ2 reflectance R for a perpendicular irradiation on graphite is R = (n−1) . The value of R (n+1)2 +κ2 14 2 Molecule Source Characterization cap plasma pulsed helium supply rotating graphite rod reaction channel jet vacuum He 10 bar GV graphite plasma UV−puls adiabatic expansion excimer UV−pulse 248 nm Figure 2.4: The Cologne Laser Ablation Source is between 0.1 and 0.5, depending on the orientation of the light in respect to the graphite structure [18], yielding a fraction A ≡ 1 − R of absorbed radiation. However, the power of the laser light is also an important factor in the absorption process [70, 86]. If the laser power exceeds a threshold value of 108 W/cm2 , a plasma is built on the graphite surface which almost completely absorbs the laser power and partially transfers the energy to the material. This energy transfer takes place due to radiation in the visible and ultraviolet spectral range [86] or via compression waves [70] where the plasma pressure can easily achieve 100 kbar. For the ablation source discussed here, typical ablations of 30-50 ng carbon per pulse are achieved, which corresponds to roughly 1.5 · 1015 carbon atoms [7]. A graphite rod of 1 cm in diameter was used as target which consisted to more than 99.5% of 12 C at a density of 2.25 g/cm3 (Goodfellow/Cambridge). While being exposed to the laser beam, the graphite rod is slowly rotated to ensure a stable and uniform ablation process. 2.1.1 Excimer Laser Ablation The standard technique to produce pure carbon clusters at the Cologne IR experiment is to use a pulsed excimer laser beam at 248 nm wavelength which is focused onto a rotating graphite rod. Thereby, a plasma consisting of carbon particles is produced which accelerates the ablation process. Helium at a backing pressure of 10 bar flushes the vaporized graphite in an adiabatic expansion into the vacuum chamber causing fast condensation of single atoms to small carbon clusters. Within a few µsec the temperature of the carbon vapor drops down from several thousand Kelvin plasma temperature to a few Kelvin rotational temperature of the condensed clusters. A total amount of 1013 –1014 clusters of different sizes are produced with every single laser pulse [58]. A two stage roots blower unit and a vacuum rotary pump keep the chamber pressure below 10−1 mbar. Carbon clusters with up to 13 atoms have been produced in sufficient amounts for infrared absorption detection [60]. Carbon clusters seeded in a flow of buffer gas usually readily separate down stream, most likely due to differing formation times. It is thus possible to clearly distinguish between clusters of different sizes, i.e. small clusters come first while larger ones come later. 2.1 Ablation Technique 15 Figure 2.5: Jet produced by excimer laser ablation technique. The molecular probe region for spectroscopy is 1-2 cm downstream the source exit. Left: Side view of the ablation jet with a sketch of the source. The jet dimensions are roughly 17 cm both in height and length. A compression zone caused by the edge next to the source is visible at the lower boundary of the jet. Right: Top view of the jet. Jet width is roughly 1-3 cm. For the Cologne IR experiment, a KrF excimer laser (Lambda Physics, LPX200) is used. This laser can operate at repetition rates of up to 50 Hz with pulse lengths of 25-50 ns and pulse energies up to 500 mJ. For the production of carbon cluster pulse energies of 200 mJ result in an optimal yield. The evaporation heat of graphite is λ = 716.9 kJ/mol [155], i.e. the maximal number of particles that can be ablated is 1.7 · 1017 , assuming a total absorption of laser power in the vaporization process. The excimer laser has a ’square’ beam profile. When the light is focused by a MgF2 lense (f=50cm), the beam has a size of 0.45 x 0.05 mm2 on the surface of the graphite rod (see Fig. 2.8). Energy densities of 1 · 109 W/cm2 are therefore achievable. The laser ablation source is made of stainless steel and consists of two parts, the main body and the cap, as can be seen in Fig. 2.4 (left). At the rear side, a solenoid valve can be attached and a hole of 2 x 2 mm2 allows for inflow of buffer gas. The gas preexpands into a cross sectional area of 12 mm x 1mm at the graphite rod. The visible graphite rod area at the channel surface is 70mm2 . The source exit consists of a slit like reaction channel sized 12mm x 0.9mm in cross section and 8mm in length. The excimer pulse enters the ablation source through the exit slit. The square form of the UV beam profile results in a line focus at the graphite rod. After several hours of operation the reaction channel is clogged by soot and has to be cleaned. The laser ablation process with subsequent adiabatic cooling is shown in Fig. 2.4 (right). First, the valve is opened and a buffer gas (e.g. He) with a backing pressure of 10 - 15 bar flows through the nozzle. During this flow, an excimer pulse is released which causes the formation of a dense plasma of vaporized carbon atoms 16 2 Molecule Source Characterization C3 C9 C13 density of particles nρ [#/cm3 ] 2.3 · 1012 4.2 · 1010 9.9 · 109 particles NC per pulse 13 ≈ 10 ≈ 1.7 · 1011 ≈ 4 · 1010 mass [ng] 0.6 0.03 0.01 rel. abundance Ci / C3 1 1/55 1/230 Table 2.1: Particle numbers of C3 , C9 and C13 using excimer laser ablation and ions. For each individual measurement, the time delay between the opening of the valve and the excimer pulse has to be adjusted to achieve optimal jet conditions. In case of an incorrect time delay the molecule yield drops because of insufficient cooling. The reason for this is that the hot atomic carbon vapor is cooled by the buffer gas at room temperature and additionally small molecules can form via three-body collisions in an endothermal process. The formation of molecules and radicals mainly happens in the reaction channel where the density and pressure is high enough for a condensation process. The most efficient cooling process is adiabatic expansion of the gas into the vacuum chamber where the molecules are then available for spectroscopic detection (see Fig. 2.5). From IR measurements on linear carbon clusters [58] it has been shown that mainly small molecules are produced in the excimer jet (see Tab. 2.1). C3 is found to be much more abundant than C9 and C13 . It is remarkable however, that long carbon chains like C13 are also produced in sufficient amounts for IR detection. So far, only linear carbon chains [189] or silicon-carbon clusters [188, 187] have been found. Interest in the detection of molecules with cyclic, polycyclic or cage like structure which probably correspond to larger species is increasing, though. There has never been a mass spectroscopic analysis of the here introduced Cologne laser ablation source and there is no information at hand whether larger molecules are formed in the excimer jet or not 2 . These facts led directly to the question whether it is possible to change conditions in a way to increase the yield of medium sized to large molecules (e.g. C10 - C60 ). Murray et al. [138, 197] employing material research experiments investigated the effect of lasers with different wavelength on graphite targets. They compared pulsed laser depositions of carbon films using a KrF excimer laser at wavelength 248 nm with pulse lengths of 15 ns with those of a Q-switched Nd:YAG laser at 1064 nm, also having a pulse length of 15 ns. The fluence of the excimer laser at the pyrolytic graphite target was 3 J/cm2 while that of the Nd:YAG was 2.7 J/cm2 . In a time-of-flight (TOF) mass spectrum positive ions ejected from the target were investigated3 . In the Nd:YAG spectrum, many peaks corresponding to carbon clusters C+ n of size 1 ≤ n ≤ 27 appeared + + with most intense peaks for C11 and C15 . Contrary to that, the TOF mass spectrum 2 Fullerenes can also be produced in a laser ablation source [169, 41] but their production has not yet been proved in the case of the Cologne laser ablation source. 3 No buffer gas was applied. 2.1 Ablation Technique 17 pulsed Nd:Yag @ 1064 nm IR beam vacuum chamber skimmer / orifice Quadrupole Jet ion detector quadrupole mass filter pulsed ionization chamber / ion optics amplifier Figure 2.6: QMS experimental setup + for the excimer laser ablation only had significant peaks at C+ 2 and C3 . Murray et al. concluded that “the ejected species [...] are dependent upon the laser wavelength” [138]. Since these results are in agreement with the investigations for the Cologne excimer laser ablation shown in Tab. 2.1, application of a Nd:YAG laser rather than an excimer laser appeared to be useful for the production of medium sized and large carbon clusters. 2.1.2 Nd:YAG Laser Ablation For the characterization of a Nd:YAG ablation it seemed desirable to directly investigate the mass spectrum. For this reason, a new experimental setup had to be build (see Fig. 2.6). The Nd:YAG laser (Baasel BLS 700) used for this work operates at a wavelength of 1064 nm with pulse energies of up to 15 Joule. The laser has no option of Q-switching and thus has pulse widths from 0.1 to 1 ms instead of 25 - 50 ns. Important technical data of the Nd:YAG and the excimer laser are given in Tab. 2.2. As detection device, a quadrupole mass spectrometer was applied that works in two modes, the first of which has a mass range of 1 - 100 amu with high mass resolution while the second works between 100 - 400 amu. This mass spectrometer had previously been used to analyze the electric arc spectrum of graphite and is described in [58]. It is designed to detect neutral species and contains an electron impact ionizer, a quadrupole mass filter and an 18 2 Molecule Source Characterization ion detector. The ion source works with an adjustable emission current between 0.04 and 5 mA and electron energies of ∼ 90 eV. The ionized molecules are mass selected in a mass filter which consists of four rods with 4 mm diameter and 200 mm length. Field radius of the latter is 3.45 mm 4 . Mass resolution m/∆m at mass 100 is better than 100 for the first modus (1 - 100 amu) and 50 for the second modus (10 - 400 amu) 5 . A multiplier is used as ion detector with an output current between 10−6 and 10−12 A which is subsequently amplified and transformed into a voltage signal which is then send to a computer. The first question to answer was whether a 0.1 -1 ms Nd:YAG laser pulse ablation onto a graphite target without the use of buffer gas results in a similar mass spectrum as reported by Murray et al. [138], i.e. can clusters of the size C11 and C15 be found or not? Murray et al. had looked for the ionic species and found that C+ n clusters produced by Nd:YAG laser ablation have kinetic energies less than 5 eV. In a plasma the mean kinetic energies of neutral species is usually well below those of the corresponding ions. The potential settings - which are important for the guiding of the ions through the quadrupole - were well adjusted for particles with kinetic energies of less than 1 eV but could not be adjusted to arbitrary potentials. A mass spectrum of a Nd:YAG ablation is shown in Fig. 2.7. The hydro-carbon molecules in the spectrum originate from the interaction with the rest gas, i.e. with H2 O. Other species like OH and O are produced by fragmentation processes in the electron impact source due to the high electron energies. Since ionization potentials of pure carbon chains Cn with n=3-25 are between 7 - 12 eV [189] a mean kinetic electron energy Ee of 20 eV would probably have been sufficient. Nevertheless, Ee was kept at 90 eV. A lot of soot was produced during the experiment and a protection glass in front of the IR mirror became polluted very quickly 6 . Therefore the vacuum chamber had to be opened regularly. As a consequence, it was not possible to let the vacuum chamber be evacuated over a longer time, e.g. several weeks, to achieve background pressures below 10−8 or 10−9 mbar which would have been useful to avoid interaction of the ablated material with the rest gas during measurements. The measurements reveal the predominance of atomic C while C2 was less abundant by a factor of 10 and C3 was already close to the detection limit. Larger clusters could not be detected. The relative abundance C:C2 :C3 was found to be 100:6.6:2.4. A comparison of different kinds of carbon cluster sources reveals significant differences concerning the amount and distribution of produced carbon molecules as can be seen in Tab. 2.3. Typically, usage of laser ablation sources results in a broad distribution of cluster sizes in the lower mass region from C1 to C10 [11, 107] with abundances that allow for absorption spectroscopy. Gas aggregation sources like thermal vaporization of graphite (vaporization in an oven) [193, 53, 198] or the Langmuir-method (surface method, electric arc) [89, 168, 54] produce 4 The field radius denotes half the distance between the rods. The mass resolution is dependent on the operation frequency and dc- and ac voltage of the mass filter. As a first approximation it can be said that the better the mass resolution the lower the sensitivity. Details are given in [58] 6 Mirrors can be protected with a Helium flow but this feature was not implemented. 5 2.1 Ablation Technique 19 excimer (KrF), Nd:YAG, Baasel BLS 700 Lambda Physics LPX 200 (not Q-switched) wavelength 248 nm 1.064 µm pulse duration 25 - 50 ns 0.1-1 ms repetition rate single pulse - 50Hz single pulse - 100 Hz pulse energy max. 500 mJ max. 15 Joule aver. power max. 8 W max. 50 W 2 beam diameter/size 5-12 x 23 mm 6 mm beam divergence 1-3 ≤ 6 mrad (full angle) mode multi mode multi mode Table 2.2: Technical data of applied ablation lasers. soot in large abundances but nearly no small carbon molecules (see Tab. 2.3). Using these schemes, the Nd:YAG laser ablation process resembles more a thermal or electric arc process than an typical laser ablation process. This can be explained by the following considerations. Nd:YAG laser can produce long laser pulses with high pulse energies which increase the ablation rate while the intensity, i.e. the power per area, decreases. If a laser does not exceed an intensity of 108 W/cm2 , the ablation process is not dominated by direct sublimation but rather by melting-vaporization processes. Qswitched Nd:YAG or excimer laser have pulse widths of 15-50 ns and thus high intensities at the focus point. The target material absorbs the energy during a time interval that does not allow for heat conduction effects. Thus, most of the energy is directly used for sublimation. If the pulse width of the laser is increased, heat conduction dominates the energy transfer and melting occurs. The intermediate step through a liquid phase can be very efficient in terms of ablation yield, however, most of the material is ejected as soot rather then in molecular form. To gain more insight into these processes, the ablated material was investigated under a microscope (see Fig. 2.8 and 2.9). Aluminum has been used to determine the focus size of the Nd:YAG laser beam. The beam focus area is about 3 · 10−5 cm2 in the case of the Nd:YAG laser and roughly 2 · 10−3 cm2 in the case of the excimer laser. If a pulse energy of 200 mJ with a pulse length of 100 µs is applied, maximum intensity at the target can reach 7 · 107 W/cm2 which is still below the threshold value of 108 W/cm2 . If instead the excimer laser with a 200 mJ pulse energy is applied, the intensity is 1.8 · 109 W/cm2 . As Fig. 2.8 (e,f) and 2.9 (e,f) show, melting of graphite is much more emphasized by using a Nd:YAG laser compared to an excimer laser. This indicates that much more soot is produced using a Nd:YAG laser instead of an excimer laser. However, meltings occur in both cases as the small carbon blebs indicate. Graphite is difficult to analyze under an optical microscope because of its low contrast and strong light absorption in the visible frequency range. Therefore, differences concerning ablation processes can better be investigated if silicium or glass are used as target materials rather than graphite. The melting zone of an silicium wafer 20 2 Molecule Source Characterization Figure 2.7: Mass spectrum of Nd:YAG laser ablated graphite rod. Background pressure was 1.4 · 10−5 mbar. after being exposed to a Nd:YAG laser pulse is clearly visible in Fig. 2.9(d) whereas the same energy applied by an excimer laser shows no sign of melting. Ablation of graphite without the use of buffer gas is of no use for absorption spectroscopy since the molecule yield is not high enough for that purpose. Therefore, as a next step the Nd:YAG laser irradiation had to be combined with a buffer gas flow in order to achieve high molecule production rates. The attempt to record a mass spectrum of an ablation jet by using a Nd:YAG laser together with Helium as buffer gas failed completely. This can be explained by two reasons. 1) If no buffer gas is used, the pressure in the vacuum chamber (about 10−4 - 10−5 mbar due to the use of a turbo molecular pump with 1500 l/s) changes to maximal 10−3 mbar which can still be handled by the mass spectrometer without the need of using a skimmer 7 . If a buffer gas is used, even at moderate backing pressures of 1 bar a skimmer is absolutely necessary to avoid damage of the secondary electron multiplier (or channeltron) as well as to avoid collisions and interactions of the ions in the mass filter. Furthermore, a skimmer is needed to avoid turbulences of the gas flow in front of the hole. Turbulences cause collisions and thus a change in the chemistry of the jet which is undesired. On the other hand, the introduction of a skimmer brings in certain difficulties and disadvantages. For example, the skimmer has 7 This type of experiment resembles more a drift experiment in which the ablated species have time to interact with the rest gas in the vacuum chamber before they are ionized and detected. Therefore, also species like C2 H and C2 H2 can be found (Fig. 2.7). 2.1 Ablation Technique C-Cluster C1 C2 C3 C4 C5 C6 C7 C8 C9 C10 21 Laser thermal electric arc [11] [107] [175] [53] [89] [168] [168] 37.0 56.0 99.0 22.2 35 100.0 100.0 31.4 35.0 74.4 62.5 13 20 35-48 100.0 100.0 100.0 100.0 100.0 61 7-10 2.13 2.5 14.5 3.8 . . 0.6-1.0 7.9 1.6 31.6 5.7 . . . 0.413 . 26.1 . . . . 1.11 . 34.7 . . . . 0.19 . 11.0 . . . . 0.11 . 0.9 . . . . 0.32 . . . . . . Table 2.3: Relative C-cluster concentration of different production techniques. Values of 100.0 in each column indicate clusters with largest abundance whereas all other entries were set in relation to those. [133] to be perfectly aligned with respect to the molecule flow and the ionization chamber of the mass spectrometer. Since the opening diameter of the skimmer has to be chosen small, i.e. of the order 100 -200 µm this can be very difficult. In addition to that, the supersonic gas flow is not stable, i.e. might have slightly different shapes or internal structures. Therefore, proper alignment is not easy. In first approximation, loss of intensity due to usage of a skimmer is primarily determined by a 1/r2 law with r the distance between skimmer and ionizer. Thus r should best be a few mm to cm 8 . In the experimental setup described here the distance between skimmer and ionization region could not be reduced below 15 cm. This results in a loss of signal intensity of two orders of magnitude. 2) If a buffer gas is used to form a supersonic expansion, the plasma spreads far beyond the focus point of the laser ablation. Although a plasma in its entireness is neutral, it can cause significant electrical interferences even through a skimmer. A typical problem was that for masses < 1 a huge artificial signal occurred but no signal was recorded at other masses. The main goal to characterize a supersonic ablation jet with the help of a mass spectrometer could therefore not be achieved. It has to be stated that the mass spectrometer used in the experiment, i.e. the mass filter and ion detector, worked perfectly. Also, laser ablation did not cause principal problems, although there happened to appear soot depositions on the diagonal mirror. What seemed to be more critical for the experiment was the device consisting of skimmer and ionizer which was used to extract the molecules and radicals out of the ablation jet as well as to ionize them. However, this device serves as a mediator between molecule source and molecule detection. Consequently, this part of the setup had to be rearranged and optimized. For the optimization process it is advantageous to have a stable supersonic molecular flow. Laser ablation technique as it was implemented in this specific setup was not able to 8 It seems to exist an optimum distance between skimmer and ionization region as can be read in [165]. 22 2 Molecule Source Characterization serve this purpose. As a consequence, a slit nozzle discharge source was used in further experiments to produce molecules and radicals in a supersonic jet. Although the discharge jet was stable and easy to apply, disturbances very similar to those of Nd:YAG laser ablation appeared. As a result, it was not possible to record mass spectra with the current experimental setup. Further improvements were necessary. 2.1 Ablation Technique 23 Figure 2.8: Excimer laser ablation of different kinds of material. a) and b) show ablations on aluminum foil (thickness 10 µm). In a) the whole area of ablation is shown for which 16 shots at 150 mJ were needed to produce the hole. In b) one shot at 450 mJ results in a very sharp edge, no signs of melting are visible. In c) an ablated glass sheet (thickness 2 mm) is shown. The laser was applied 30 sec at 50 Hz with a pulse energy of 400 mJ. Bubbles (diameter 17 - 40 µm) indicate a melting phase. In d) a silicium wafer after one shot of a 330 mJ laser pulse is shown. For pictures e) and f) a laser was applied to a graphite rod over a period of 10 sec at 50 Hz repetition rate with a pulse energy of 400 mJ. Left side of e) shows the ablation hole with an estimated width of 190 µm; the bright vertical stripe (44 µm in width) indicates the edge of the ablation hole. f) shows the ablation edge in more detail which marks the border line of the melting zone to the unperturbed region. The blebs and bubbles indicate a liquid phase during ablation process close to the ablation hole. 24 2 Molecule Source Characterization Figure 2.9: Nd:YAG laser focused on different kinds of material. Photos a) and b) show ablations on aluminum foil (thickness 10 µm). On the right hand side of a) the effect of a single laser shot (200 mJ during 100 µs) is demonstrated. The hole on the left side of a) is caused by the same laser light applied 50 times. In b), 100 shots of a 1 ms (230 mJ) laser pulse were applied. In the lower left corner it can be seen that the ablation hole is surrounded by a melting zone and a transition zone which has an outer border wall towards the unperturbed aluminum. In c), an ablated glass sheet (thickness 2 mm thick) is displayed. The dark zone at the lower right corner indicates the melting zone of the center ablation region. Picture d) shows an ablated silicium wafer. The square structure in the upper region can be interpreted as focus region of the laser pulse (1 shot, 100 µs, 330 mJ). A segment line in the lower left area marks the border between melting region and unperturbed silicium structure. Pictures e) and f) show an ablated graphite rod. In e), melting of graphite can clearly be seen. In f), 50 laser pulses (1 ms) with an energy of 330 mJ were applied. Small blebs (8-11 µm in diameter) surround the edge of the ablation hole. 2.2 Discharge Plasma 25 enlargement slit (30 x 0.25 mm 2 ) expansion applied potential grounded metal plate insulator (−600 to −1200 V) discharge pulsed valve gas gas Figure 2.10: Discharge slit nozzle 2.2 Discharge Plasma Discharge nozzles are of great importance for the investigation of molecular species and can by applied in many spectroscopic fields, see Fig. 2.11. The discharge slit nozzle used for the mass spectroscopic analysis of the described experiment was developed at Basel by Linnartz et al. [111, 135] and is a versatile tool concerning the production of molecules, radicals and ions. It has been shown that pure carbon clusters can be formed in such a device (see Fig. 2.1) [113, 136] as well as hydro-carbons of many kinds [110] and ions [112]. In Fig. 2.10, a cross section of the molecule source is shown. The orifice of the slit consists of an insulator, a metal plate, a second insulator and two sharp plates which form the actual slit (30mm x 250 µm, 60◦ exit angle). A solenoid valve is used for the inflow of buffer gas and precursor gas with a backing pressure of 2-5 bar and a mass flow between 15 and 30 sccm (standard cubic cm per minute). A stable discharge can be achieved by applying a voltage of 400 - 1000 V (300-400 µs) to the jaws during the gas flow. The discharge strikes to the inner metal plate which is grounded 9 . During this process, the flow is heated up which causes fragmentation of the precursor gas, and thus new species start to form. During the subsequent expansion the gas is cooled down and can reach temperature as low as 100 K (see Chapter 4.2). The investigation of the discharge mass spectrum was performed at Balzers/Lichtenstein in November 2000 using a Plasma Monitor PPM 422 (Inficon AG). For this purpose, the electric discharge nozzle, a gas mixing device and precursor gases were implemented into a setup as shown in Fig. 2.12. The plasma monitor (PM) operates in the mass range between 1 - 512 amu. It is optimized to detect ions in a plasma but is also able to detect neutrals. The PM consists of an entrance orifice, ion optics, an ion source, energy analyzer, intermediate focus, mass filter and an ion counter. In this test setup 9 If the outer plate is grounded and the negative voltage is applied to the inner plate the discharge is instable and arcing to the solenoid valve occurs. 26 2 Molecule Source Characterization Figure 2.11: Jets produced by a discharge slit nozzle. The molecular probe region for spectroscopy is 1-2 cm downstream the source exit. Left: Angular view of the discharge jet. Right: Top view showing a broad zone with two diverging areas of high density. In the Cologne Carbon Cluster experiment the IR laser probes the jet parallel to the slit. The resultant absorption length (4-5 cm) is larger than in the case of the laser ablation source (Fig. 2.5). the discharge (500 - 600 V) nozzle was used with repetition rates of 1-2 Hz and gas pulses of 1 ms. Discharge duration was less than 300-500 µs. If the PM is used in a pulsed operation mode as it is described here, a 103 CPS (counts per second) signal at the PM corresponds to a sensitivity of 106 CPS in a cw plasma. The average pressure in the vacuum chamber was ∼ 3 · 10−3 mbar (N2 equivalent) during operation. An electrically isolated extraction orifice10 instead of a skimmer was used as entrance hole to the PM and placed 2-3 cm apart from the front of the discharge nozzle. Ions are more easily detected than neutrals because they are guided into the PM by an electric field, and alignment of the PM orifice with respect to the jet is less crucial. Thus a large extraction volume can be probed. The detected mass distribution of ions is not effected by any ionization process as it would be the case for neutral species and fragmentation processes are negligible. Kinetic energies of ions in a plasma are in the order of a few to hundred eV. A quadrupole without any energy filter can only handle ions with a certain energy distribution. It is therefore necessary to first determine the energy distribution of the ions with an energy analyzer (± 500V) and then to select the species of a certain 10 Could be pos. and neg. biased or on floating potential. 2.2 Discharge Plasma 27 Vacuum Chamber Plasma−Monitor Discharge Source Sample Gas Jet Ion Detector Quadrupole Mass Filter Energy Filter Ionization Chamber / Ion Optics 24.0 23.9 Mass Flow Controller Voltage Multi Channel Figure 2.12: Experimental setup for mass spectrometry on a molecular beam at Balzer/Lichtenstein, Nov 2000 energy range. For example, if neutral He is ionized in the PM (ion source 70 eV, 1 mA), the energy distribution is Emean = 98 eV with ∆E = 2-4 eV (see Fig. 2.13). If necessary, the PM potentials can be set appropriate to Emean so that even high energy ions (± 500 eV) can be detected. The energy filter can select ions within a certain energy range ∆E for a further mass selection in the mass filter. If only nitrogen was used in the discharge slit nozzle, Emean was found to be 4.40 eV. In Fig. 2.14 (top), cations of the N2 discharge have been investigated and nitrogen complexes N+ n with a broad mass distribution have been found. The mass resolution m/∆m at mass 55 amu was 23. Acetylene (C2 H2 ) diluted in a buffer gas is a precursor for the production of many radicals and ions like Cn H, HCn H, H2 Cn ,... . Fig. 2.14 (bottom) shows a 20 sec scan of a discharge using 0.5% C2 H2 in Ar. The mean kinetic energy of the ions was 4.40 eV. In the lower mass + + + regions, ions like C2 H+ 2 , Ar , ArH3 and C4 H2 could be found but the corresponding count rates of max. 80 CPS are rather sobering. The signal-to-noise ratio of the ions is between 6 - 17 and the ion signals are well resolved. In a second experiment anions have been detected. The measurements showed that anions are less abundant than cations. In both cases the detected ion signal was far from being satisfying. Nevertheless, mass spectra of positive and negative ions could be obtained. For measurements on neutral species a multi-channel analyzer was used to allow for integrations over the relevant time intervals when the discharge was switched on, thereby reducing the noise. Again, the detection of even the faintest rest gas molecules worked 28 2 Molecule Source Characterization Figure 2.13: Energy distribution of He+ produced in the ion source of the plasma monitor, ∆E = 2-4 eV. perfectly but no correlated discharge signal could be recorded. The reason for this failure might be that the PM was optimized for ion detection and hence no skimmer was used since this is less important for ions than it is for neutral species. This lack of a skimmer clearly lowered the chance of detecting neutrals. Another reason may be that the ion optic precedes the ion source in the PM setup. Therefore, the ion source has a distance of 3-4 cm to the orifice which corresponds to an intensity decrease in the order of 101 -102 . 2.2 Discharge Plasma 29 Figure 2.14: Mass spectra of a molecular beam. Top: Discharge on N2 molecules (E=-4.4 eV). Bottom: Measurement of cations on a discharge of 0.5% C2 H2 in Ar (E=-4.4 eV). 30 2 Molecule Source Characterization 2.3 Conclusions and Prospects Detection of ions and neutrals in a supersonic jet makes high demands on mass spectrometers. This is especially true if one considers that the use of a mass spectrometer is intended to monitor conditions within the jet during optimization processes, i.e. with integration times of max. 5 min or less. The main challenge is the development of an effective extraction device to probe molecules and ions of a few mbar pressure gas flow. The most important region to probe the jet is 1 - 3 cm downstream the nozzle exit. Depending on the backing pressure of the buffer gas and the type of nozzle used, pressures up to 10 mbar may occur in regions where the extraction orifice or skimmer is placed. In this experiment the pumping speed is not high enough to reduce the mean vacuum pressure below 10−1 -10−2 mbar during operation with 50Hz discharge/ablation repetition rates. Mass filter usually require operation pressures of 10−4 - 10−5 mbar. The use of a single pressure reduction stage can not be sufficient to achieve this pressure drop. To use a mass spectrometer at high repetition rates a second stage pressure reduction is required. The Distance d between the two orifices (or skimmer and orifice) of the reduction stage should be much smaller than the mean free path of the incoming particles. The exact distance d depends on the Knudsen characteristics of the extracted beam [165]. It has been shown that there is a difference between the detection of ions and that of neutrals. Ions require no further ionization process and can be guided by electric or magnetic fields. Nevertheless it is important to place the ion optics very close to the extraction orifice to enhance the yield. If the focus is on neutral detection, an ionization source has to be placed right behind the skimmer (or orifice). Hence it is clear that a mass spectrometer can only be optimized on either ions or neutrals. Displacement of the ionization source or of the ion optics result in a significant decrease of the signal intensity. In a pulsed experiment with pulse lengths of a few µs up to several hundred µs, a mass spectrometer which intrinsically works in a pulsed mode, like a TOF mass spectrometer is advantageous, i.e. a better duty cycle can be achieved by using a TOF instead of a quadrupole mass spectrometer. This is because a TOF records all masses simultaneously whereas a quadrupole can only detect one species at a time. If a mass scan from 1 to 400 amu is desired, the effectiveness of the TOF relative to a quadrupole is 400. Another principal disadvantage of a quadrupole is that the mass range is usually limited, e.g. from 1 - 400 amu in the here described experiment. TOF have no such mass limits. It is the mass resolution that decreases for higher masses and thus limits the TOF mass range. A quadrupole mass spectrometer needs an energy filter to detect species with high energies or broad energy distributions. This is not needed if a TOF mass spectrometer is used. The use of an energy analyzer provides additional information on the plasma conditions. In most cases this information is less important, and TOFs usually make no use of this possibility because it restricts the mass spectrum to species within a certain energy range. For measurements on less stable molecular beams and supersonic jets, i.e. when fluc- 2.3 Conclusions and Prospects 31 tuations between peak times of a beam occur, the use of an additional ion storage quadrupole device can be advantageous. Independent on the kind of mass filter which is used (i.e. quadrupole or TOF), this device can collect ions for a certain time, e.g. for a time slightly longer than the duration of the plasma pulse, before they are released to the mass filter section. A precise timing is therefore not as crucial as without using this device. Mass spectrometers can be used to optimize experimental conditions or they can serve as monitoring devices for experiments which require time consuming searches for certain species or long integration times. They can also be applied to develop new molecular sources. Although the nowadays available molecule sources are very efficient, it is very likely that the production of larger radicals or reactive molecules in their gas phase, e.g. molecules with a carbon chain backbone between C20 and C60 , will only be possible if new production techniques are developed and tested. This can be of great importance for the understanding of grain growth or for the identification of larger cyclic or polycyclic molecules. It seems that even the limits for the detection of linear species can be pushed further away to allow discoveries beyond molecules like C13 , HC17 N, HC13 H, H3 C12 N, C14 H, ... [61, 123]. 32 2 Molecule Source Characterization 3 The Cologne Carbon Cluster Experiment: The New Setup Good name in man and woman, dear my lord, Is the immediate jewel of their souls: Who steals my purse steals trash; ’tis something, nothing; ’Twas mine, ’tis his, and has been slave to thousands; But he that filches from me my good name Robs me of that which not enriches him And makes me poor indeed. William Shakespeare (1564 - 1616), “Othello” Molecules like pure carbon clusters Cn or other species with no permanent dipole moment can only be observed due to their vibrational or electronic properties. C3 and C5 have been astrophysically detected towards the late-type star IRC+10216 [87, 13]. The knowledge of the ro-vibrational transitions of small carbon clusters in the IR region is therefore important for future astronomical detections of these species. The Cologne carbon cluster experiment combines an effective excimer laser ablation source (see Chapter 2.1.1) with a sensitive, high resolution infrared tunable diode laser (TDL) spectrometer. TDL’s provide a fairly monochromatic beam which guarantee high spectral resolution and high sensitivity, with detection of absorbance as low as 10−6 -10−7 . Radicals like linear C8 and C10 have been observed with this spectrometer for the first time [61, 14]. The experimental setup has previously been described in the work of Berndt [14], Giesen [60] and Winnewisser et al. [194]. TDL’s are not well suited for large scans over wide frequency ranges. Thus spectral searches for the asymmetric stretching modes of gas-phase carbon molecules is essentially guided by the infrared vibration spectra of cold matrix isolated carbon clusters (see e.g. Shen et al. [166]) and their assignment to asymmetric stretching modes by ab initio calculations (e.g. Martin et al. [115]). In the recent years the group of Prof. J.P. Maier from Basel has developed a novel matrix isolation technique that allows to deposit mass selected carbon clusters. The infrared spectra of such a matrix can undoubtedly be 34 3 The Cologne Carbon Cluster Experiment: The New Setup Figure 3.1: The Cologne carbon cluster experiment. assigned to a cluster of a certain size. Once the vibrational bands of the radicals are known an aimed search for ro-vibrational transitions in the gas phase can be conducted. 3.1 The Cologne Carbon Cluster experiment Fig. 3.1 shows the new experimental setup of the Cologne carbon cluster experiment. A TDL spectrometer with a spectral resolution better than 5×10−4 cm−1 is used to record the rotationally resolved spectra of asymmetric stretching modes of small carbon clusters. Using a number of different diode lasers the spectral region between 1750 – 2100 cm−1 , where most of the carbon clusters are predicted to have characteristic IR active vibrational bands, can be covered. The IR laser beam intersects the pulsed cluster jet of a laser ablation source (see Chapter 2.1.1 for details of the source) 10 mm downstream from the nozzle. An increase in sensitivity is gained by using a multi-pass optics of Herriott type to obtain 24 passes through the jet. Since most of the diode lasers have multi-mode laser performance, a monochromator with 1 cm−1 resolution has to provide sufficient mode separation. The exiting IR beam is focused on a HgCdTe photo-conductive detector. A fast AC-coupled amplifier allows time resolved detection of the weak absorption signal on a pair of gated boxcar integrators before storing the data by a PC. Part of the IR beam is used to monitor the absorption spectrum of a 3.2 The New Setup 35 Figure 3.2: Sensitivity of IR detectors [97]. In this experiment InSb (J10D) detectors are used in the 5 µm region. reference gas and fringes of a germanium etalon simultaneously with the cluster signal. The 1m reference cell is used to record Doppler limited lines of a reference gas at low pressures. Frequency calibration of the data is accomplished by referencing the lines to the fringe spectrum of the etalon with a free spectral range of 0.016 cm−1 . 3.2 The New Setup Measurements of the linear radicals C13 , C10 and C8 have revealed that for the detection of higher members of the Cn chain further improvements in the sensitivity of the spectrometer are needed. The main problems occurred from the use of a cold head system for cooling the laser diodes, mechanical instabilities in the optical system and electrical interference of the IR signal with the excimer pulse signal. In the course of this thesis a completely new experimental setup was build. Important improvements have been done including: • Laser diodes mounted in a liquid nitrogen dewar instead of using a cryogenic cold head 36 3 The Cologne Carbon Cluster Experiment: The New Setup Figure 3.3: Frequency calibration using CCCS and Calib. The TDL laser was frequency modulated at 5kHz. • The use of InSb detectors for the 2000 cm−1 frequency region instead of HgCdTe detectors • Enhancement of mechanical stability • Setting up a new optical system to avoid astigmatism • Reduction of electrical interferences • Development of new calibration software With the new setup a better signal-to-noise can be achieved and the sensitivity is increased in the frequency region of 2000 cm−1 where vibrational bands of C10 and C8 occur. In the new setup the laser diodes are mounted in a liquid nitrogen dewar instead of a cryogenic cold head using closed cycle technique. Mechanical shocks are completely avoided and baseline fluctuations are significantly smaller. The frequency of the TDL 3.3 Conclusions and Prospects 37 diodes is tuned by temperature and current. In the new setup the minimal temperature of the diodes can not drop below 77 K so that special diodes are required. The laser power of these diodes is between 0.1 and 10 mW as it was the case with the previously used diodes at temperatures of . 4K. The spectral line widths of these lasers are about 30 MHz. A new laser control device and complete new electrical connections of the electronic devices reduced electrical interference with the Excimer pulse significantly. The optical system has been improved by using stable mountings for the optic components and by mounting both optics and vacuum chamber on the same optical bench. The arrangement of the optic components is chosen to minimize astigmatism and loss of signal. The HgCdTe detectors for the reference and etalon signal were replaced by InSb photovoltaic detectors for the measurements at 5 µm wavelength (i.e. at 2000 cm−1 ). In Fig. 3.2 it can be seen that HgCdTe (J15D12, Judson) detectors have their peak sensitivity between 10-11 µm whereas InSb (J10D, Judson) detectors peak at 5 µm. Reference and etalon signal can be modulated in two modes. Mode one (M1) uses a chopper wheel with a modulation frequency at 1.2 kHz whereas mode two (M2) modulates the frequency of the TDL laser (up to 6 kHz). M1 is used for measurements with the laser ablation source. M2 can be used for fast and high precision frequency calibration where the lock-in amplifiers operate in the second derivative mode. The new calibration software CCCS (Carbon Cluster Control Software) is based on the linux operating system using the in-house program dada as program language. CCCS receives and converts the data into ascii format and displays the measurement on the monitor screen. The program is nearly self explaining and comprises of help functions. With an additional program Calib the data can be immediately frequency calibrated, see Fig. 3.3. Test measurements on C3 have been performed. Fig. 3.4 shows the R(28) transition of C3 . 3.3 Conclusions and Prospects High resolution absorption spectroscopy is one of the most powerful tools to characterize small gas-phase molecules. The Cologne carbon cluster experiment has provided highly accurate data on asymmetric stretching transitions of small linear carbon clusters during the last few years. With the new setup even more challenging problems can be approached. At first, searches for hitherto undetected linear chains like C11 and C12 can be started. The vibrational spectra of cyclic C6 and C8 trapped in a cold matrix have been found by Graham et al. [190, 191]. It is thus most likely to find these species in the gas-phase with the Cologne infrared spectrometer. The infrared spectra of combination bands also provide good predictions for the low bending modes of carbon clusters, which are as low as 1 - 4 THz. 38 3 The Cologne Carbon Cluster Experiment: The New Setup Figure 3.4: Rovibrational transition of C3 at 2067 cm−1 (left line). The line at the right hand side is probably due to a hot band transition of C3 . Part II C3N Isotopomers, C4N, and C6N 4 Experimental Setup “ The principle of science, the definition almost, is the following: The test of all knowledge is experiment. Experiment is the sole judge of scientific ’truth’. But what is the source of knowledge? Where do the laws that are to be tested come from? Experiment, itself, helps to produce these laws, in the sense that it give us hints. But also needed is imagination to create from these hints the great generalizations - to guess at the wonderful, simple, but very strange pattern beneath them all, and then to experiment to check again whether we made the right guess. ” Richard P. Feynman, ”The Feynman Lecture on Physics” The rotational spectra of the molecules presented here were detected in a supersonic molecular beam by a Fourier Transform Microwave (FTMW) experiment carried out at the Harvard-Smithsonian Laboratory, Cambridge MA/USA. This spectrometer (Fig.4.1) has been used to detect over 80 new reactive molecules during the past six years [179] and the experimental setup has previously been described in McCarthy et al. [118, 122]. The basic principle of the spectrometer is as follows. A pulsed supersonic molecular beam of an organic precursor gas heavily diluted in an inert gas is produced by an commercial solenoid valve (General Valve Co.). Reactive molecules of many kinds are made by applying a small electrical discharge in the throat of the supersonic nozzle, prior to expansion of the gas into the large Fabry-Perot cavity of the spectrometer. As the molecular beam traverses the cavity the molecules are irradiated by a short (1 µs) microwave pulse at frequency ν0 . The microwave pulse induces polarized resonant transitions of the molecules, ions or radicals. At that time the molecules are under far different conditions than they were prior to the expansion. Important properties which have to be considered are the translational, rotational and vibrational temperatures, as well as the pressure, and density of the molecules. After the polarizing pulse is switched off the polarized gas coherently emits at its resonant frequencies. This free-induction decay (FID) is detected time resolved by a superheterodyne receiver and subsequently Fourier transformed to 42 4 Experimental Setup give the spectrum in the frequency domain. The spectrometer operates from 5 to 40 GHz and is fully-computer controlled to the extent that, in spite of the small spectral coverage of each setting of the Fabry-Perot (< 0.5 MHz), automated scans covering wide frequencies and requiring many hours of integration, can be conducted. Rotational lines of known molecules are routinely monitored for calibration. The experimental part can be divided in three main sections, i.e. the production of the radicals (see Chap. 4.1) using a gas flow controller and a discharge nozzle, their cooling in an adiabatic expansion (see Chap. 4.2) and the subsequent detection in the FTMW spectrometer (see Chap. 4.3). 43 . 70 cm liquid Nitrogen Fabry−Perot cavity Gas input 5 4 ν + δ o ν o Pulsed discharge nozzle Signal antenna Polarization pulse antenna 3 Vacuum chamber PIN switch +30dB 6 Diffusion pump trigger PIN switch Mirror 36 cm 2 30 MHz 9 1 ν − 30 MHz o ν + δ o . 30 MHz source Frequency synthesizer ν − 30 MHz o 30 MHz Bandpass δ + 30 MHz 7 +20dB 0 to +90 dB filter δ Low−pass filter 8 Figure 4.1: Block diagram of the FTMW low band system, 5 - 25 GHz with supersonic jet. (1) The control computer sets the synthesizer frequency (5 - 26.5 GHz) to νo -30 MHz. (2) The output is mixed (single sideband modulator) with a 30 MHz signal which results in a frequency of νo and is amplified (’drive’ amplifier) (3) A 1 µs pump pulse triggered by a ’polarization’ PIN switch enters the FP-cavity via an antenna and (4) polarizes the incoming molecules in the vacuum chamber. The TE00q mode of the radiation field is indicated by the dotted lines between the mirrors. At this stage the molecules have undergone significant cooling due to the adiabatic expansion in the supersonic jet and rotational temperatures of a few Kelvin can be achieved. (5) During the relaxation process the molecules emit signals at certain transition frequencies νo +δ During this process the ’polarization’ PIN is switched off to suppress noise from the drive amplifier. (6) The Free Induction Decay (FID) signal of the molecules is received at the signal antenna and amplified (front-end amplifier). (7) It passes a second PIN switch and is then mixed with a νo -30 MHz signal from the synthesizer. The image rejection mixer has an internal amplifier (+20dB). (8) The resulting δ+30 MHz signal passes a bandpass filter which discriminates the broadband noise and a second mixer transfers the signal down to the 1-MHz video band using a 30 MHz signal. (9) The computer records the filtered and amplified signal in the time domain. 44 4 Experimental Setup 4.1 The Production of CnN Radicals Radicals can be produced in many ways [58], e.g. by laser ablation, cw glow discharge, and pulsed discharge followed by a supersonic expansion. In the present work the latter method was used and the production of the radicals consisted mainly of two steps. (1) To find and produce the appropriate stable precursor gases, which are then mixed through a gas flow controller (see 4.1.1). (2) To choose and build an effective discharge nozzle and optimize the conditions for a maximum radical yield (see 4.1.2). 4.1.1 The Precursor Gases To achieve an efficient way of producing long carbon chain molecules like HCn N, CH3 Cn N, H2 Cn , and even for open-shell molecules like Cn N with n≥3 it is advantageous to use moderately large organic precursors (e.g. cyanoacetylene and diacetylene), rather than commercially available gases such as acetylene (C2 H2 ) and nitrogen (N2 ). Cyanoacetylene (HC3 N) was used for the production of C4 N and C6 N as well as for the production of all 13 C isotopic species of C3 N (in this case 13 C isotopic acetylene had to be added). An appropriate synthesis of HC3 N has been described in Murahashi et al. [137] and Moravec [134] together with suggestions for further readings. HC3 N can be stored at -5◦ C for several months without evidence of polymerization and a vacuum distillation process prior to use was not required. The sample can be produced to a purity of more than 98% as has been shown by 1 H NMR spectroscopy [66]. Once the precursor gases are produced they are heavily diluted in an inert gas like argon or neon which is necessary to maintain a steady gas discharge and to achieve low rotational temperatures of 3 K in the adiabatic expansion following the discharge, (see Fig.4.2). In general there seems to be no ready made recipes for the optimum mixing ratio of precursors to yield a certain radical in sufficient amounts; the work has nearly always to be done by trial and error. In the case of Cn N chains the pioneering work was already done by Gottlieb et al. [66, 121] and Ohshima et al. [142] and had only slightly to be changed. A good mixture for supersonic expansions proofed to be quit different from mixtures used in gas flow cells (see Gottlieb et al. [66]). The best results for C3 N production in a glass tube using a glow discharge where achieved by using HC3 N and N2 or He with an equal molar mixing ratio at 3.3·10−5 bar (3.3 Pa). For experiments using a supersonic expansion the best conditions were found to be the following: 13 C-isotopic C3 N: In this work the strongest lines of the 13 C-isotopes of C3 N were obtained with a mixture of 0.02% HC3 N and 0.07% 13 C-enriched acetylene in Ne. The 13 C-enriched acetylene is a 50% H13 CCH, 25% H13 C13 CH and 25% HCCH mixture produced by hydrolysis of Li2 C2 containing an 1:1 mixture of 12 C and 13 C that was prepared by the NIH Stable Isotope Resource at Los Alamos National 4.1 The Production of Cn N Radicals 45 Laboratory [121]. For speed and convenience all measurements on isotopic species were done with enriched samples and in the case of the 13 C isotopic species an enhancement of the signal by a factor of 2-3 was achieved. An other, but less effective, way to produce isotopic cyanoacetylene is to use a C2 H2 /13 CO/inert gas mixture. For the CC13 CN species a lot of energy is needed to break the strong C≡N bond in the HC3 N precursor. This problem can be circumvented by using HCC13 CN as precursor which can be produced using chemical reactions like K13 CN + H3 PO4 → H13 CN + other and H13 CN + C2 H2 → HCC13 CN + other. CCC15 N: For the production of CCC15 N only a mixture of 0.2-0.3% CH3 C15 N (Acetonitrile) in Ne was used and for the strongest lines a 6 times stronger signal than those of the 13 C isotopic species was observed. C2 N: Ohshima et al. [142] used a 0.15% CH3 CN or CCl3 CN mixture diluted in Ar where the CCl3 CN/Ar sample produced a two times stronger signal. 1 C4 N and C6 N: For the C4 N and C6 N production it was sufficient to take 0.05% HC3 N in Ne. The relation of the signal intensity between CCCN and C4 N is ∼15 and under optimized experimental conditions, the absorption intensities of the strongest lines of the C6 N are only two times weaker than those of C4 N. The total mass flow which entered the vacuum chamber was 14 sccm (standard cubic cm per minute) for C4 N and 32 sccm for isotopic C3 N and C6 N. 1 C2 N can also be produced in a glass tube experiment using a microwave discharge of CF4 with CH3 CN as Kakimoto & Kasuya have shown in [98]. 46 4 Experimental Setup . 4.3 cm teflon insulation Mirror Discharge Source copper electrode 3mm thick Sample Gas Jet 30 sccm gas 5−6 mm . Vacuum chamber 24.0 hot trigger 23.9 Discharge Mass Flow Controller Voltage 1100−1250 V time Voltage trigger on/off . H13 CCH 2.4% HC 3 N Ne 2.1% 95.5% Current oscilloscope Figure 4.2: The Production of Cn N radicals. On the left: precursor gases (H13 CCH, HC3 N) are mixed with a buffer gas (Ne) and fed to a pulsed discharge nozzle before it adiabatically expands into the vacuum chamber. On the top right: the discharge nozzle as it is used for the production of intermediate sized radicals. On the bottom right: The discharge voltage and current is monitored. In most cases an efficient radical production is indicated by a fine fringed current line. The dashed vertical cursor which frame the current line within the 2ms voltage supply marks a separation of 1-1.5 ms. 4.1 The Production of Cn N Radicals 47 . . copper electrodes 5 − 10 mm Teflon spacer Gas copper electrodes 10 − 20 mm Teflon spacer Gas . . . Pulsed nozzle Teflon spacers . Pulsed nozzle Teflon housing . 10 mm Teflon spacer Teflon housing . Figure 4.3: Nozzle for the production of molecules and radicals. On the left: Discharge nozzle optimized for the production of radicals like C4 N. On the right: Discharge nozzle for the production of molecules like HC17 N. The left part on each figure is a solenoid valve with an orifice of 1mm. 4.1.2 The Discharge Nozzle Discharge nozzles similar to the ones used for the here presented experiments were first described by Schlachta et al. [164] in 1991. Since then many other work groups used discharge nozzles of the same type to produce radicals and molecules. The first long carbon chain molecule of astrophysical interest was the cyanopolyyne HC9 N observed in the laboratory by Iida et al [92] in 1991. The same group showed that also reactive molecules which previously were detected in glow discharges in glass tubes (e.g. C3 N, C4 H [66], and C6 H [146]) can also be effectively generated in supersonic expansion discharge nozzles. The basic geometry of the discharge nozzles which were used for this work is as follows. A solenoid valve through which the sample gas enters the nozzle is mounted on a Teflon housing which contains a Teflon spacer (5-18 mm in length), two oxygen-free highconductivity copper electrodes (1mm thick) separated by a spacer (4-10 mm) of the same dielectric. In some cases a third spacer is added to function as a reaction channel but this is usually only the case if long closed-shell molecules are to be studied. The dimensions of the electrodes and the insulators are quite critical for the efficient formation of discharge products. Applying a short pulse discharge in a small region between the electrodes where the pressure of the flow gas is still high and a subsequent expansion of the gas in a supersonic free jet, where almost no collision occurs, generates a relatively high concentration of transient species. These conditions are very different to a dc discharge in a flow gas cell system which is far inferior in terms of the plurality of radical production as known so far. This emphasizes the importance of the right discharge conditions, such as dimension of the discharge unit, the applied discharge voltage, discharge timing, and as already discussed the composition of the sample gas. 48 4 Experimental Setup . Pin 1 2 3 4 5 6 7 8 solenoid valve gas 11 mm i.d. 3 mm 1 mm copper 4 mm teflon discharge zone reaction channel ground −HV Figure 4.4: Test nozzle to optimize the geometry of the discharge nozzle. Eight electrodes (PIN) give the possibility to freely choose the distances of the electrodes in use, i.e. the discharge zone can be varied by just reconnecting the electrodes from outside the vacuum chamber. Also the length of the reaction channel is adjustable from 0-30mm. Highly reactive molecules as well as long carbon chain molecules like HC17 N can be produced with this nozzle. Restricting oneself to the here described geometry of the nozzle there is a possibility to systematically vary the length of the discharge zone and of the reaction channel during the operation of the nozzle. As part of this thesis a test nozzle as seen in Fig.4.4 was build to study the most effective electrode and reaction channel positions and lengths. Two electrodes were used at a time which could be chosen freely out of the 8 electrodes of the test nozzle. The region between cathode and the grounded electrode is called the ’discharge zone’ which is followed by the ’reaction channel’, i.e. the space left until the flow exits the nozzle. With this nozzle many different combinations of discharge zone lengths and reaction channel lengths can be tested and in-situ electrode changing during an experiment can be done from outside the vacuum chamber which ensures similar conditions for each electrode setting in terms of mass flow, valve opening times and gas mixing. The testing and optimization usually works by monitoring the signals of known species; starting with easily to produce ones like HC5 N or C3 N and then refining with more rare species of the same type. There seems to be a principal difference in the production of stable and unstable molecules. In the case of the closed-shell molecules like HC9 N [141] and HC2n+1 N with n=5-8 [118] which have been generated from simple molecules such as (CN)2 and HC4 H (or CH2 CHCN and C2 H2 as used in [141]) more collisions are required to form longer carbon chains. This might explain why for such molecules a nozzle with an additional reaction channel is favorable. Once an optimum setting of ’discharge zone’ and ’reaction channel’ lengths is found an easier to assemble and to clean nozzle can be build (see Fig. 4.3, right and Fig. 4.7). For highly reactive species such as open shell free radicals it is important to avoid any quenching process on the surfaces of electrodes, insulators (spacers) and other molecules. Collisions apparently 4.1 The Production of Cn N Radicals 49 drive the hydrocarbon chemistry to the more stable closed-shell polyynes [118]. The optimal production conditions for the Cn N chains were achieved by using no reaction channel 2 and a 10 mm discharge zone in the test nozzle. This corresponds to a short nozzle as shown in Fig. 4.3 (left) 3 . This nozzle was used with a low-current dc discharge of 1100-1300 V synchronized with a 300-480 µs long gas pulse at a total backing pressure of 2.5 bar and a total gas flow of 30-32 sccm. The discharge current (see Fig. 4.2 bottom right) was typically 10-100 mA and lasted from 1 - 1.5ms per pulse corresponding to an energy per pulse of the order of 100mJ. For the production of open-shell molecules the first (inner) electrode was used as the cathode and the second (outer) electrode was grounded, whereas long closed-shell carbon chains were obtained by using the second (outer) electrode as cathode. The strength of rotational lines can decrease by a factor of 2-4 when the polarity of the electrodes is reversed [118]. For radicals the loss factor can be much higher by changing the polarity. The short discharge nozzle was used for the measurements of C4 N, C6 N and the C3 N isotopomers (see Fig. 4.5 and Fig. 4.6) and it is also described in [37]. 2 3 except the tip of the nozzle These conditions were found to be similar to those that optimize production of the acetylenic free radicals Cn H. 50 4 Experimental Setup Figure 4.5: The short ’radical’ nozzle during a discharge. This nozzle was used for the measurements of C4 N, C6 N and the C3 N isotopes. Figure 4.6: The supersonic jet expansion of the short nozzle. The picture was taken in Cologne using a pure He discharge at 5 bar stagnation pressure and 10−1 mbar background pressure. The straight line in the middle of the picture marks the center line of the gas flow. The jet boundary is indicated by the upper line which arises from the nozzle exit at an angle of roughly 23◦ with respect to the exit plummet. The dashed line indicates the barrel shock. On the right hand side the vacuum chamber window limits the further view on the still expanding jet. 4.1 The Production of Cn N Radicals 51 Figure 4.7: The long nozzles during a discharge. The pictures were taken in Cologne using pure He discharges at 5 bar stagnation pressure and 10−1 mbar background pressure. The discharge nozzle has a total length of 55mm (with tip), both electrodes were separated by a spacer (13mm length, 5mm hole diameter) and the last electrode has a distance of 29mm from the nozzle exit. At the top: The first (on the very left which can hardly be seen) electrode was grounded and the second (the dark, vertical stripe in the middle of the nozzle) on a negative high voltage. Bottom picture: The first electrode (left to the bright discharge zone) had a negative high voltage and the second (in the middle of the nozzle) was grounded. This kind of nozzle is mainly used in the mode shown in the upper picture and is very effective in the production of long carbon chain molecules with an electronic close shell structure, e.g. HC17 N. 52 4 Experimental Setup 4.2 Adiabatic Expansion After the precursor gases have passed the discharge region the particles in the gas undergo an adiabatic expansion into the vacuum chamber, i.e. they form a molecular beam. Spectroscopy on a pulsed molecular beam is used in a number of laboratories because the low rotational temperature of the molecules within the beam allows interesting experiments in the field of Van-der-Waals complexes, clusters and radicals. Compared with experiments in glass tubes under room temperature and medium low pressure 4 the spectra of stable and unstable molecules in a beam are very simple since they are usually free of high rotational and highly excited vibrational states. Although the primal interest is in the molecules created in the discharge the first and main part of this section deals with the buffer gas atoms which usually contribute to more than 98% of the flow particles and which set constraints on the flow conditions in which the molecules of interest are seeded. In the last part of this section the focus is on the interaction of the buffer gas with the molecules and radicals and the spectral behavior in a supersonic expansion of the latter. The problem is mainly approached in a theoretical way to clarify the basic prosseses. It has to be mentioned that deviations from experiment are known but that up to now no theory comprising jet properties, chemical reactions and spectroscopic properties of certain molecular species is available. The experiment described here is set up in a pulsed mode with repetition rates of 2 Hz which requires smaller and less expensive pumping systems than a cw mode. In this experiment a 35cm diameter diffusion pump (Varian) backed by a dual-stage mechanical pump is used to maintain a background pressure of 2.7·10−9 bar (0.27 mPa). Typical peak pressures at 2 Hz repetition rate of the nozzle are 6.7·10−8 bar (6.7 mPa). Assuming a perfect adiabatic process one could argue that the properties of the molecules are mainly determined by the adiabatic equations, like T = T0 p p0 γ−1 γ (4.1) with T0 ≈ 300 K the stagnation temperature, p the background pressure during the expansion 5 and p0 = 2.5 bar the stagnation pressure result in a minimum temperature of Ta & 1.2 K. This however does not consider that the adiabatic expansion happens as a ’free jet’, i.e. a supersonic flow which can show complicated features as seen in Fig. 4.8. Free jets have been investigated by many authors and the main reference for this section is the ’new classic’ book edited by Scoles ’Atomic and Molecular Beam Methods’ [165], the work of Balle & Flygare [6], and McClelland et al. [124]. Molecular beams with pulses longer than 100µs in a vacuum chamber of 1m length are technically rather ”gated” or ”modulated” than pulsed (R. Gentry in [165], p.54), i.e the 4 5 usually in the µbar - mbar region This is not necessarily the mean background pressure but the lower limit can be set to the achieved pressure if there is no gas load. 4.2 Adiabatic Expansion 53 Background Pressure P b Compression Waves M>1 Reflected Shock Expansion Fan P0 ,T0 M<<1 M=1 Mach Disc Shock Zone of Silence M>>1 M<1 Flow Slip Line M>1 Barrel Shock Jet Boundary Figure 4.8: Continuum free-jet expansion [165]. mean free path of molecules moving at a speed of typically 800-1000 m/s is larger than the dimension of the vacuum chamber itself. In the further discussion the molecular beam although pulsed is treated as a continuum free-jet for the time between 500 - 1100 µs after the valve has opened and released the gas because in that time it is assumed that the beam exposes all important features which characterize a continuum free jet. A gas expansion from an region with a stagnation pressure p0 into an area with a background pressure pb becomes supersonic if the equation p0 >G≡ pb γ+1 2 γ γ−1 = 2.05 using γ = 5 3 for atoms (4.2) is fulfilled. G is less than 2.1 for all gases and in this experiment the requirement of Eq. 4.2 is easily achieved. A supersonic flow increases velocity as the gas expands and the Mach number M , with v M≡ (4.3) vs where v is the flow velocity and vs is the speed of sound, is a measure of this, i.e the beam is supersonic if M ≥ 1. Eq. 4.2 ensures that the pressure at the source exit or nozzle ’throat’ is well above the background pressure and the flow is said to be ’underexpanded’. In the subsequent expansion the gas is accelerated so strongly that M can be much larger than 1, e.g. 40 or even larger which means that the particles in the beam have a higher velocity than the local speed of sound. On the other hand information can only propagate at the speed of sound 6 . This means that for a certain region in the flow, the so called ’zone of silence’ (ZOS), the particles in the flow are not influenced by 6 Here, information refers mainly to the pressure and density distribution in the vacuum chamber or the molecular flow. 54 4 Experimental Setup any external conditions imposed on them like the ambient pressure pb which they have to meet downstream. The jet over-expands with M continuously increasing in the ZOS. Of course at some point the supersonic flow is adjusted to the boundary conditions via shock waves, which are regions between the ZOS and the rest of the vacuum chamber that are very thin non-isotropic regions of large density, pressure and temperature. The re-compression usually happens as a barrel shock at the sides and the Mach disc shock at the normal to the centerline of the flow, see Fig. 4.8. In this regions the flow becomes subsonic (M < 1) and can react on the background pressure, walls or other obstacles. Assuming that only background gas reacts on the flow the distance nozzle exit - mach disc location xM is given by r xM p0 = 0.67 (4.4) d pb where d is the nozzle diameter. This formula together with a similar one for the width of the barrel shock has proved to give results which are in fair agreement with observations at the Cologne Cluster Experiment. In this experiment a laser ablation or alternatively a discharge is used where many molecules are optically excited which make the jet visible. In the Cologne Cluster Experiment p0 is 10 bar and pb is of the order 5·10−2 mbar with a clearly visible jet of ca. 17 - 20 cm length within the vacuum chamber which agrees well with Eq. 4.4. On the contrary because of the very low background pressure as it occurs in the here described FTMW experiment the location of the Mach disc would be far beyond the walls of the vacuum chamber. The main constrains on the jet are therefore the walls and mirrors of the vacuum chamber. The shock wave thickness is of the order of the mean free path λ and for the centerline region it can be estimated using kB T (x) λ(x) = √ 2σp(x) (4.5) with kB the Boltzmann constant, T (x) the temperature at the position along the centerline axis x, p(x) the pressure and σ the hard sphere collision cross section of the buffer gas. The distance between the mirrors where the free jet can expand is 70 cm and to estimate the shock front position xS which limits the free jet in the x-direction a detailed knowledge of the flow properties is necessary. Scoles ([165], p.23, Tab. 2.2) gives numerical formulas for the Mach number M as a function of (x/d), with x the downstream coordinate along the centerline axis, assuming an isentropic, compressible flow of an ideal gas 7 (see Fig. 4.9). For a pin nozzle and with (x/d) 1 and gases with γ = 5/3 like He, Ne, Ar, etc. this formulae 8 reduces to x 23 M = 3.232 · (4.6) d This equation is independent of the background pressure and is only valid in the ZOS where M > 1. Once the Mach number is known in that region all the other important 7 8 It is also assumed that there are negligible viscous and heat conduction effects. [165], p.23, Tab. 2.2 4.2 Adiabatic Expansion 55 flow properties can be calculated if the stagnation values like T0 , p0 , n0 (particle density) are given and the following equation −1 T γ−1 2 (4.7) = 1+ M T0 2 together with the adiabatic equations, e.g. Eq.4.1, are used, see Fig.4.9. To get an idea of the numbers involved in the flow process a simple model is applied to calculate some flow properties, see Fig. 4.10, and a small program was written 9 to calculate the Mach number, temperatures, pressure, velocity, collision numbers, etc. at various points downstream the flow axis. The discharge is not explicitly included in the model but is assumed to correspond to a higher stagnation temperature T0 . In the simple model the gas pre-expands into the discharge nozzle so that the pressure drops from the stagnation pressure p0 to the pressure at the nozzle exit pn ∼ 2 mbar 10 . This pressure drop can be estimated by assuming a constant mass flow rate ṁ = ρvA, with ρ the density, v the one-dimensional flow speed, and A the cross-sectioned area of the flow which then can be used to calculate the increase of the flow velocity, i.e the Mach number via ([165], p.19, Eq. 2.9) (γ+1)/2(γ−1) 1 2 γ−1 An 2 = 1+ M A0 M γ+1 2 (4.8) where An is the flow area at the nozzle exit and A0 the flow area at the valve exit. According to that the flow gains already ca. 97% of is final speed in the discharge nozzle and the temperature drops to ∼20K. It follows a free expansion into the vacuum chamber. For the calculation of the free jet properties, as it is summarized in Tab.4.1, the discharge nozzle was now not included and the free jet starts directly behind the valve orifice 11 . The expansion is split into 3 zones. The first is the region of continuum flow where all particles can interchange energy via collisions and one parameter, the Mach number, characterizes all important properties at each point along the flow axis, e.g. the isotropic equilibrium Boltzmann distribution of the velocity. Because the density of the gas decreases rapidly in the expansion the collision frequency cannot maintain continuum flow and a smooth transition to free-molecular flow begins (zone 2). The low background pressure in this experiment causes this transition to happen without any conspicuous continuum shock structure. A measure of this transition is the point xF at which only one collision Z ' 1 is left for each particle to interact during the rest of the expansion. Usually one collision is sufficient to achieve a transitional relaxation but beyond xF the particles remain in their states and the temperature TF of the flow does not change after the transition to a free molecular flow so that TF is the lowest possible temperature which can be achieved during the 9 based on the formulas [165], Eq. 2.3, 2.4-2.6, Tab. 2.2, and [124], Eq. 4 assuming the input values of Tab. 4.1 11 This means that the distances xi given in the Tab.4.1 have no direct meaning to the experimental setup but are included to estimate the order of magnitude. 10 56 4 Experimental Setup Figure 4.9: Top: Mach number along the centerline axis of a free expansion. Bottom: Temperature along the centerline axis of a free expansion. 4.2 Adiabatic Expansion 57 x F beginning transition zone . . . . discharge nozzle 2 gas tube TF A0 . . 5. free jet Tn , p n , v n A nozzle discharge energy . . . . .. .. . . . background . . . gas . . . . T0 , p 0 . . 4 3 1 . . . free−molecular flow . . . . . . . . . . . . TS . . . . . . . . . . . x S begin of shock zone Figure 4.10: Model of flow development. 1. The stagnation region (tube and solenoid valve). 2. The gas enters the discharge nozzle. The pressure drops to pn (Eq. 4.8) A0 is the entrance nozzle area and Anozzle the exit area. Because of the boundary conditions enforced by the walls the expansion is strictly speaking not adiabatic but this fact is neglected here. To calculate the temperature drop Eq. 4.1 is used and Tn is the temperature at the nozzle exit. The gas pre-expands in the nozzle and is gaining flow speed. When a discharge is applied the energy of the flow increases to Tn,d . 3. Free jet. Isentropic flow region. Many collisions occur. 4. At xF transition to free-molecular flow. The translational temperature is ’frozen’. 5. Shock waves appear at xS and re-thermalize the flow gas. expansion. Because of the walls or mirror a shock wave will appear at xS which marks the transition to zone 3. Unwanted collisions with molecules scattered by the surfaces in front of the expansion re-thermalize or broaden the velocity distribution, i.e heat the beam, in terms of an effective temperature. Assuming T0 = 300 K xS can be estimated to be ≈ 58 cm because the mirror distance is 70cm and at that point the mean free path is of the order of 12cm. The highest possible mach number would still be M (xF ) ≈ 43, so that a translational temperature of Ttrans = 0.47K should be achievable. Because the translational temperature changes along the x-axis the mean temperature between the mirrors should be slightly larger than 0.5 K. An alternative approach to estimate the translational temperature based on experiments with iodine in seeded supersonic beams is given in McClelland et al. [124] where the terminal Mach number can be estimated by Mt = F (γ) Kn − γ−1 γ (4.9) 58 4 Experimental Setup with F (γ) = 2.03 for γ = 5/3 and Kn = λ0 /d the Knudsen number, λ0 the mean free path in the stagnation zone 12 , and the ’maximum fractional change in the mean random velocity per collision’ ( (H2 , He) = 0.02, (Ar) = 0.25, and (N e, D2 , N2 , CO, . . . ) = 0.1 − 0.5, [124]). In the case of Ne this results in a minimum translational temperature of 0.1 - 0.3 K depending on which is taken 13 . The precursors are heavily diluted in the buffer gas and can usually be neglected by the calculations of the flow properties. From Table 4.1 it can be seen that there are large discrepancies between the different kind of estimations of the translational temperature of a free jet, i.e. 30% between TF (xF ) (using Eq. 4.6) and Tt (xt ) (using Eq. 4.9)14 . The buffer gases can not be detected by the Fourier transform spectrometer and an direct verification of the here presented flow values is not possible. Instead the translational velocity of polar species (precursors or radicals which are created in the discharge, see Section 4.3) can be measured and compared to the theoretical results as seen in Tab. 4.1, e.g. a velocity of 840 m/s was measured for a molecule which was cooled by Ne atoms in a supersonic jet. If this molecule was in thermal equilibrium with Ne a stagnation temperature of 343 K has to be assumed, see Tab. 4.1. Because the Ne as well as the precursor gases were used at room temperature it has to be concluded that it was the discharge that heated the gas to ca. 50◦ C prior to the adiabatic expansion. 12 In McClelland et al. [124] the formula is described with a mean free path at the nozzle throat, which in Balle & Flygare [6] is interpreted as stagnation value. To calculate λth with the pressure at the nozzle throat p0 has to be replaced by p0 /G, ([165], p.15) which yields higher temperatures of a factor 2. 13 T0 was assumed to be 300 K. For Ar and He the results are T(Ar)=0.6 K and T(He)=7.3 K respectively. 14 For comparison also the values Ta derived by using the pure adiabatic equation 4.1 are given in Table 4.1. Because Trot (molecule) > Ttrans (molecule) and Ttrans (molecule) ≥ Ttrans (N e) it is clear that for Trot (molecule) less or equal 0.5K (see Fig. 4.11) Ta can not be ∼ Ttrans (N e). Therefor Ta does not correctly describe the physical conditions of the buffer gas in the jet. 4.2 Adiabatic Expansion 59 Table 4.1: Free jet flow properties for Ne Experimental / particle parameters γ 5/3 ∼0.2 - 0.5 mN e [amu] 20.18 σN e−N e [nm2 ] 0.24 T0,1 [K] 293.15 p0 [bar] 2.5 T0,2 [K] ∼ 343 pb [mbar] 2.7·10−6 T0,3 [K] 500 pb,load [mbar] 6.7·10−5 gas mass flow [sccm] 30 repetition rate [Hz] 2 d [mm] 1 D [mm] 5 18 Under standard conditions there are ≈ 6.7·10 particles per pulse. Jet properties at xF [mm] M (xF ) TF (xF ) [K] ≈ Ttrans,min Tt (xt ) [K] ≈ Ttrans,min TZ<5 [K] & Trot,min Ta [K] p(xF ) [mbar] v(xF ) [m/s] vt [m/s] ρ(x = 350) [1/cm3 ] T0 (293.15 K), p0 49 48.4 0.5 0.1-0.3 1.1 1.2-4.3 2.3·10−4 776.5 777.1 ∼ 1014 T0 (300 K),p0 44 47.8 0.47 0.1-0.3 1.68 1.2-4.4 2.4·10−4 785 786 T0 (343 K), p0 42 44 0.6 0.18-0.4 2.13 1.4-5.1 3.1·10−4 839.7 840.5 Jet boundaries xM [m] >4 ∅barrel shock [m] >2 γ = cp /cv the heat capacity ratio, from [124] p.950-951, mN e atomic mass of Ne, σN e−N e hard sphere collision cross section of Ne [5], T0,i stagnation temperature i, p0 stagnation pressure, pb = background pressure without gas load, pb,load background pressure with gas load, d diameter of valve orifice, D diameter of nozzle exit, x distance from nozzle exit, xF freezing point with Zr,binary (xF ) ' 1, xt point where terminal Mach number is reached [124], M is taken from [165], the following temperatures T are all translational temperatures, TF (xF ) temperature at freezing point, Tt (xt ) temperature corresponding to Mt with Mt as in [124], TZ<5 temperature at point where only less than 5 collision remain per particle (a rotational relaxation needs less than 5 collisions), Ta as in Eq. 4.1 using pb (first value) and pb,load (second value), p pressure at xF [165], xM location of Mach disk, ∅barrel shock diameter of the barrel shock, v flow velocity at x, vt terminal velocity, ρ density at x. It should be xF ≈ xt and T (xS ) < T (xF ) ≈ T (xt ) < Ta . 60 4 Experimental Setup Molecular collision rates are important for the translational, rotational and vibrational relaxation process as well as for chemical reactions and cause the most important deviations from ideal predictions based on the continuum properties. Two-body collision rates scale with p0 d, whereas three-body collisions required to build molecules out of atoms scale with p20 d. This means that any chemical reaction takes place in the discharge nozzle or shortly after the particles enter the vacuum chamber (x/d) ∼ 1 15 where the density of the flow is still high. After the radicals have formed kinetic processes, such as energy exchange for cooling of internal states, first decrease and finally ’freeze’, i.e. terminate. The effectiveness of such cooling processes depends on the number of collisions experienced by each particle. The total amount of binary collisions Ztot can be calculated by integrating the collision rates Z(x) from x = 0 to x = ∞ and is typically of the order of 102 (pin nozzle) to 103 (slit nozzle) collisions. From this number the number of collisions Zr remaining in the expansion at a given point x along the flow axis can be calculated and gives a measure of the positions of the ’transition zones’ where an average particle will not experience enough collisions to achieve translational or rotational equilibrium. In the case of the translation relaxation only a few collisions Z ' 1 are required and this transition often occurs beyond the point where the velocity ratio v(x)/v∞ ∼ 0.98 [165] 16 , see Tab.4.1 (xF ). Rotational relaxation of small molecules like C3 N need slightly more collisions Z . 5. Simple diatomic molecules may require ca. 104 for the vibrational relaxation. The needed vibrational collision number for large polyatomic molecules and the rotational collision number of most diatomic molecules are of the order 10 to 100 so that the vibrational modes of diatomic do not participate in the expansion [165]. It is not only the collision number which determines the effectiveness of cooling but also the cross section σ of the energy transfer (e.g. Erot → Etrans , Evib → Etrans ) compared with Etrans → Etrans . Usually it is σ(Etrans → Etrans ) > σ(Erot → Etrans ) > σ(Evib → Etrans ) (4.10) so that the rotational energy is smaller than the vibrational energy but higher than the translational energy. In most cases is σ(Erot → Erot ) > σ(Etrans → Etrans ) and an equilibrium within the rotational levels is achieved with a temperature Trot . The population of the vibrational levels usually do not follow a Boltzmann distribution but it is nevertheless standard to speak of a vibrational temperature Tvib which is a temperature corresponding to a Boltzmann distribution which has been approximated to the real distribution. After the expansion the gas has usually the following order of temperatures: Ttrans < Trot < Tvib [43]. The Ttrans , Trot , Tvib values depend on the expansion parameters and are typically Ttrans < 5 K, Trot < 10 K, Tvib < 100 K (Demtröder et al. [43], 3 bar Ar with 5% NO2 with a d=50 µm nozzle.). Grabow et al. [69] who did measurements on the SO 15 16 see [165], p. 25, Fig. 2.10 For this experiment v(x)/v∞ ∼ 0.99 . 4.2 Adiabatic Expansion 61 radical (with X3 Σ− ) reported that the effectiveness of the cooling by a beam expansion depends inversely on the heights of the energy levels to be cooled. Rotational transitions with energy differences of only a few cm−1 are cooled much more effective, i.e. the rotational temperatures Trot can be as low as a few degree Kelvin, whereas vibrational states where found to have Tvib of a few hundred K. If for the rotational relaxation only 3 or less collisions are required the rotational temperature Trot is expected to be between 0.2 - 2.1 K 17 . In Fig. 4.11 a measured C4 N spectrum is plotted together with a theoretical spectrum corresponding to 0.5 K 18 . In the work of Grabow et al. [69] molecules in a beam expansion without a discharge are found to have transitions with low vibrational energies, e.g. OCS, SO2 or SO, which become much stronger in intensity when the discharge is turned on within the nozzle orifice. This is in agreement with Schlachta et al. [164] who studied a variety of diatomic radicals like OH, NH, CN, and C2 and report rotational temperatures of 5-50 K, depending on the expansion parameters, and in many cases vibrational temperatures of several thousand Kelvin when applying a discharge. The main advantages for the analysis of the spectrum due to the cooling of the molecules in the jet can be seen by the formula describing the number of molecules in a certain state νi , Ji −Evib −Erot N0 Y ) ( ) ( N (νi , Ji ) = (2Ji + 1)e kTrot e kTvib (4.11) Z i , where νi is the vibrational quantum number of the vibrational mode i and Ji the rotational quantum number within the vibration νi , and Z the overall partition function: • The population number N (νi , Ji ) reduces to a few ro-vibrational levels, so that the number of the absorption lines decreases drastically. • Because the absolute number N0 remains constant the population of the lower levels increase and instead of many weak lines a few strong lines are conceived in the spectrum, see Fig.4.11 17 18 see Table 4.1 with Tt (xt ) = 0.2 K and TZ<5 = 2.1 K. It is however not unproblematic to make a straight forward comparison between the measured and the calculated intensities. Some people including myself believe that to estimate the rotational temperature of a molecule measured with a FTMW spectrometer the effect of the Q-factor (see section 4.3) has to be considered. The result including this effect would be a rotational temperature of 0.2K. 62 4 Experimental Setup Figure 4.11: Theoretical intensities of the C4 N rotational transitions corresponding to 0.5K together with the measured lines as discussed in Chapter 7.2. 4.3 The Fourier Transform Microwave Spectrometer 63 4.3 The Fourier Transform Microwave Spectrometer The Fourier Transform Microwave (FTMW) Spectrometer at the Harvard/DEAS Spectroscopy Laboratory exists since 1995 and was build following the classical experiment from Balle & Flygare [6, 29] published in 1981. At the moment the FTMW spectrometer operates in two modes, the low band mode between 5 and 25 GHz and the high band mode between 25-40 GHz. Both modes were used for the measurements of the radicals presented in this work. A block diagram of the used FTMW spectrometer in its low band operational mode is shown in Fig. 4.1. A schematic diagram of the FTMW was already published in [122] but Fig. 4.1 reflects some rearrangements which were done since then and it also shows the cooling of the mirrors. Between 1997 and 2000 the sensitivity could be improved by more than an order of magnitude and the results were published in [118]. The spectrometer can only be used within resonant frequencies ν of a TEMmnq mode of the Fabry-Perot (F.-P.) resonator 19 " # 1 (m + n + 1) ν = ν0 (q + 1) + (4.12) π cos(1 − Rd ) with ν0 = c/2d, d the distance between the mirrors, R radius of curvature of both spherical concave mirrors, c speed of light. The experiment is set up in a confocal arrangement of the mirrors (R=d) with d=70cm and the TEM00q modes which are the dominant modes are therefore 1 ν = ν0 (q + 1) + (4.13) π and separated by ν0 =214 MHz. One of the mirrors, the ’drive’ mirror, can be moved and allows the frequency tuning of the spectrometer. To change the resonance frequency from ν1 to ν2 = ν1 + ∆ν by moving the drive mirror from position 1 with d1 to position 2 with d2 = d1 − ∆d the Eq. 4.12 transforms to c 1 1 d d2 ≈d2 c(q + 1) ∆d ∆ν = − (q + 1) 1 = 2 d1 d2 2 d2 yielding ∆d = c∆νd2 c(q + 1) . (4.14) At 20 GHz Eq. 4.13 results in (q + 1) ≈ 90 half wavelength between the mirrors and if a frequency search with step sizes of ∆ν = 500 kHz is desired Eq. 4.14 yields a separation ∆d for each step of ca. 20 µm 20 . 19 20 nomenclature and formulas of the following section are mainly taken from Balle & Flygare [6] This means that also mechanical vibrations can vary the cavity resonant frequency and thereby contributes to the low frequency noise. 64 4 Experimental Setup Using separate antennas, i.e the ’drive’ antenna to connect the F.-P. cavity to the oscillator injecting the 1µs pulse and the ’signal’ antenna to couple the electric field of the resonator to the detector circuit, which can be tuned externally and independently critically coupling can be obtained. Both antennas are ’L’ shaped and can have different geometrical dimensions (varying from 0.5 to 2 cm) depending on the frequency range they are used for. Various antenna separations from the mirror surface were tested. During the measurements for the Cn N chain molecules two equal antennas specified for 9.5 GHz were used and both were separated from the mirror surface by 0.08 mm to yield an optimum coupling. An important parameter of the experiment is the Q factor (quality factor) which is defined as Q = ωW/P (4.15) where ω = 2πν is the angular frequency of the radiation, W is the total energy stored in the cavity and P is the power dissipation, i.e. the energy loss per time −dW/dt. In a quasi-optical treatment of a resonator there are mainly two reasons for the power inside the cell to dissipate. (1) The power can be dissipated by diffracting on the mirrors. This problem can be made arbitrarily small by increasing the radius of the mirrors a. The a2 Fresnel number F ≡ λR with F&1 is a measure of the Fresnel diffraction and making √ F=1, i.e. a = Rλ insures a good Q because the mirror captures more than 95% of the wave amplitude at any point [6]. The spectrometer has mirrors with radii a of 36 cm which cause a cut-off frequency at ∼5 GHz, see Fig. 4.13. A general discussion of this problem is given in the classical work of Boyd & Gordon [21] and Kogelnik & Li [105] 21 . The upper limit at about 40 GHz is not due to diffraction but is imposed by cut-offs in amplifiers and the high-band PIN switch. (2) Ohmic losses in the metallic mirrors. Based on the assumption that the dissipations are only due to ohmic losses a theoretical Qth can be calculated with d Qth = (4.16) 2δ where δ is the skin depth, i.e. the distance in the conductor at which the amplitude of an electromagnetic wave has decreased to 1/e of its value at the surface. In this experiment the mirrors are made of aluminum which has a δ of 8.5·10−5 cm at 10 GHz thus resulting in a Qth ≈ 4.1·105 . The resonant cavity mode has a Lorentzian line shape and the full width ∆νc at half height is given by νc νc ∆νc = , or Q = (4.17) Q ∆νc The decay time constant τc of the energy W in the cavity can be obtained by considering that the energy W decays at a rate proportional to W and 1/τc defined as the proportionality constant W · (1/τc ) = − 21 dW (4.15) 2πνc W ≡P = , thus following dt Q τc = Q (4.17) 1 = . (4.18) 2πν 2π∆νc Storm et al. [171] describe a way to circumvent this problem by using a cylindrical resonator to operate in TE01q modes instead of TEM00q as used in the spectrometer described here. 4.3 The Fourier Transform Microwave Spectrometer 65 power ν Q= ∆ν Q2 > Q1 Lorentzian line shape signal 1 MHz band width Fabry−Perot Q1 Q2 polarization ν 400 kHz νo νpol Figure 4.12: Cavity mode Lorentzian line shape. The quality factor Q is a measure of the full width ∆ν at a given frequency ν of the cavity mode. Different Q’s result in different power values for a given frequency. If the mirrors are adjusted for a cavity mode center frequency νo of 10 GHz the bandwidth of the Fabry-Perot is ∼ 1 MHz. The polarization pulse has a 400 kHz offset from the expected molecule emission frequency. In this experiment high-quality factors (Q up to ∼ 104 ) have been obtained over much of the centimeter-wave band. The highest unloaded Q0 22 now achieved is 2·105 , which is close to the above calculated theoretical limit of about 4·105 . When the cavity is critically coupled, the loaded QL 23 is about 105 . For a cavity with Q=104 at νc = 10 GHz the decay time τc is 0.16 µs which corresponds to a cavity band width ∆νc of 1 MHz, see Fig. 4.12. This means that in praxis the maximum frequency region which can be recorded at a time is of the order <1 MHz, i.e. 500-600 kHz. After this frequency region is examined a new frequency can be set and a computer moves the mirrors appropriately to match the desired center frequency with the cavity mode. Line intensities I are sensitive to the cavity Q. In [6], Eq.(40), a relation between the emitted electric field in the cavity and the Q factor is given in form of a proportionality E(r, t) ∼ Q, so that I ∼ Q2 . However the measurable Qef f , in which also influences of the detection circuit are considered, is no smooth function of the frequency ν and can vary rapidly between adjacent frequency regions so that measured intensities can differ strongly from theoretically predicted. During this thesis many new Q-values cor22 23 The unloaded Q, Q0 , accounts for power dissipation in ohmic losses The sum of all dissipative elements defines the loaded Q, QL , i.e diffraction, antenna coupling, etc.. 66 4 Experimental Setup Figure 4.13: Measured Qef f and calculated Qth for the Fabry-Perot cavity. Set 1 - 4 are effective Q values including effects of the electrical detection circuits measured by Sam Palmer (April 1998). Set 5 was measured during this work. Between 6 and 13 GHz the Q-value can change rapidly by stepping to an nearby frequency which is best seen in Set 5. Therefore a comparison between absolute intensities of even close lying emission lines is nearly impossible. ’s=1.3mm’ and ’d=203µm’ means that the signal antenna was separated by 1.3mm from the mirror surface and the drive antenna by 203µm.’L=1.9cm’ is the length of the antenna and the frequency in brackets is the optimum coupling frequency for that antenna. The theoretical Q values were calculated by √ Sam Palmer using the formulae Q = 2πd νc α1 , with αref lectiv = 4π λδ = 1.43 · 10−4 νGHz , a2 δ the skin depth and αdif f ractiv = 16π 2 F e−4πF = 23.79νGHz e−1.893νGHz , F = λd the Fresnel number in cgs units. 4.3 The Fourier Transform Microwave Spectrometer 67 responding to the used antenna set and frequencies of interest were determined and used as reference values, see Fig. 4.13. The essential cavity optics and first-stage amplifier are cooled with liquid nitrogen, thereby reducing the noise equivalent system temperature from 800 K to about 200 K, which is nearly a factor of four better sensitivity 24 . Above 10 GHz diffraction losses in the open resonator are negligible, and roughly two-thirds (110 K) of the receiver noise is from the cold amplifier and one-third from the 77 K mirrors. Below 10 GHz diffraction from the open resonator contributes significantly to the cavity Q, and the system temperature rises to about 400 K. The cooling is done separately for each mirror by continuously flowing liquid nitrogen through a copper coil soldered to a copper disk making good thermal contact with the mirror’s back surface. Thermal isolation of the mirrors is achieved by suspension on epoxy strips. Condensation of gas from the supersonic jet does not appreciably degrade the reflectivity of the cold mirrors. The supersonic molecular beam is oriented parallel, rather than perpendicular as described in [6], to the Fabry-Perot axis and can be removed via a gate valve assembly to service the discharge nozzle even when the spectrometer is operated at 77 K. The transition lines measured by this spectrometer appear Doppler shifted, see Fig. 4.14. The Doppler separation is a measure of the relative speed ±vmolecule of the incoming molecules with regard to the back and forward traveling microwaves and is symmetric with respect to the rest frequency νo of the molecules. ∆ν vmolecule 2 = , νo c Hz/2 Here 0.107M · c = 840.45 ms = speed of the molecule. see Fig. 4.14. The line separation 19083.2M Hz and thus the speed of the molecules can differ due to their size or charge, i.e ions are faster than neutral species. 24 ”The sensitivity of the present liquid-nitrogen-cooled FTM spectrometer is far from the fundamental limits. Liquid helium cooling of the optics and the first stage of receiver amplification might improve the sensitivity by nearly an order of magnitude. ” [179] 68 4 Experimental Setup time domain E t /µ s ν o = 19083.2139 MHz signal / µV 11 kHz Doppler ∆ν = 107 kHz frequency ( 400 kHz − ν o ) Figure 4.14: Time and Frequency Domains. The computer receives a FID signal in the time domain and Fourier transforms it into the frequency domain. The molecules emission frequency νo is the mean value of the two doppler shifted lines. These originate from the el.mag. waves which travel back and forth in the FP-cavity relative to the molecule’s flight direction. The Fourier transform FID is displayed as a frequency offset from the pump frequency νpolarization . The plotted line is an unidentified line measured during a 13 CCCN survey. 5 Linear CnN, Cyanide Radicals “ The important thing in science is not so much to obtain new facts as to discover new ways of thinking about them. ” Sir William Bragg Cn N are open shell molecules, radicals in the technical sense of the term, and therefore paramagnetic. All members of the even numbered radicals are found in the 2 Π electronic ground state through theory [144] and experiment (e.g. this work) whereas in the case of odd membered cyanide radicals there is an expected change from 2 Σ to 2 Π ground state for increasing n [99, 144]. Dipole moments have not been measured for any of the Cn N chains discussed here, see Tab. 5.1 and 5.2 but instead ab initio calculations have been performed [144, 20]. In each series the dipole moment is found to increase steadily with chain length; given that the radicals have the same ground state. Other carbon chains like Cn H and HCn N which both are readily detected in space are of great importance regarding the Cn N chains. The first ones, Cn H (or Cn CH, with CH playing the role of N), is isoelectronic to the Cn N chains and in this work is often used for references and comparison of molecular constants and properties like magnetic hyperfine constants or electronic energy separations, see Appendix A. The second ones, the cyanopolyynes HCn N are interesting because mainly of two reasons. (1.) The role of Cn N chains in the production and depletion of interstellar cyanopolyynes, see Chapter 8. (2.) On a theoretical point of view the odd Cn N chains can be thought of daughter molecules of acetylenic HCn N by removal of the terminal hydrogen atom which results in an π 4 σ 1 2 Σ electronic state for the cyanide radicals with the radical electron localized at the terminal carbon [144]. An electron transfer from a π orbital into the half-filled σ orbital will cause a shift in the charge distribution and generates a nonpolar or in praxis nearly nonpolar 2 Π state radical which is close in energy, (see Fig. 5.1 top and bottom right). This causes a principle difference between Cn N chains with odd n or even n. The electron configuration of even Cn N chains can be explained by adding the electrons of an extra C atom to an odd numbered Cn N chain after the π → σ electron transfer, (see Fig. 5.1 bottom). 70 5 Linear Cn N, Cyanide Radicals In the following part a brief introduction to the Cn N molecules known so far is given. The odd numbered Cn N (n=1,3,5,...,13) members: CN Emissions of CN in the violet band first measured in the 1920’s were later (1937 by Adel [2], 1941 by McKellar [126]) rediscovered in the tails of comets (see Herzberg [85]) and in absorption towards the star ξ Ophiuchi (in 1941 by Adams [1]). The exploration of the µm-wave spectrum of the CN radical was widely promoted by radio astronomical observations. Here CN was first measured by Jefferts et al. [94] in 1970 towards Orion A and W51 and Penzias et al.[149] and Turner & Gammon [184] were able to determine the hyperfine (hf) structure of CN in the vibrational ground state to high precision. At this point it is worthwhile mentioning that the rotational transition N= 1 → 0 at 113 GHz is also of cosmological interest as is shortly described now. In 1964 Penzias and Wilson [148] discovered during their work at the Bell Laboratories the existence of the isotropic cosmic background radiation (CBR) which was soon interpreted by Dicke et al. [44] as the relic of an early stage of our universe when the electromagnetic field decoupled from matter. Interpretation of the intensity ratios of the R(0) violet transition X 2 Σ+ → B 2 Σ+ at 3874.6 Å and the R(1) transition by Thaddeus et al.[177] resulted in a rotational temperature [147] that came close to the today excepted value for the CBR of 2.725 K [71] and led to the conclusion that CN is thermally excited by the background microwave radiation; an interpretation which is in sharp contrast to the one given in 1950 by Herzberg [83] where it is said that ”the rotational temperature of 2.3◦ K [...] of course only [has] a very restricted meaning”. This example shows nicely that the relevancy of observations can change significantly with the progress of science. Also vibrational excited states have been investigated. Important work was done by Skatrud et al. [167], who determined the Dunham coefficients, the spindoubling, and the hf parameter for the v = 0,1,2,3 states. Johnson et al. [96] have studied the v = 2 state and Ito et al.[93] gave an analysis of the excited states up to v = 10. A good overview of this work for states from v = 0 to 7 and additional measurements up to the THz region is given in the PhD thesis of E. Klisch [104]. Isotopic measurements of 13 CN have been done by Bogey et al. between 1984-86 and included ground state [16] and excited vibrational states ν ≤ 9 [17] as well as a study of the isotopic dependence of the molecular constants. 12 C15 N was first measured in space with the KOSMA 3m telescope towards Orion A by Saleck et al. [160] and little later spectroscopically characterized by Saleck et al. [161] in 1994. Thomson et al. [180] determined the dipole moment to be 1.45 Debye. 71 Figure 5.1: Top: Schematic diagram of electron configuration of CN π 4 σ 1 (X 2 Σ). Bottom: CCN σ 2 π 1 (X 2 Π) electron configuration. On the right: CN changing from π 4 σ 1 (Σ) into a π 2 σ 3 (Π) state after electron transfer. (see Engelke [52] and Rajendra et al. [158]). 72 5 Linear Cn N, Cyanide Radicals An important result of the isotopic measurements is that CN as many other diatomic molecules like O2 , CO, and NO deviate significantly from the BornOppenheimer approximation. C3 N The linear carbon chain radical C3 N (cyanoethynyl) was first detected in the gas phase with a radio telescope by Guélin and Thaddeus [78, 79] in 1977 in the molecular envelope of the carbon-rich star IRC+10◦ 216. The identification was based on two emission line doublets at 89 and 99 GHz and the B, D and |γ| molecular constants have been determined. The detection was later confirmed by observations in many other interstellar sources (e.g. [56, 10]) and also the ρtype doubling and magnetic hyperfine constants were determined by astronomical observations [76]. The laboratory detection of C3 N had to wait until 1983 when Gottlieb at al. [66] made first measurements in the mm-range. Mikami et al. [130] studied the ν5 vibrationally excited bending mode state of C3 N by performing measurements at 208 to 278 GHz which resulted in the determination of several parameters including the l-type doubling parameter q. Sadlej et al.[159] studied the A2 Π, B 2 Π, C 2 Σ and X 2 Σ ground state of C3 N by ab initio calculations and obtained a dipole moment of 3.0 D. First measurements on the isotopic species were done by McCarthy at al. [121] in the mm-range (near 250 GHz). These efforts resulted in a detailed knowledge of the geometry of C3 N and also a preliminary estimation of the hyperfine coupling constant bF (13 C) for the 13 CCCN and C13 CCN radicals. In the same work Botschwina presented calculations of the vibration frequencies of C3 N. Magnetic hyperfine structure (hfs) is a sensitive probe of the electronic structure of open shell molecules and in this work a detailed measurement of the magnetic hyperfine structure of 13 CCCN, C13 CCN, CC13 CN, and CCC15 N is presented with the aim of getting information on the distribution of the unpaired electron spin density along the carbon chain, (see Chapter 7). The electric quadrupole hyperfine structure of the 13 C-isotopes was also determined. During a line survey of the C-star envelope IRC+10◦ 216 carried out by Cernicharo et al. [34] 10 lines of low intensity were assigned to 3 isotopic species of C3 N. Two out of four lines assigned to 13 CCCN are not in agreement with the laboratory work presented here, (see Chapter 8). C5 N In 1991 Pauzat et al. [144] noted that for odd numbered Cn N chains with n>3 a change from 2 Σ to 2 Π electronic ground state is expected and C5 N was predicted to be in a 2 Π ground state. In detail this means, that if the energy of the molecular orbital 7σ is higher than that of 3π, C5 N should be in a 2 Σ ground state. A 2 Π state would apply if the energy of the 3π orbital is higher than that of the 7σ. High level coupled cluster calculations by Botschwina [19] predicted the ground state to have a 2 Σ symmetry. In 1997 Kasai et al. [99] published the first detection of C5 N by Fourier transform microwave spectroscopy between 5 and 17 GHz and the ground state was determined to have 2 Σ symmetry. Shortly after Guelin et al. [77] 73 detected C5 N in the dark cloud TMC-1, (see Chapter 8). Electronic transitions of 2 Π ← X 2 Σ of C5 N in a neon matrix have been measured in the visible by Grutter et al. [72] between 427 and 471 nm. C7 N The absorption bands of the 2 Π ← X 2 Σ electronic transition of C7 N have been measured in the visible in a neon matrix by Grutter et al. [72]. They find a ” great deal of similarity [...] in the vibrational pattern of the C5 N and C7 N” and conclude that C7 N is probably also in a 2 Σ electronic ground state. An ab initio calculation by Botschwina et al. [20] using restricted Hartree-Fock and partially restricted open-shell coupled cluster theory yield an opposite result for which C7 N has not 2 Σ but instead a 2 Π electronic ground state. During this work a measurement campaign was started for linear C7 N but ended without any result. Our conclusion is that C7 N might be in a 2 Π ground state because if C7 N has a 2 Σ ground state with a dipole moment of roughly -3.1 D to -4.2 D but is less abundant than C5 N it still should have been detected during our search. The dipole moment of the radical in a 2 Π state is calculated to be between -0.5 and 1.0 D [20] and could have caused the emission of the radicals to fall under the detection limit of the FTMW spectrometer, (see Chapter 7.3). C9 N - C13 N Grutter et al.[72] published 6K neon matrices measurements of C5 N up to C13 N radicals in the visible and near IR between 470 and 830 nm. For C9 N, C11 N and C13 N the 2 Π ← X 2 Π states were studied with several vibrational bands including the C≡C, C≡N, and C-C stretching modes. The even numbered Cn N (n=2,4,6,8) members: C2 N CCN was first observed in the laboratory by Merer and Travis [128] in 1965. The experiment was done with a flash photolysis of diazoacetonitrile and they observed an absorption spectrum of X̃ 2 Π → Ã2 ∆, X̃ 2 Π → B̃ 2 Σ− , and X̃ 2 Π → C̃ 2 Σ+ in the region 4710-3480 Å. They deduced that CCN is linear in the ground as well as in the excited states. Furthermore they found a Renner-Teller (R-T) interaction for both X̃ 2 Π and Ã2 ∆ states. Kakimoto et al. [98] reinvestigated the (000)(000) band of the Ã2 ∆-X̃ 2 Π system through Doppler-limited dye laser excitation spectroscopy and extended this study to the (010)-(010) and (020)-(020) sequence band [100] in 1984. The experimental R-T splitting measured by Oliphant et al. [143] is ∼ 144 cm−1 . Theoretical work including high level ab initio calculations was performed by Mebel and Kaiser [127], Pd and Chandra [145] and Martin et al. [116] which showed that CCN(2 Π) does not seem to be the most stable isomer but CNC(2 Πg ) by 3.1 kJ/mol [116]; cyclic C2 N(2 A1 ) lies 50.7 kJ/mol above CNC, see Fig. 5.2; Merer and Travis also observed CNC [129]. Measurement of laser-induced fluorescence spectra yielding ground state vibrational frequencies at 1923 cm−1 (ν1 ), 325 cm−1 (ν2 ), and 1051 cm−1 (ν3 ) were done by Brazier et al. [23] and Oliphant et al. 74 5 Linear Cn N, Cyanide Radicals Figure 5.2: Calculated geometries of C2 N. The bond distances are given in Ångstroms, bond angles in degrees for the B3LYP/6-311G∗∗ and CCSD(T)/TZ2P (bold numbers) calculations [116]. The point groups and electronic states are also given. [143]. Measurements of vibration-rotation spectra were published by Suzuki et al. [173] including hyperfine transitions in the electronic A state by a microwaveoptical double resonance technique. Feher et al.[55] measured the ν1 vibrationrotation transitions by infrared absorption spectroscopy using a tunable diode laser. The first pure rotational spectrum of CCN in the 2 Π electronic ground state was measured by Ohshima and Endo [142] in 1995 by Fourier transform spectroscopy at 35 GHz. They determined the magnetic hyperfine constants a− , b, d, eQq0 and eQq2 and refined the known rotational, centrifugal distortion, and fine structure constants to high precision. Pd et al. [145] and Ohshima et al. [142] gave a nice overview of the field including comprehensive reference tables. The ν2 bending fundamental of the CCN radical in its X̃ 2 Πr state was studied by Allen et al. [3] using infrared laser magnetic resonance spectroscopy. C4 N Ding at al. [46] suggested that C4 N might be the first member among the Cn N radicals with even n having stable low-lying cyclic isomers with the three-membered ring isomer NC-cCCC only 2.8 kcal/mol higher in energy than the linear CCCCN but with a larger dipole moment of 0.62 D, (see Fig. 5.3). In this work the first observation of the pure rotational spectrum of linear C4 N in the 2 Π electronic ground state (as mentioned in [4]) is presented. Only the Ω = 21 spin sub levels were examined. Transitions of C4 N with ∆ F = 0,-1 from J= 32 − 12 up to J= 92 − 72 were measured and yield precise hyperfine coupling constants due to the nitrogen nucleus. C6 N The C6 N radical is the largest member among the Cn N chains for which the pure rotational spectrum was observed. In this work the molecule was found to have a ground state with 2 Π symmetry as expected and the Ω = 21 spin sub levels were − 19 were examined. Here transitions with ∆ F = -1 from J= 92 − 72 to J= 21 2 2 75 Figure 5.3: Theoretical geometries of C4 N. The bond distances are given in Ångstroms and the angles in degrees for the B3LYP/6-311G(d) and QCISD/6-311G(d) (in bold) calculations, see [46]. The point groups are written under each structure. measured. C8 N The effort undertaken during this thesis to measure C8 N remained unrewarded. It was expected to find a radical in a 2 Π ground state with a dipole moment larger than 0.3 D. The rotational constant B was estimated to be ∼410 MHz. In contrast to the odd-membered Cn N radicals (see Tab. 5.1) where already CN, C3 N, and C5 N were detected in space none of the even-numbered molecules (see Tab. 5.2) were observed so far beside the laboratory. There is no doubt that the detailed information on the hyperfine structure of C4 N and C6 N is indispensable for their future astronomical detection but the main obstacle is seen in their small ground state dipole moments of 0.14 - 0.33 D [144]. 76 5 Linear Cn N, Cyanide Radicals molecule ground state CN X2 Σ+ C3 N X2 Σ C5 N X2 Σ C7 N probably X2 Π Table 5.1: Cn N, n odd. dipole exp B value first moment (predicted) detection [Debye] [MHz] [astro-source] -1.45 56693.47 astro,visible astro,radio lab,mm-µm lab,THz -2.2 4947.62 astro,mm lab -3.4 1403.08 lab astro 0.8 (583±1) not yet detected year References 1941 1970 1977 1995 1977 1983 1997 1998 Adams [1] Jefferts [95] Dixon [47] Klisch [103] Guélin [78] Gottlieb [66] Endo [99] Guélin [77] Botschwina [20] Dipole moments are taken from Pauzat et al. [144] molecule C2 N C4 N C6 N C8 N Table 5.2: Cn N, n even. ground dipole exp B value first state moment (predicted) detection [Debye] [MHz] 2 XΠ 0.40 11938.58 lab,UV-vis lab,µm X2 Π 0.14 2422.70 lab,µm 2 XΠ 0.31 873.11 lab,µm X2 Π (410±2)∗ not yet detected year References 1965 Merer [128] 1995 Ohshima [142] 2000 this work 2001 this work Dipole moments are taken from Pauzat et al. [144] and Pd & Chandra [145] in the case of C2 N. * estimated by the author 6 Theoretical Considerations “Die Überlegung ist lustig und bestechend; aber ob der Herrgott nicht darüber lacht und mich an der Nase herumgeführt hat, das kann ich nicht wissen...” Albert Einstein, letter to C.Habicht, 1905 Although the history of spectroscopy dates back to the mid 19th century 1 the emission and absorption of light by atoms and molecules can only be correctly understood in a modern quantum mechanical treatment of the phenomenon. Today the basic theory of the interaction of particles like atoms, molecules, and radicals with electro-magnetic waves is on wide parts well understood but still not complete. Radicals for instance are not only difficult to produce in the laboratory but show also special features in their spectra which make the work with these molecules especially challenging. It is not simply the rotational and vibrational motions which have to be considered but also the electronic and magnetic behavior of these molecules which because of their open shell structure can be very complicate. To demonstrate this a hypothetical energy level diagram of an radical in a 2 Π ground state is given in Fig. 6.1 including an inlet showing a spectra of several hyperfine components of one rotation transition. Transitions between these energy levels are governed by selection rules and in many cases spin statistics determine the intensity of the measurable lines. As has been seen in Chapter 4 the used FTMW spectrometer has a very high frequency resolution and an extremely high sensitivity so that in many cases even the faintest lines can be detected. For the analysis of a measured spectra all of these lines have to be assigned and labeled by quantum numbers corresponding to the energy levels. The focus of this chapter is to introduce an appropriate Hamiltonian to describe the rotational spectra of linear radicals in a 2 Σ or 2 Π electronic ground state; in particular for the Cn N radicals. The basic references are the standard text books of Bernath, “Spectra of Atoms and Molecules” [12]; Edmonds, “Angular Momentum in Quantum Mechanics” [51]; Gordy & Cook, “Microwave Molecular Spectra” [63]; Herzberg, “Molecular Spectra and Molecular 1 In 1859 Kirchoff and Bunsen discovered that each element has its own characteristic spectrum. 78 6 Theoretical Considerations spectrum Π Π 7B 3/2 J=5/2 N=3 f e intensity 2 5B J=3/2 6B f e S=+1/2 frequency N=2 4B 2B Aso N=1 N=0 e J=5/2 f 5B S=−1/2 e J=3/2 J=1/2 F=5/2 F=1/2 F=3/2 F=3/2 F=5/2 F=1/2 F=3/2 f e f 3B Rotation Π F=1/2 F=3/2 F=1/2 1/2 Fine structure Hund(b) Λ −doubling Hyperfine structure Hund(a) Figure 6.1: Hypothetical energy level diagram of a 2 Π radical including rotational, fine structure, Λ-doubling and hyperfine structure effects. The inlet on the upper right shows an exemplary spectrum of one rotation transition with several hyperfine components. Structure I & II” [83, 82], “The Spectra and Structure of Simple Free Radicals” [84]; Townes & Schawlow, “Microwave Spectroscopy” [181]. Important contributions came also from papers and works of Brown & Schubert [27], Frosch & Foley [57], Kawaguchi et al. [102], and Klisch [104]. The frequency ν of emission lines correspond to the difference of two energy levels ∆E = E 00 − E 0 = hν. If the energy levels are known all transitions can be computed and frequency predictions can be made. Using the stationary Schrödinger equation the energies are the eigenvalues of the Hamilton operator Ĥ ĤΨ = EΨ (6.1) which reflects the properties of the molecule and also consists of all interactions that may occur. Electrons are generally much faster than nuclei and their motion can very often be treated separately. The generalization of this concept leads to the Born-Oppenheimer 6.1 Pure Rotation of Linear Molecules 79 approximation where the wavefunction Ψ of a molecule can be separated in a product of sub-functions each representing a certain motion or property Ψ ' Ψel Ψvib Ψrot Ψns (6.2) with Ψel the electron wavefunction, Ψvib and Ψrot the vibrational and rotational wavefunction respectively, and Ψns the nuclear spin wavefunction. Each of the sub-functions is dependent on a certain set of quantum numbers but not necessarily on all, e.g. the electron wave function Ψel (nLS) only depends on n the principal, L the orbital, and S the electron spin quantum number but not on v the vibrational, J the rotational and M the magnetic quantum number. Ψ forms a set of basis functions in which the Hamiltonian can be written as a sum of sub-Hamiltonians, e.g. Ĥ = Ĥel + Ĥvib + Ĥrot + Ĥns (6.3) If only pure rotational transitions are considered the expectation value hĤel + Ĥvib i = Eev can be treated as a constant. In the case of radicals the electrons and nuclei have several possibilities to interact with each other or with the overall motion of the molecule which leads to extra terms in the Hamiltonian. But instead of refining the Hamiltonian 6.3 it is more convenient to re-express the Hamiltonian in the form Ĥ = Ĥrot + Ĥf + ĤΛ,l + Ĥhf s (6.4) that is to focus on the structure of the energy levels and to restrict oneself to the relevant terms for the MW- and mm-wave measurements; here Ĥrot is the rotation term, Ĥf the fine structure term caused by the electron spin and orbital angular momentum, ĤΛ,l representing the Λ- and l-type doubling effect caused by rotation-electron orbit interaction and effects of the bending vibration respectively, and Ĥhf s the hyperfine structure term mainly caused by the nuclear spins and electric quadrupole interactions, see Tab. 6.1. In the subsequent sections the effect of each term of the Hamiltonian 6.4 will be explained in more detail with special focus on the radicals relevant for this thesis. 6.1 Pure Rotation of Linear Molecules As a first approach and if no vibration and electronic effects are considered a linear molecule can be seen as a rigid rotor where the distances between the nuclei are fixed. The energies resulting out of an end-over-end rotation can be obtained using Ĥrot = L̂2 h2 = 2 J(J + 1) 2I 8π I (6.5) with L̂ the angular momentum operator and J the eigenvalues of L̂ obeying L̂2 = ~J(J + 1). 2 The Hamiltonian can be simplified by introducing the rotational constant BSI = 8πh2 I 80 6 Theoretical Considerations Table 6.1: Selection of important interactions and their constants parameter interaction (IA) Hamilton term / quantum numbers / energy term / relations fine structure γ ρ-type doubling ĤN S = γ N̂ · Ŝ (1) electr. spin - rotation IA N, S γD distortion constant of γ λ electr. spin -electr. spin IA ĤSS = 23 λ (3 ŝ2z − Ŝ 2 ) S Aso electr.spin - orbit IA ĤSO = Aso (L̂ · Ŝ) (1) S, L Aef f Aef f = ASo + γ Λ and l-type p (= qΛ ) Λ-type doubling e.g. EΛ = ±p 12 (J + 12 ) (2) doubling rotation - electr. orbit IA for 2 Π1/2 in a pure Hund’s R, L case (a) q l-type doubling El = ±qi 14 (vi + 1)J(J + 1) bending vibration for Π states (2) , vi , l pef f pef f = p + 2q pDef f distortion constant of pef f magnetic hf a, a+ , a− nuclear spin - orbit IA ĤIL = a ΛIˆ · k̂ (1) , I, L a+ = a + 12 (b + c) a− = a − 12 (b + c) b, bF Fermi-contact ĤF = b Iˆ · Ŝ, bF =b+ 3c (1) I, S c, t elect.dipole- nucl.dipole IA Ĥ = c (Iˆ · k̂)(Ŝ · k̂) (1) I, S t = 3c d hyperfine Λ-doubling Ĥd = d 12 (exp(2iφ)I− S− + (only for 2 π1/2 -states 6= 0) exp(−2iφ)I+ S+ ) (3) I, S CI nuclear spin - rotation IA ĤCI = CI Iˆ · N̂ I, N electr. eQqo 1. and 2. order elec.quadr.IA (3) ĤeQqo = eQqo (3Iz2 − I 2 )× (4I(2I − 1))−1 (3) ĤeQq2 = eQq2 /(4I(2I − 1)) ×(exp(2iφ)I−2 + exp(−2iφ)I+2 ) 1 from Townes & Schawlow [181]; 2 energy term calculated by perturbation theory, the + sign yields the upper Λ or l -doublet level and the - sign the lower level, see Gordy & Cook [63], energy terms are expressed in MHz, i.e. normalized by 1/h · 10−6 (h Planck constant); 3 Kawaguchi et al. [102] quadrupole eQq2 I 6.1 Pure Rotation of Linear Molecules 81 [Joule] to give Ĥrot = BSI J(J + 1). It is however customary to express B in MHz (or cm−1 ) rather than in the SI-units B= h × 10−6 8π 2 I [MHz] . (6.6) For the rest of this thesis all molecular constants and energies are expressed in MHz (if not otherwise indicated). The eigenvalues of Eq. 6.5 can thus be written as hĤrot i = Erot = BJ(J + 1) (6.7) so that the energy is expressed in MHz. The molecular spectrum consists of transitions between these energies, i.e. νE 0 →E 00 = E 0 − E 00 = 2BJ 0 (6.8) with E 0 the upper state energy and E 00 the lower state energy. However, a molecule is not strictly a rigid rotor and centrifugal forces have to be considered so that the rotational energy is of the form Erot = BJ(J + 1) − D[J(J + 1)]2 + H[J(J + 1)]3 + ... (6.9) with D and H the first and second order centrifugal distortion coefficient respectively. Only the centrifugal distortion constant D could be determined for the molecules discussed in this thesis because the FTMW spectrometer has a upper frequency limit of 40 GHz which in these cases correspond to low rotational excitations. 6.1.1 Selection Rules Not every transition between rotational energy levels is allowed. For electronic dipole transitions with dipole moment µ̂ the transition moment Z µ = Ψ0 (J 0 M 0 )∗ µ̂Ψ00 (J 00 M 00 )dτ (6.10) determines the intensity and if µ = 0 the transition is “forbidden” or if µ 6= 0 “allowed” with I ∼ |µ2 | 2 . The transition moment µ depends on the wave functions and thus on the quantum numbers which determine Ψ0 and Ψ00 in order that µ 6= 0 or not. Quantum numbers that yield allowed transitions are determined by so called “selection rules”. There are two types of selection rules: a) the rigorous electric dipole selection rule and b) the approximate electric dipole selection rule. In a) the selection rules are independent of the degree of approximation introduced in the wave function and in b) they are not. Disregarding the effect of nuclear spin the resulting selection rule for linear molecules are that only transitions with ∆J = J 0 − J 00 = ±1 2 (6.11) Double primes (Ψ”) indicate the lower state and single prime (Ψ’) stands for the upper state . 82 6 Theoretical Considerations are allowed 3 . This means that measurable transitions should have frequencies of the form νJ 0 ←J 00 = 2BJ 0 − DJ 03 . (6.12) 6.2 Fine Structure Any atom or molecule with unpaired electrons reveals some sort of sub-structure to the rotational structure. These effects are mainly due to the electronic spin-orbital, spinspin, and the spin-rotational interaction which are called fine structure interactions. and can be expressed in the Hamilton term Ĥf of Eq.6.4 as Ĥf = ĤN S + ĤSS + ĤSO . (6.13) The first term on the right hand side of Eq. 6.13 takes account of the electron spinmolecular rotation interaction HˆN S = γ N̂ · Ŝ (6.14) which is always present for radicals with an electronic multiplicity of 2S + 1 ≥ 2. This effect can be explained by the electrons participating in the rotation of the molecule and thus inducing a weak magnetic field which then can interact with the electron spin moment. In general this induced magnetic field is rather weak and the splitting of the ĤN S energy levels small. On the other hand the spin-rotation interaction is the only contribution to the fine structure for molecules in a 2 Σ state, e.g. C3 N and its isotopomers. If a molecule has more than one unpaired electron, e.g. O2 or C4 , the dominating term of the fine structure is 2 (6.15) HˆSS = λ(3ŝ2z − Ŝ 2 ). 3 This is somehow the corresponding quantum mechanical expression to the classic magnetostatical equations of two interacting dipoles. If a molecule also has an electronic orbital angular momentum Λ 6= 0 an usually strong interaction between the spin and the orbital angular momentum occurs: HˆSO = ASO Ŝ · L̂ (6.16) Since a molecular electron does not move in a spherically symmetric field (as it is possible in an atom), torques are exerted on it by the field which in general causes the angular momentum not to be constant. In a diatomic or linear molecule the fields are symmetric about the molecular axis and no torque is exerted on a molecular electron about the 3 It is also ∆M = 0, ±1. Furthermore, there is a (+/-) parity rule found by Laport in 1924. Here each energy level is labeled according to its inversion symmetry property with a plus or minus. The transition rules are: + ↔ −, + = +, − = −. Irrespective of the presence or absence of nuclear spin, this rule is strictly valid for dipole radiation. 6.2 Fine Structure 83 internuclear axis. The component Λ of the electronic angular momentum L is therefore constant in this direction (other important angular momenta are listed in Tab. 6.2). In a Hund’s case (a) the spin S is coupled to the molecular axis and Ω is a good quantum number so that the spin-orbit term is separated by the amount ASO in fine structure blocks with Ω = |Λ + S| and Ω = |Λ − S|. The labeling of these blocks are the Ω itself, e.g. a 2 Π state has a Ω = 1/2 and a Ω = 3/2 block and a 2 ∆ state can be split into a Ω = 3/2 and a Ω = 5/2 block. ASO is a constant of a pure electronic interaction and usually much larger than the rotational constant B so that transitions with different Ω usually do not occur in the microwave region due to the big energy gap between the Ωblocks. This was also the case in this work where only transitions of C4 N and C6 N were measured with ∆Ω = 0. The typical energy ladder BN (N + 1) of rotational transitions known from the Hund’s case (b) can not be found in a Hund’s case (a). Instead the energy levels are widely separated in the two fine structure blocks which both reveal a ladder structure due to the rotation. Only if molecules are considered with Λ ≥ 1 and 2S + 1 ≥ 3 (e.g. in 3 Π states) all three terms of Eq. 6.13 are simultaneously needed. For the C3 N isotopes in a 2 Σ state only ĤN S had to be considered and also for the 2 Π molecules C4 N and C6 N only one fine structure term (ĤSO ) is needed. The latter is due to a strong correlation between ASO and γ in a strong Hund’s case (a) so that instead of using ASO and γ separately a new constant Aef f = ASO + γ can be introduced to describe the structure of these radicals. 6.2.1 Hund’s Coupling Cases a) and b) Radicals have one or more unpaired electrons and therefore a total spin S 6=0. In addition the molecules can have a non vanishing electronic orbital angular momentum and nuclear spins. The rotational energy expression (Eq. 6.9) is not strictly valid for such molecules and a systematic way of including the extra angular momenta has to be found. In 1926 Hund [91] introduced a method to deal with the coupling of angular momenta which is basically a classification of ideal cases which are very often closely approximated by real molecules. The Hund’s coupling cases are often correlated to the electronic ground states of the radicals which follow the notation 2S+1 |Λ| (6.17) with S the total electron spin and 2S + 1 the electronic multiplicity, here Λ is the electronic orbital momentum in units of ~, e.g. 0,1,2,3,..., which is represented by the characters Σ, Π, ∆, Φ. The used angular momenta are summarized in table Tab. 6.2. The Cn N (n odd) radicals have a 2 Σ electronic ground state, i.e. Λ = 0 and S = 1/2, and the even membered radicals have a 2 Π ground state, i.e. Λ = 1 and S = 1/2. Hund’s case (a): This case applies if Λ 6= 0 and the electronic spin-orbit (LS) coupling is assumed to be strong, while the coupling of the rotation of the nuclei with the electronic motion is very weak, i.e. the magnetic field generated by the rotation is small. Ω is then a good quantum number even if R 6= 0, see Fig. 6.2. Ω and the rotational angular 84 R L Λ S Σ N J Ω I F 6 Theoretical Considerations Table 6.2: Angular momenta and their projections angular momentum of the end-over-end rotation of the molecule resultant electronic orbital angular momentum component (projection) of L along the molecular axis Λ = |ML | = 0, 1, 2, ..., L resultant electronic spin projection of S along the molecular axis, Σ = S, S − 1, ..., −S = R + L angular momentum without elect. spin total angular momentum without nuclear spin = |Λ + Σ| resultant angular momentum along molecular axis nuclear spin total angular momentum (with nuclear spin) . R J J J N S case a) Λ Σ L S Ω Λ R N S L case b) case b), molecule in Σ state . Figure 6.2: Vector diagram of Hund’s coupling cases a) and b). On the left: Hund’s case (a) with only the total angular momentum J fixed in space. The nutation of the internuclear axis about J is indicated by the blue ellipse; the precession of L and S about the figure axis are assumed to be much faster (green dash-dotted ellipses). In the middle: Hund’s case(b) in the general case Λ 6= 0. Also here only the total angular momentum J is fixed in space. The precision of N and S about J (green ellipse) is much slower than the nutation of the internuclear axis about N (dash-dotted blue line). On the right: Hund’s case(b) with Λ = 0. 6.2 Fine Structure 85 momentum R result in an angular momentum J. A criterium for a molecule to be in a Hund’s case (a) is (Gordy & Cook [63]) 2JB |ΛASO | . (6.18) Hund’s case (b): This case can be best explained in the most prominent case were Λ = 0 but S 6= 0, i.e. Σ states almost always can be expressed in a Hund’s case (b). Because there is no orbital field the spin moment can not couple to the internuclear axis. The strongest influence to the spin moment comes from the weak magnetic field generated by the end-over-end rotation of the molecule which causes S to couple with N (which is the same as R because Λ=0). In some cases it can happen that light molecules, e.g. OH, in a high rotational state J generate a large magnetic field by rotation that is strong enough that the electronic spin S is rather resolved in direction of J than to the internuclear axis even for Λ 6= 0. In general the Hund’s case (b) is defined as the ideal case where S is coupled to N to form J as seen in Fig. 6.2. Examples of the more complicate cases of the coupling of one or two additional nuclear spins are discussed in Chapter 7. Selection Rules The distinction in a Hund’s case (a) or (b) involves the usage of different “good” quantum numbers, i.e. J and Ω or N in which the selection rules can be expressed. In the Hund’s case (a) the quantum number Σ is a good quantum number and the selection rule ∆Σ = 0 (6.19) is valid. This rule holds for combinations of the sets of rotational levels of various multiplet components. The usage of Ω is more common and thus the selection rules for linear molecules 4 are ∆J = ±1 and ∆Ω = 0, ±1 (6.20) In the microwave region there are mainly transitions with ∆Ω = 0 to observe (because of the higher energies involved in the other types of transitions) but this is not due to a forbiddance of this transitions. In the Hund’s case (b) the selection rule is ∆J = ±1 and ∆N = ±1 . (6.21) In the more complicate case where the molecule has atoms with nuclear spin, e.g. I1 and I2 , the total angular momentum is not J but F = J + I1 + I2 4 and considering only cases of pure rotation (6.22) 86 6 Theoretical Considerations Now the analog of Eq. 6.11 holds rigorously for dipole radiation: ∆F = 0, ±1 with F = 0 = F = 0 (6.23) Because the interaction with J and the nuclear spin is usually very weak the rule 6.11 for the quantum number J is still very strong even though not rigorous. The selection rules 6.19 and 6.21 for N and S do not hold in intermediate coupling cases between (a) and (b). A nicely summarized presentation of this topic including the selection rules for magnetic dipole and electric quadrupole radiation is given in Herzberg [84]. 6.2.2 Λ-type Doubling, and l-type Doubling Λ-type doubling. For radicals with an orbital angular momentum Λ 6= 0 the fine structure energy terms split in doublet states labeled with e and f with reference to the parity of the wave function Ψ. At this stage of development these doublet states would be exactly degenerate but the influence of the molecular rotation on the electronic orbital momentum lifts the degeneracy. This can be interpreted as a decoupling from the electron orbital angular momentum from the internuclear axis 5 . The bigger the Λ-type splitting the stronger the coupling of L towards R so that the molecule takes an intermediate state between Hund’s case (a) and Hund’s case (d) which would be the extreme case of a L-R vector coupling scheme. The Λ-doubling splitting is significantly smaller than the fine structure splitting. Perturbation theoretical considerations are leading to the following proportionalities (Landau-Lifschitz [109]) ∆EΛ ∼ (me /M )2Λ (6.24) with me the mass of the electron and M of the molecule. The ratio of masses (me /M ) 1 and with Λ > 1 the chances of a measurable splitting of the Λ terms decrease rapidly so that radicals with a ∆ or Φ electronic ground state are of nearly no interest concerning this symmetry effect. On the other hand the Π states can have quite large splittings. In Table 6.3 the approximate energy Λ-type splitting is given for 2 Π states in pure Hund’s case (a) and (b). In the case of C4 N and C6 N it was not the Λ-type doubling constant which was fitted but an effective Λ-type doubling constant pef f , with pef f = p + 2q 6 which also includes q the l-type doubling constant. l-type doubling. Polyatomic linear molecules have vibrational bending modes νi which can be excited. If a degenerate bending mode is excited an additional angular momentum pz = l~ about the internuclear axis with l = vi , vi − 2, vi − 4, ..., −vi has to be taken into account and the rotational energy can be written as Erot = B[J(J + 1) − l2 ], 5 6 In a classical picture this would be interpreted as an effect of the Coriolis force. p is used here as the Λ-doubling constant irrespective of the applied Hund’s case. (6.25) 6.3 Hyperfine Structure 87 Table 6.3: Theoretical Λ-type doubling for 2 Π state, Gordy&Cook [63]. ∆E Hund’s splitting of levels approx. theor case (approx.) coupling constant (a) Ω= 1 2 b 2 1 )(J 4 pa = 3 ) 2 4ASO B νe b a B ASO 2 3 = νe8B p (J − + p = 2p ASO (b) pN (N + 1) νe represents the transition frequency between the ground level and the lowest Σ state; all constants are in frequency units. Ω= 3 2 pa (J + 12 ) with J the total angular momentum including l, so that J = |l|, |l| + 1, |l| + 2, ... . The Coriolis coupling force which is proportional to v × ω between the vibrational motion v and the angular rotation motion ω in it’s orthogonal plane, lifts the ±l degeneracy of Eq. 6.25 and a doublet splitting of the rotational lines is produced. For |l| = 1, i.e. a vibrational Π state, the l-type splitting is approximately 1 ∆E|l|=1 = q (vi + 1)J(J + 1) 2 (6.26) where vi is the vibrational quantum number of the ith degenerate bending mode and q the vibration-rotational coupling constant. In the first excited state the energy splitting simplifies to ∆E|l|=1 = qJ(J + 1) (6.27) In most cases q can be approximated to be (Gordy & Cook [63]) q ≈ 2.6 Be2 νbend (6.28) with Be the equilibrium rotational constant and νbend the degenerate bending frequency. 6.3 Hyperfine Structure If a molecule has one or more nuclear spins the spectra will expose a further splitting of the energy levels due to the interaction of the nuclear spins with the other angular momenta of the molecule. This is most commonly called the hyperfine structure because it is usually much smaller than the above discussed fine structure. However, in some cases the substructure induced by the nuclear magnetic moment can be of the same order or even larger than the electric fine structure interactions, e.g. the Fermi-contact interaction in the case of 13 CCCN is larger than the electronic spin-rotation interaction. 88 6 Theoretical Considerations For the hyperfine transitions different kinds of electromagnetic radiation can be involved. A photon always carries an angular momentum of l~ with l=1,2,3,... which corresponds to a classical radiation field of a 2l -pole. During an emission or an absorption process not only the overall angular momentum but also the parity has to be conserved in the photon-molecule system. The transformation properties (~r) → (−~r) of electric and magnetic multi-pole radiation are not the same, i.e electric multi-pole radiation has a (−1)l parity whereas magnetic multi-pole radiation has a (−1)l+1 parity. A transition between two molecular states with the parity π1 and π2 can only occur if π1 = (−1)l π2 , π1 = (−1)l+1 π2 , for El-radiation for Ml-radiation (6.29) is fulfilled. For example: In the Hund’s case (a) the selection rule is ∆J = ±1 and with a nuclear spin ∆F = 0, ±1 so that only magnetic dipole (l = 1) and electric quadrupole (l = 2) are allowed and electric dipole and magnetic quadrupole transitions are forbidden [104]. A molecule containing an atom with a quadrupole moment, e.g. 14 N in C4 N or 13 C in CC13 CN, will always have a hyperfine structure in their spectrum. If the molecule has an open shell structure with S 6= 0 the magnetic dipole (hfs) transitions usually dominate upon the electric quadrupole transitions in terms of the line intensities. The Hamilton term Ĥhf s of Eq. 6.4 can be written as Ĥhf s = Ĥmag,hf s + ĤQ (6.30) with Ĥmag,hf s the magnetic interaction term and ĤQ the electric quadrupole term. 6.3.1 Magnetic Hyperfine Structure The magnetic hyperfine structure Hamiltonian as it is used in this thesis can be written as Ĥmag,hf s = ĤIL + ĤF + ĤIS + Ĥd (6.31) with ĤIL the nuclear spin- electronic orbit interaction term, ĤF the Fermi contact term, ĤIS the nuclear spin - electronic spin interaction (or short: dipole-dipole interaction), and Ĥd the hyperfine Λ-doubling term. In general the Hamiltonian for an interaction between a magnetic dipole and a magnetic field is of the form Ĥ = µ̂ · Ĥm where Ĥm is the magnetic field and µ̂ the dipole, i.e. µI = gI µn I the magnetic spin moment with gI the dimensionless gyromagnetic ratio (g-factor), µn the nuclear magneton, and I the nuclear spin. The magnetic field Hm can be caused by the electronic orbital or spin angular momentum and depending on the direction of the I,L, and S the Hamilton equation can be written as ĤIL = aΛIˆ · k̂ ĤF = bIˆ · Ŝ ĤIS = c(Iˆ · k̂)(Ŝ · k̂) (6.32) (6.33) (6.34) 6.3 Hyperfine Structure 89 with k̂ a unit vector along the molecular axis. The expressions Eq. 6.32 - 6.34 apply accurately only when Λ is a “good” quantum number and holds for Hund’s case (a) and (b). The introduced constants a, b, and c are 2µB µI 1 a = (6.35) I r3 U 2µB µI 8π 2 3 cos2 θ − 1 b = Ψ (0) − (6.36) I 3 2r3 U 3µB µI 3 cos2 θ − 1 c = (6.37) I r3 U with θ the angle between the molecular axis and r the radius from the nucleus to the electron. Ψ(0) is the electron wavefunction at the interacting nucleus. (. . .)U denotes that the mean value is taken over the unpaired electron. The Eq. 6.32 - 6.34 apply to each electron in the molecule. It is evident that for electrons in the inner shells and for those who are paired the terms nearly cancel out each other so that in most cases only the unpaired electrons have to be considered. Nuclear spin - electronic orbit interaction. The quantity a refers only to electrons with an orbital angular momentum. According to the Biot-Savart law the unpaired electrons with L 6= 0 generate a magnetic field at the nucleus which interacts with the nucleus magnetic moment. Fermi contact interaction. It is common to use the “Fermi contact” constant bF instead of b which is defined as bF = b + c/3 with 2µB µI 8π 2 bF = Ψ (0) . (6.38) I 3 U (Ψ2 (0))U is the probability to find the unpaired electron at the nucleus which for an electron in an p atomic orbit is negligibly small but for a s-type orbit can be quite large. Hence, whenever there is a appreciable amount of s character to the wavefunction of an unpaired electron the magnetic hyperfine interaction which is proportional to Ψ2 (0), may be expected to dominate, i.e. bF > a. Dipole-Dipole interaction. Because of the (cos2 θ)U angular dependence this interaction cancels for a spherical electron density distribution, e.g. s-type orbitals. Hyperfine Λ-doubling. The magnetic hyperfine interactions discussed so far are all identical for the two energy levels of a Λ-doublet. For a Π state a certain type of electron spin-nuclear spin interaction results in a different structure for the Λ doublet. Qualitatively this can be explained best in a Hund’s case (b). The electron wave function has a eiφ ± e−iφ angular dependence with φ the angle of rotation about the internuclear axis. The probability distribution Ψ2 of the electrons is therfore proportional to sin2 φ and cos2 φ. For the lower Λ-doublet state with a sin2 φ distribution, the field of the electron at the nucleus is parallel to I and 90 6 Theoretical Considerations rotation of molecule rotation of molecule R R s φ s I s I molecular axis molecular axis lower Λ− doublet state s upper Λ− doublet state Figure 6.3: Unpaired electron distribution for a 2 Π state in Hund’s case (b) for the two Λ-doublet states. for the upper state it is directed oppositely to I, see Fig. 6.3. This causes the spin-spin interaction energy to be different for the two Λ states, i.e. the hf Λ-doubling is parity depending. For a 2 Π state in a Hund’s case (a) the splitting due to the hyperfine Λ-doubling is ∆Ed = ± d(J + 12 ) ˆ ˆ I ·J 2J(J + 1) (6.39) for the 2 Π1/2 state and ∆E = 0 for the 2 Π3/2 state. The upper sign in Eq. 6.39 applies to the upper Λ-doublet state. d is defined as 3µB µI d= I sin2 θ r3 (6.40) U In all molecules with a nuclear spin there is an interaction between the nuclear magnetic moment and the rotation of the molecule which can be calculated by Van Vlecktransformations of higher order terms and usually results in a small splitting which is expressed in the constant CI . This type of interaction was not considered in this work for the Cn N radicals. Useful references for the definition of the magnetic interaction constants are Frosch & Foley [57] and Steimle et al. [170]. 6.3 Hyperfine Structure 91 6.3.2 The Electric Quadrupole Interaction There can also be a hyperfine structure due to electric charge distribution in the nucleus. If the nucleus is not assumed to be a point charge the charge distribution has to be considered which may be in motion and produces magnetic fields which gives the nucleus an angular momentum in quantities of I~ (I is an integer or half integer). If V is the electrostatic potential produced at the nuclear center of mass by all electronic charges in the atom the electrostatic energy of a nuclear charge ∆q = ρ(x, y, z)∆x∆y∆z, with ρ the nuclear charge density, is ∆W (x, y, z) = ∆q(x, y, z)V (x, y, z). Quantum mechanical considerations on the multi-pole expansion of the electrostatic potential V V0 + x ∂V0 ∂ 2 V0 ∂V0 ∂V0 1 2 ∂ 2 V0 1 2 ∂ 2 V0 1 2 ∂ 2 V0 +y +z + x + y + z + xy + ... (6.41) ∂x ∂y ∂z 2 ∂x2 2 ∂y 2 2 ∂z 2 ∂x∂y reveal that the nucleus normally has no inherent dipole moment and that also all terms involving odd powers of the coordinates will be zero ([181],p.132) thus leaving only a term which is independed of the nuclear size or shape (i.e. ZeV , with Z= atomic number , e=proton charge) and a term associated with the quadrupole moment of the nucleus. The energy due to the electric quadrupole moment is than 1 WQ = − Q : ∇E 6 which is the inner product between the quadrupole moment dyadic Z Q = (3~r ⊗ ~r − r2 1) ρ d(3)~r (6.42) (6.43) and the gradient of the electric field due to the electrons. Using a coordinate system with z in the direction of the nuclear spin all non diagonal terms of Q vanish and the entire quadrupole moment can be expressed in terms of one constant Z 1 (3z 2 − r2 ) dx dy dz (6.44) Q= e called “the” nuclear quadrupole moment [181]. For nuclei with a spherical charge distribution Q is zero and the quadrupole moment can thus be seen as a measure of the deviation from a spherical shape, i.e. if the nuclear charge distribution ρ is somewhat elongated along z then Q is positive; if it is flattened along the nuclear axis, Q is negative. All isotopes with I=0 or 1/2 have a quadrupole moment equal zero because of their spherical symmetry. From Eq. 6.42 it is clear that the energy also depends on the gradient of the electric field at the nucleus. In a linear molecule the charge distribution is symmetric around the molecular axis but varies along these axis so that the gradient of the electric field depends on the position within the molecule. Nitrogen 14 N with I=1/2 has a Q of +0.02 · 1024 cm2 [181] but the electric quadrupole energy from C2 N is different from that of CNC because of the different ∇E experienced by the nitrogen nucleus. In a quantum mechanical treatment the Hamiltonian can be build out of Eq. 6.42 with Q and 92 6 Theoretical Considerations e − I 2 1 has the same angular d Because 3 (II + II) ∇E replaced by operators Q̂ and ∇E. 2 dependence with respect to nuclear orientation as 3~r ⊗ ~r − r2 1, Q̂ can be expressed as 3 eQ 2 e −I 1 Q̂ = (II + II) (6.45) I(2I − 1) 2 d can be shown to be and ∇E 7 d= ∇E q 3 2 f (JJ + JJ) − J 1 J(2J − 1) 2 (6.46) 2 with q = (JJ| ∂∂zV2 |JJ) depending on J. The energies of the Hamiltonian ĤQ = − 16 Q̂ : d are therefore of the form ∇E E = −eQq f (J, Ω, I, F ) (6.47) where eQq is called the quadrupole coupling constant and f (J, Ω, I, F ) a function 8 which involves the coupling of the angular momenta. In the tensor notation the quadrupole interaction can be written as HQ = X eQq2 61/2 eQq0 2 ˆ ˆ 2 ˆ ˆ T0 (I, I) − e2iqΦ T2q (I, I) 4I(2I − 1) 4I(2I − 1) q=±1 (6.48) ˆ I) ˆ a standard with eQq0 and eQq2 the two electric quadrupole parameters and T 2 (I, 2nd-rank spherical tensor with components expressed in the molecule-fixed axis system, see [27]. It is 2 sin θ 2 (6.49) eQq2 = −3e Q r3 T with the index T indicating that the mean value is taken over the total (i.e. the paired and unpaired) electrons, see Ohshima et al. [141]. This equation (6.49) together with Eq. 6.37 enables an estimation of the non-axial distribution of the wavefunctions. The matrix elements of HˆQ in a Hund’s case (a) are discussed in 6.4.1 in more detail. 7 8 Eq. 6.46 is only valid if J is a “good” quantum number. Under certain assumptions f is the Casimir function but in general Eq. 6.48 has to be applied. 6.4 Matrix Representation of the Hamiltonian 93 6.4 Matrix Representation of the Hamiltonian For non-singlet states even the rotational Hamiltonian Ĥrot = BR2 can be significantly more complicate than in the singlet case because only with the usage of “good” quantum numbers is the calculation of the energy expression meaningful. If an electronic orbital angular momentum and an spin is present the Hamiltonian can be written as Ĥrot = B(Jˆ − L̂ − Ŝ)2 . (6.50) On way to proceed is to re-express Eq. 6.50 as Ĥrot = B(Jˆ − Jˆz2 )2 + B(Ŝ − Ŝz2 )2 + B(L̂ − L̂2z )2 −B(Jˆ+ L̂− + Jˆ− L̂+ )2 − B(Jˆ+ Ŝ − + Jˆ− Ŝ + )2 + B(L̂+ Ŝ − + L̂− Ŝ + )2 (6.51) using Jˆ+ , Jˆ− as lowering (+) and raising (-) operators respectively and L̂+ , L̂− , Ŝ + , Ŝ − as raising (+) and lowering (-) operators in the common sense of definition. With Eq. 6.51 it is already evident that in some cases, e.g. 2 Π states, a mixing of the states is necessary to build the correct Hamiltonian. As an example the basis set of a 2 Π state is given here as |2 Π3/2 i, |2 Π1/2 i, |2 Π−1/2 i, |2 Π−3/2 i, (6.52) which can be written as e/f parity basis functions: |2 Π3/2 i ± |2 Π−3/2 i √ , | Π3/2 e/f i = 2 |2 Π1/2 i ± |2 Π−1/2 i √ | Π1/2 e/f i = 2 2 2 (6.53) with (+) referring to the e parity and (-) referring to the f parity. The rotational Hamiltonian can now be expressed as |2 Π3/2 e/f i |2 Π1/2 e/f i B[(J + 1/2)2 − 1] −B[(J + 1/2)2 − 1]1/2 Ĥ = 2 1/2 −B[(J + 1/2) − 1] (6.54) 2 B[(J + 1/2) + 1] Every Hamiltonian term in Eq. 6.4 can be written in such a matrix form. For 2 Σ states the off-diagonal terms in the matrix are very often zero and a matrix representation is not necessary in such a case but for 2 Π states the situation is completely different. 6.4.1 The Matrix Representation of the 2 Π-Radicals The fit of the 2 Π radicals C4 N and C6 N proved to be not as easy as expected and it is therefore useful to examine the 2 Π matrix elements in more detail. The matrix elements of the rotation, spin-orbit, Λ-doubling and magnetic hyperfine interaction of 2 Π states in the Hund’s case (a) are calculated in Brown et al. [26]. The most influential term of 94 6 Theoretical Considerations Table 6.4: Matrix with Spin-Orbit, Rotation and Λ-doubling h2 Π± 3/2 JIF |...... h2 Π± 1/2 JIF |...... |2 Π± 3/2 JIF i |2 Π± 1/2 JIF i 1 1 2 2 {A + AD [(J + 2 ) − 1]} − 21 γD Z +B[(J + 21 )2 − 1] − DZ(Z + ∓ 12 qD (J − 12 )(J + 12 )(J + 32 ) −{B − 12 γ − 12 γD (Z + 2) − 2D(J + 12 )2 ∓ 12 q(J + 12 ) ∓ 14 (pD + 2qD )(J + 12 ) ∓ 12 qD (J + 12 )3 }Z 1/2 1) − 12 {A + AD (Z + 2)} −γ − γD 12 (3Z + 4) +B[(J + 12 )2 + 1] − D(Z 2 + 5Z + 4) ∓ 12 (p + 2q)(J + 12 ) ∓ 12 (pD + 2qD )(J + 12 )(Z + 2) ∓ + 12 qD (J + 12 )Z (Hermitian) with Z(J)= (J- 12 )(J+ 32 ) = (J+ 12 )2 -1 Upper and lower sign choice refer to e and f levels respectively. the Hamiltonian is one containing the rotational, spin-orbit, and Λ-doubling interaction and the matrix representation is given in Tab. 6.4. As an example and because of the importance in the fit of the C4 N radical the matrix elements of the electric quadrupole interaction is given here. HQ of Eq. 6.48 in the Hund’s case (a) can be expressed using 3j- and 6j-Symbols 9 : 0 0 −1 F J I (−) 2 I J0 0 J 2 J 0 1/2 J 0 −Ω ×[(2J + 1)(2J + 1)] δΛ0 Λ δΩ0 Ω eQq0 (−) −Ω 0 Ω 0 X 0 0 J 2 J + δΛ0 ,Λ∓2 (6)1/2 eQq2 (−)J −Ω (6.55) 0 −Ω −q Ω 1 hηΛ SΣJ Ω IF |HQ |ηΛSΣJΩIF i = 4 0 I 2 I −I 0 I J+I+F q=±2 This formula was directly used in the fit program to analyze the C4 N and C6 N spectra. A disadvantage of this representation is that it is not very intuitive and it is therefore 2 ± desirable to express HQ in a matrix of the form h2 Π± Ω0 JIF |HQ | ΠΩ JIF i as it can be seen in Tab. 6.5. The derivation is given in the Appendix B. In short: 2 Π states have Ω or Ω0 values of 3/2 or 1/2 so that Ω’=Ω ± 1. For the energy the ∆J = 0 elements were 9 see Brown & Schubert [27], Eq. 2 6.4 Matrix Representation of the Hamiltonian 95 Table 6.5: Matrix with electr. hf interaction |2 Π± 3/2 JIF i h2 Π± 3/2 JIF |...... h2 Π± 1/2 JIF |...... with K(F ) = eQq0 27 2 K(F )[ 4 | 2 Π± 1/2 JIF i − J(J + 1)] 1 1 3 1/2 2 2 ± eQq 4 K(F )[(J − 4 )(J + 2 )(J + 2 )] eQq0 3 2 K(F )[ 4 (Hermitian) − J(J + 1)] 3R(F )[R(F )+1]−4J(J+1)I(I+1) I(2I−1)(2J+3)(2J+2)(2J)(2J−1) R(F ) = F (F + 1) − J(J + 1) − I(I + 1) Matrix elements in non-parity conserving basis derived by Tom C. Killian and Guido Fuchs from Brown & Schubert, [27]. calculated using the basis functions |2 Π± |Ω| , Ji = 10 |Λ, Σ, J, Ωi ± | − Λ, −Σ, J, −Ωi √ 2 (6.56) with ± referring to the e/f parity respectively and following the convention of Brown et al. [25]. It was necessary to include the off-diagonal term eQq2 in the analysis of the C4 N spectrum to achieve a good fit, see Chapter 7.1. Another problem appeared when fitting the C6 N spectrum. In this case a correlation between the magnetic hfs constant a− and eQq0 seemed to jeopardize the analysis. The matrix for the magnetic hfs is given in Tab. 6.6 in terms of the constants a, bF and c as it appeared in Brown et al. [26]. If only transitions between one of the Π3/2 or Π1/2 states are measured it is advantageous to use a transformation (a,bF ,c) → (a+ ,a− ,b) which separates the magnetic hf constants for the Ω=3/2 and Ω=1/2 states: a a+ a + 12 (b + c) a + 12 bF + 31 c bF → a− = a − 1 (b + c) = a − 1 bF − 1 c 2 2 3 c b b bF − 3c The new matrix is given in Tab. 6.7 and it can be seen that a+ is the interaction constant for Π3/2 states only and a− and d for Π1/2 states only. b can be determined using Π3/2 or Π1/2 transitions either. The analysis of the C6 N spectrum was based on this Hamiltonian and includes only the a− constant, see Chapter 7.2. Some other than the here mentioned approach from Brown & Schubert [27, 26, 24] to the theoretical study of the 2 Π Hamiltonian are made by Frosch & Foley [57], Kawaguchi et al. [102], Davies et al. [42]. 10 The matrix elements are in a non parity conserving basis. 96 6 Theoretical Considerations Table 6.6: Matrix with magnetic hyperfine interaction |2 Π± 3/2 JIF i h2 Π± 3/2 JIF |...... h2 Π± 1/2 JIF |...... R(F ) [3a 2J(J+1) 2 + 34 bF + 12 c] (Hermitian) |2 Π± 1/2 JIF i R(F ) [(J− 12 )(J+ 32 )]1/2 1 [ 2 bF 2J(J+1) R(F ) [1a 2J(J+1) 2 − 14 bF − 16 c ∓ 12 d(J + 12 )] with R(F ) = F (F + 1) − J(J + 1) − I(I + 1) Table 6.7: Transformed matrix with magnetic hyperfine interaction |2 Π± 3/2 JIF i h2 Π± 3/2 JIF |...... R(F ) 3 a 2J(J+1) 2 + h2 Π± 1/2 JIF |...... (Hermitian) | 2 Π± 1/2 JIF i R(F ) − 16 c] [(J− 12 )(J+ 32 )]1/2 1 b 2J(J+1) 2 R(F ) 1 [a 2J(J+1) 2 − ∓ d(J + 12 )] with R(F ) = F (F + 1) − J(J + 1) − I(I + 1) 7 Measurements and Analysis “ The great tragedy of Science - the slaying of a beautiful hypothesis by an ugly fact. ” Thomas H. Huxley (1825 - 1895) The production of the C3 N isotopomers and the C4 N, C6 N production has been already described in Chapter 4.1. In short: For speed and convenience, all of the C3 N isotopic measurements were made with enriched samples; with the 13 C-HCCH sample. Line intensities were typically 2-3 times stronger than those of the same lines observed in natural abundance. When the 13 C-methylcyanide sample was employed instead, lines of CC13 CN were three times more intense than those observed with 13 C-acetylene 1 . In this chapter the results of the measurements are presented and an interpretation of the data is given. CCC15 N is discussed first because of its relative simple spectrum which is in fairly good agreement with the theoretical predictions, i.e. the b(15 N) and c(15 N) values, derived from the CCC14 N measurements done by Gottlieb et al. [66] but which had not been resolved in the CCC15 N measurements of McCarthy et al. [121] so that only the constants B, D and γ were known. CCC15 N as well as the 13 C mono-substituted C3 N isotopomers have a 2 Σ ground state but in the case of the 13 C isotopic species of C3 N two nuclear spins have to be considered which causes an extra splitting of the energy levels compared to that of the CCC15 N species. Only the B, D, γ and bF (13 C) values were known for the 13 CCCN, C13 CCN, and CC13 CN species from millimeter-wave measurements done by McCarthy et al. [121]. In the same work the c(13 C) values were not determinable and set to zero. Also the bF (14 N ) and 1 Short summary: 9 GHz antenna set at room temperature, A/D delay around 16 or 20 µs, discharge voltage of 1100 - 1250 V, general valve opening time 300 - 480 µs, gas entrance pressure of 2.5 atm, total flow rate of 30 - 32 sccm with 9 sccm Ne, 1.25 sccm 0.18 % HC3 N in Ne and 0.5 sccm 1.5 % H13 CCH (statistical mixture), for CCC15 N it was 0.5 sccm 5% CH3 C15 N, 200 - 5000 shots were integrated for one spectrum. A “short-nozzle” with a “normal” tip was used. 98 7 Measurements and Analysis Figure 7.1: Measured CCC15 N transition with Zeeman splitting c(14 N ) constants were not fitted but assumed to be close to the values for the CCCN radical which were bF (14 N )=-1.20(3)MHz and c(14 N )=2.84(9)MHz. In this work all relevant magnetic hyperfine constants could be fitted plus the electrical quadrupole constant due to the 14 N nucleus. C4 N and C6 N do not have a 2 Σ but a 2 Π electronic ground state and had therefore to be treated separately with a computer program written by John Brown, see Chapter 6.4.1. Both radicals are detected in the laboratory for the first time and no spectroscopic data were available for this molecules sofar. 7.1 The C3N Mono-Substituted Isotopomers 7.1.1 CCC15 N CCC15 N was measured in the 9.5 - 38.4 GHz region corresponding to rotational transitions from J = 1 → 0 up to J = 4 → 3 respectively, see Table 7.1. The analysis 7.1 The C3 N Mono-Substituted Isotopomers 99 Table 7.1: Measured Rotational Transitions of CCC15 N in the X 2 Σ+ State Frequencya O − C b Transition J’ F’ N J F (MHz) (kHz) N’ 1 3/2 2 0 1/2 1 9593.486 −4 2 5/2 3 1 3/2 2 19195.842 0 2 5/2 2 1 3/2 1 19195.863 1 2 3/2 2 1 1/2 1 19213.911 4 2 3/2 1 1 3/2 1 19243.521 0 3 7/2 4 2 5/2 3 28798.215 1 3 7/2 3 2 5/2 2 28798.215 −1 3 5/2 2 2 3/2 1 28816.243 −2 3 5/2 3 2 3/2 2 28816.350 −3 4 9/2 5 3 7/2 4 38400.553 2 4 9/2 4 3 7/2 3 38400.553 4 a Estimated experimental uncertainties (1σ) are 2 kHz. b Calculated frequencies derived from the best fit constants in Tab. 7.2 was done with a standard Hamiltonian for a linear molecule in a 2 Σ electronic ground state similar to Eq. 7.8 and 7.9 for the 13 C isotopic species but with one nuclear spin. The angular momenta coupling scheme was J = N + S and F = J + I(15 N ). 11 lines were sufficient to fit 6 molecular constants but for the final fit also the mm-data from McCarthy et al. [121] were included, see Table 7.2. Radicals are paramagnetic and react on outer magnetic fields like the Earth’s magnetic field as can be seen in the spectrum of CCC15 N Fig. 7.1 where the J 3/2→ 1/2, F 2→1 transition is split due to the Zeeman effect. CCC15 N is the only one of the isotopic species of C3 N which shows a notable Zeeman splitting of 20-30 kHz in its spectrum. Helmholtz coils were mounted around the FTMW cavity to cancel outer magnetic fields but was not used during the measurements for this work. Unlike the cases of CCCN or the 13 C isotopic C3 N species there has to be no electrical quadrupole to be considered for CCC15 N. The bF (14 N) and c(14 N) values from the mm-measurements [121] could be used to estimate the values for the 15 N isotope (see Townes & Schawlow, p.196 [181]): µI bF (14 N ) bF ∼ ⇒ = I bF (15 N ) µI (14 N ) I(14 N ) µI (15 N ) I(15 N ) = µI (14 N ) I(15 N ) = −0.7131 µI (15 N ) I(14 N ) (7.1) so that bF (15 N)est. ≈ 1.68 MHz and c(15 N)est. ≈ -3.98 MHz using µI (14 N)= 0.4036, µI (15 N)= -0.2830, I(14 N)=1, and I(15 N)=1/2. This prediction was good enough to find the actual transition lines within 1-2 MHz. CCC15 N obeys Hund’s case (bβJ ) which can be seen using the formulas in Townes & Schawlow (p.199) were an estimation of the magnitude for the hyperfine splitting for the Hund case (bβJ ) is given. In the case of a molecule in a 2 Σ ground state the formula 100 7 Measurements and Analysis Table 7.2: Molecular Constants of CCC15 N (in MHz). Data reduction was done by using the Pickett-program [153]. a Constant this workb mm-data onlyc recommendedd values B 4801.2277(5) 4801.2264(4) 4801.2267(1) −3 D ×10 0.76(2) 0.7062(3) 0.7064(1) γ −18.218(3) −18.17(4) −18.208(1) γD ×10−3 0.6(2) −0.02(2) ... 15 bF ( N) 1.87(1) 3(17) 1.883(9) c(15 N) −4.26(3) 12.(73) −4.30(3) w-rmse 0.43 0.42 0.97 a Uncertainties (in parentheses) are (1σ) in the last significant digit. b 11 lines were used, see Tab. 7.1. The uncertainties of the lines is estimated to be 2 kHz. c 28 lines from [121] were used. The uncertainties of the lines are estimated to be between 22-86 kHz. d Total fit with all measured 39 lines. e w-rms weighted rms, i.e. rms normalized with uncertainties of measured lines. (8-10)2 can be reduced to (J=N-1/2): ξ WJ=N −1/2 z = − }| { b c + I ·J 2N + 1 (2N − 1)(2N + 1) (7.2) with I · J = 12 (F (F + 1) − J(J + 1) − I(I + 1)). ∆F=1 with J=J’ and N=N’ determines the hyperfine splitting ∆Whf s : ∆Whf s 1 0 0 1 = ξ (F (F + 1) − ...) − (F (F + 1) − ...) 2 2 1 = ξ [(F + 1)(F + 2) − F (F + 1)] 2 = ξ(F + 1) (7.3) With I(15 N)=1/2 and F=(N-1/2)-1/2 we obtain ∆WJ=N −1/2 = ξN = − 2 Townes & Schawlow, p.199 N N ·b+ · c, 2N + 1 (2N − 1)(2N + 1) N 6= 0(!) (7.4) 7.1 The C3 N Mono-Substituted Isotopomers 101 In the case of CCC15 N b=bF -c/3=3.3154 MHz so that 1 (c − b) = −2.54MHz 3 2 1 ∆WJ=N −1/2 (N = 2, F = 1 → 2) = ( c − b) = −1.90MHz 5 3 and ... b N → ∞ : ∆WJ=N −1/2 = − = −1.66MHz 2 ∆WJ=N −1/2 (N = 1, F = 0 → 1) = (7.5) The corresponding formulae for Eq. 7.4 for J=N+1/2 is c b + (F + 1), with F=J-1/2 ∆WJ=N +1/2,hf s = 2N + 1 (2N + 1)(2N + 3) N +1 N +1 = ·b+ ·c (7.6) 2N + 1 (2N + 1)(2N + 3) and thus 1 ∆WJ=N +1/2 (N = 0, F = 0 → 1) = b + c = bF = 1.883MHz 3 2 1 ∆WJ=N +1/2 (N = 1, F = 1 → 2) = (b + c) = 1.637MHz 3 5 and ... b N → ∞ : ∆WJ=N +1/2 = = 1.66MHz 2 (7.7) These estimations are in fairly good agreement with the measurements, i.e. the energy output-file of the fit program, and indicates that we have a nearly pure Hund case (bβJ ). The energy level of CCC15 N is given in Fig. 7.2. 102 7 Measurements and Analysis . CCC 15 N (X 2 Σ ) E GHz splitting x 183 splitting x 13.7 F=1 1.888 MHz F=2 J=3/2 60 N=3 50 γ (N+1/2) F=3 1.619 MHz J=5/2 F=2 . 30 F=0 N=2 . 2.518 MHz F=1 J=1/2 20 27.3 MHz 1.649 MHz J=3/2 F=2 F=1 10 N=1 0 F=0 N=0 Rotation 1.883 MHz J=1/2 b F F=1 Hyperfine−Structure Fine−Structure γ =−18.2 MHz b =1.88 MHz F c=−4.30 MHz . Figure 7.2: Energy level diagram of CCC15 N. The rotational transition N 1 → 0 corresponds to an energy difference of ∼ 9.6 GHz. The fine structure splitting which is due to the spin-rotation interaction is approximately γ(N+1/2). For the N=0, F 0 → 1 transition the hyperfine structure is solely determined by the Fermi contact interaction and the splitting is therefore a direct measure for the bF . The dotted vertical lines represent measured transitions and are sorted in increasing frequency. Only ∆J=∆N transitions which are the strongest are displayed here, i.e. not all measured lines are shown here. 7.1 The C3 N Mono-Substituted Isotopomers Figure 7.3: Measured 7.1.2 13 13 103 CCCN transitions. CCCN, C13 CCN and CC13 CN The hyperfine structures of 13 CCCN, C13 CCN and CC13 CN were analyzed with a standard Hamiltonian for a linear molecule in a 2 Σ electronic state including the 13 C and 14 N nuclear spins [57]: H = Hrot + Hf + Hhf s (7.8) with Hrot = BN2 − DN4 Hf = γN · S + γD (N · S)N2 1 Hhf s = bF I(13 C) · S + c[I(13 C)z Sz − I(13 C) · S] 3 1 +bF I(14 N ) · S + c[I(14 N )z Sz − I(14 N ) · S] 3 (2) 14 +eQq/4To (I( N )) (7.9) where N is the rotational angular momentum of the molecule, S is the electron spin angular momentum, and I(13 C) = 1/2 and I(14 N ) = 1 are the nuclear spins of the respective nuclei. The z axis is taken to lie along the linear carbon chain. Hrot is the 104 7 Measurements and Analysis I(14 C) Hund case (b β s ) 2 nuclear spins 13 I( C) F1 F F2 S I( 14 C) Hund case (b βJ ) 2 nuclear spins I(13C) F F1 N N Λ L S J R quantization axis Λ R quantization axis L Figure 7.4: Hund case bβs and bβJ with 2 nuclear spins. Hund case bβs (left): The subscript β indicates that the nuclear spin is not coupled to the molecular axis but to some other vector. In this case the nuclear spin I(13 C) is coupled with the electron spin S to form F2 . J = N + S does not appear, see Townes & Schawlow p.197 [181], instead a new quantum number F1 is used here. Hund case bβJ (right): In this case, which is expected to be more common, the electron spin couples to the rotation N to give J. Then J and I couple to give F1 . This coupling scheme was exploited in the fit. rotation and centrifugal distortion Hamiltonian and Hf is the fine structure Hamiltonian, i.e. the magnetic interaction between the electronic spin and the molecular rotation. Hhf s is the hyperfine structure Hamiltonian which includes the interaction between the electron spin and the 13 C nucleus as well as the interaction between the electron spin and the 14 N nucleus in terms of the Fermi contact and the electron-nuclear dipole(2) dipole constants. The last term is the electrical quadrupole interaction with To (I) the molecule fixed component of the quadrupole moment tensor. Herb Pickett’s [153] program was used to analyze the data and the coupling scheme J = N + S, F1 = J + I(13 C), F = F1 + I(14 N ), (7.10) was applied to examine the hyperfine structure of the two here measured 13 C isotopic species of C3 N. For 13 CCCN, a more natural choice for the coupling scheme would be F1 = N + I(13 C), F2 = F1 + S, F = F2 + I(N) because the 13 C hyperfine interaction is larger than that of spin-rotation [181] , see Fig. 7.4. However the fitting program can readily handle large off-diagonal terms in the Hamiltonian matrix which occur when the coupling scheme in Eq. 7.10 is used, so, for uniformity, it was adopted for 13 CCCN as well. Some of the observed rotational transitions of 13 CCCN are indicated in the energy level diagram in Fig. 7.5. The data were fitted in two steps. In initial fits to the carbon-13 Fourier transform centimeter-wave data (see Table 7.6, 7.7 and 7.8), the rotational constant B and the centrifugal distortion constant D were constrained to the values previously determined from the millimeter-wave data, and the three nitrogen hyperfine constants (bF , c, and eQq) were constrained to the values for normal CCCN [66]. The three remaining constants, the spin-rotation constant γ, bF (13 C), and c(13 C), were varied to fit the lowest-J transitions, yielding a rms of typically 7.1 The C3 N Mono-Substituted Isotopomers 13 105 2 CCC N (X Σ ) E GHz splitting x 56 arbitrary scaling (less than x 0.1) 28 MHz 50 J=3/2 45.1 MHz J=5/2 40 F=1,3 F=2 F=0 F=2 F=1 F 1 =2 F 1 =1 F 1 =3 F=2 F=1 F=3 F 1 =2 61 MHz N=2 J=1/2 20 966 MHz 27.0 MHz 10 F=2 F=1 F=2 F=0 F=1 F=2 F=3 30 J=3/2 F=3,4 956 MHz F 1 =1 F 1 =2 F 1 =0 F=1 F 1 =1 F=1 F=2 F 1 =1 F=2 F=1 F=0 N=1 0 N=0 Rotation J=1/2 Fine−Structure γ =−18.0 MHz F=0 1.2 MHz 980 MHz F=1 F 1 =0 magnetic Fine−Structure, I( 13 C) b =973.5 MHz c=139.5 MHz F Hypefine Structure Figure 7.5: 13 CCCN energy level scheme. The splitting of the order 900-1000 MHz due to the magnetic interaction of the 13 C is much stronger than those of the 14 N (or 15 N). For the low lying rotation transitions the 13 C Fermi contact and dipole-dipole interaction dominates the energy splitting and is even larger than the spin-rotation interaction (for N<50). The magnetic interaction of the 14 N nuclear spin provides the hyperfine structure in the order of 1 MHz. The dotted vertical lines (on the right) represent measured transitions and are sorted in increasing frequency. Only ∆J=∆N transitions which are the strongest are displayed here. ≤20 kHz; subsequently, B and the three nitrogen constants were varied as well, giving an rms comparable to the 2-5 kHz measurement uncertainty. After the centimeterwave transitions were assigned, global fits including the millimeterwave data [121] were done. The centimeter-wave lines were assigned a frequency uncertainty of 2 kHz, and the millimeter-wave lines uncertainties between 15 kHz and 150 kHz, with 25 kHz for most. Each hyperfine line was thus given a weight of about 100 relative to each of the 5-7 millimeter-wave lines in the range of N=11 to 29. The final hyperfine parameters derived from the global fits are nearly identical to those calculated from the initial fits, and the global rms are comparable to those obtained from the millimeter-wave data alone. As an example Tab. 7.3 shows the molecular constants of 13 CCCN during each of these steps; the analogous tables for C13 CCN and CC13 CN are in Appendix C. In Tab. 7.4 the final fits for 13 CCCN, C13 CCN and CC13 CN are summarized. For comparison, the spectroscopic constants of normal CCCN from [66] are also given. With the 13 C hyperfine constants given in Table 7.4, it is possible to make systematic comparisons between the electronic structure and chemical bonding of C3 N, isoelectronic C4 H, and isovalent CCH. Such comparisons are appropriate because all three chains 106 7 Measurements and Analysis Table 7.3: Molecular Constants of Constanta this workb 13 CCCN (in MHz). mm-data onlyc recommendedd values B 4771.218(1) 4771.2193(1) 4771.2195(2) D ×10−3 0.62(5) 0.6991(1) 0.6993(2) γ −17.96(2) −17.93(2) −17.963(5) γD ×10−3 0.2(7) −0.020(9) ... 13 bF ( C) 980.(4) 973.(3) 973.(2) c(13 C) 140.(5) 108.(1) 139.5(3) 14 bF ( N) −1.17(7) −30.(20) −1.26(6) 14 c( N) 3.1(2) 0.(10) 3.4(1) eQq0 −4.45(4) −6.(52) −4.48(4) w-rmse 0.61 2.720 1.177 a Uncertainties (in parentheses) are (1σ) in the last significant digit. b 11 lines were used, see Tab.7.6. The uncertainties of the lines is estimated to be 2 kHz. c 20 lines from [121] were used. The uncertainties of the lines are estimated to be between 22-86 kHz. d Total fit with all measured 31 lines. e w-rms is the weighted rms, i.e. the rms normalized with uncertainties of measured lines. Table 7.4: Spectroscopic constants of the 13 C isotopic CCCN species. The fit included the MWFT data and the mm-data from McCarthy et al. [121] 13 Constanta CCCNb CCCN C13 CCN CC13 CN B 4947.6207(11) 4771.2195(2) 4920.7095(2) 4929.0640(2) D×103 0.7535(16) 0.6993(2) 0.7453(4) 0.7497(3) γ -18.744(6) -17.963(5) -18.574(5) -18.648(3) γD ×103 -0.006(11) ... ... ... bF (13 C) ... 973(2) 188.6(2) 23.55(2) 13 c( C) ... 139.5(3) 52.9(1) 2.17(3) bF (14 N ) -1.20(3) -1.26(6) -1.234(6) -1.182(8) 14 c( N ) 2.84(9) 3.4(1) 2.82(3) 2.88(2) eQq -4.32(10) -4.48(4) -4.331(9) -4.323(8) w-rmsc ... 1.19 1.16 0.77 number of ... 11+(20) 13+(28) 32+(12) transitions used MWFT+(mm) a Units are MHz. The 1σ uncertainties (in parantheses) are in the units of the last significant digits. The spectroscopic constants were derived from the hyperfine-split centimeter-wave transitions in Tables 7.6, 7.7, 7.8 and the mmwave transitions in [121]. b Ref.[66] c Normalized standard deviation of the fit. 7.1 The C3 N Mono-Substituted Isotopomers 107 . 1 2 C C C N C C C N . . C C C N 4 C C C N . 3 Figure 7.6: Resonance structure of CCCN are σ-bonded radicals with 2 Σ ground states, and because the hyperfine constants are proportional to important expectation values of the valence electron, providing highly specific probes of the molecular wave function. Carbon-13 is particularly useful because it probes the wave function at all the substituted positions along the carbon chain. There are only two non-zero hyperfine parameters for a 2 Σ state: the Fermi-contact term bF and the dipole-dipole term c. The Fermi-contact constant bF is a useful measure to localize the unpaired electron, or more accurate to determine the spin density along the carbon chain in σ bonded radicals like Cn N with odd n (e.g. CN and C3 N) and Cn H with even n (e.g. C2 H and C4 H). This is because only s electrons have non-zero amplitude at (r=0) 3 and in this case the unpaired electron is expected to have significant s character, see Fig. 5.1. The magnetic dipole coupling constant c also provides useful information on the orbital occupation of the unpaired electron, because it is a function of both an angular average and the radial expectation value of 1/r3 . C3 N is isoelectronic to C2 H and C4 H and shows similar behavior in the hyperfine structure. Before comparing these radicals it is worthwhile thinking of a qualitative model for the Fermi contact interaction along the carbon chain which is basically similar for these molecules (e.g. N is replaced by CH in the case of Cn H, n=2,4). A description of the bonding in CCCN requires a superposition of several different electronic structures as it is shown in Fig. 7.6. Structure 1, with the unpaired electron localized on the terminal carbon has the highest stability of all the resonance structures because of it’s four π bonds: two between C(1) and C(2) , two between C(3) and N. Resonance structure 2, with the unpaired electron on C(2) , and structure 4, with the electron on N, are less stable because they only have three π bonds. Structure 3, with the unpaired electron on C(3) , has only two π bonds and is the least stable structure. It is therefore expected that the spin density should be greatest at C(1) , less at C(2) and N, and least at C(3) . Assuming that the electron configuration is identical in all three here relevant isotopomers 13 CCCN, C13 CCN, and CC13 CN, then bF (13 C) for each isotopomer is a measure of the 3 With (r, θ) the spherical coordinates of the electron. 108 7 Measurements and Analysis Table 7.5: bF (13 C) and c(13 C) values 13 C position → CCH CCCN CCCCH bF (13 C)-values / MHz 1 2 3 900.7(6) 161.63(10) — 973.46(2) 188.513(1) 23.555(23) 396.8(6) 57.49(5) -9.54 c(13 C)-values / MHz 1 2 3 142.87(3) 64.07(5) — 139.5(3) 53.0(1) 2.17(3) 89.12(1) -1.91(3) 9.84(8) spin density at carbon C(i) . In Tab. 7.4 the bF -values are listed with 973, 189 and 24 MHz for C(1) through C(3) , which qualitatively is the predicted decrement. Figure 7.7 shows the magnitude of the two hyperfine constants at different positions along the carbon chain for the three radicals. Although bF (13 C) and c(13 C) are nearly the same for 13 CCH and 13 CCCN and for C13 CH and C13 CCN, the same two constants are each smaller by about a factor of two or more at the same substituted positions of CCCCH. The reason for these difference may be the large zero-order mixing between the low-lying 2 Π state and the X 2 Σ+ ground state of C4 H: the 2 Π - X 2 Σ energy separation is calculated to be 3600 cm−1 for CCH [150, 151], 2400±50 cm−1 for C3 N, but only 100±50 cm−1 for C4 H [121] 4 . Owing to strong vibronic coupling between these states, the 2 Σ ground state of C4 H probably possesses significant 2 Π character, unlike the ground states of CCH or C3 N, which are nearly pure 2 Σ. With simple atomic orbitals [170] it is possible to estimate crudely the fractional 2s and 2p character in the 2σ molecular orbital of C3 N, on the assumption the unpaired electron is localized on either of the two carbon atoms furthest from the nitrogen. Within a few percent, this calculation yields the same unpaired electron spin density on the terminal carbon atom (74%) and adjacent carbon atom (26%) for C3 N as that previously derived for CCH [120], and little contribution from the pπ electronic configuration. In contrast, the relative amount of pπ character is estimated by the same calculation [121] to be about 28% for C4 H, a result in good agreement with that of Hoshina et al. [90], who concluded on the basis of LIF measurements that the admixture is about 40%. 4 The bF (13 C) of CC13 CCH is -9.54 MHz (see Tab. 7.4) which can be explained by a spin polarization effect which arises when the paired electrons in the σ orbital are slightly polarized by the electrons 2 in the nearby p orbitals [30] : bF = 2µBI µI h 8π 3 Ψ (0)iU + spin polarization. 7.1 The C3 N Mono-Substituted Isotopomers 109 CCH 1000 CCCCH 800 CCCN F Fermi contact (b , in MHz) 1200 600 400 200 0 0 1 2 3 4 5 0 1 2 3 4 5 Dipole-dipole (c, in MHz) 200 150 100 50 0 carbon atom Figure 7.7: (Top:) bF -values of different isoelectronic carbon chains. Each number n stands for the position of a 13 C-atom within the molecule as seen from the terminal C-atom, i.e. n=2 for the CCCN isotope is C13 CCN. In a pure 2 Σ ground state the unpaired electron would be completely localized at the terminal carbon atom. For a small 2 Π admixture in the 2 Σ ground state the spin density is highest at the terminal C-atom and decreases with increasing n which is reflected by the decreasing bF -values. CCCCH has more 2 Π character in the ground state and hence a smaller |hΨ(0)i|2 which results in smaller bF compared with the more pure X2 Σ species CCH and CCCN. (Bottom:) c-values of isoelectronic carbon chains. 110 7 Measurements and Analysis Table 7.6: Measured Rotational Transitions of 13 CCCN in the X 2 Σ+ State. Frequencya O − C b Transition J’ F’1 F’ N J F1 F (MHz) (kHz) N’ 1 3/2 2 3 0 1/2 1 2 9528.817 1 1 3/2 1 2 0 1/2 0 1 9542.475 −3 2 5/2 3 3 1 3/2 2 2 19073.604 −2 2 5/2 3 4 1 3/2 2 3 19073.886 −1 2 5/2 2 2 1 3/2 1 1 19084.532 −3 2 5/2 2 3 1 3/2 1 2 19084.612 3 2 3/2 2 3 1 1/2 1 2 19085.382 0 3 7/2 4 5 2 5/2 3 4 28617.158 2 3 7/2 3 3 2 5/2 2 3 28626.762 1 3 7/2 3 4 2 5/2 2 3 28626.796 4 3 5/2 3 4 2 3/2 2 3 28627.817 −3 a Estimated experimental uncertainties (1σ) are 2 kHz. b Calculated frequencies derived from the best fit constants in Tab. 7.3 Table 7.7: Measured Rotational Transitions of C13 CCN in the X 2 Σ+ State. Frequencya O − C b Transition J’ F’1 F’ N J F1 F (MHz) (kHz) N’ 1 3/2 2 2 0 1/2 1 1 9829.369 −1 1 3/2 2 3 0 1/2 1 2 9830.396 1 1 3/2 1 1 0 1/2 0 1 9839.603 −1 1 3/2 1 2 0 1/2 0 1 9840.701 −2 2 5/2 3 3 1 3/2 2 2 19672.508 6 2 5/2 3 4 1 3/2 2 3 19672.780 −5 2 5/2 2 3 1 3/2 1 2 19681.164 −7 2 5/2 2 2 1 3/2 1 1 19681.175 3 2 3/2 2 1 1 1/2 1 1 19684.220 1 2 3/2 2 3 1 1/2 1 2 19684.755 2 3 7/2 3 4 2 5/2 2 3 29521.725 3 3 7/2 3 3 2 5/2 2 2 29521.758 1 3 5/2 3 4 2 3/2 2 3 29626.764 −4 a Estimated experimental uncertainties (1σ) are 2 kHz. b Calculated frequencies derived from the best fit constants in Tab. 7.3 7.1 The C3 N Mono-Substituted Isotopomers Table 7.8: Measured Rotational Transitions of CC13 CN in the X 2 Σ+ State. Frequencya O − C b Transition J’ F’1 F’ N J F1 F (MHz) (kHz) N’ 1 3/2 2 2 0 1/2 1 1 9847.713 −3 1 3/2 2 3 0 1/2 1 2 9848.755 2 1 3/2 1 1 0 1/2 0 1 9853.055 5 1 3/2 1 2 0 1/2 0 1 9853.282 8 1 1/2 1 2 0 1/2 1 2 9873.604 2 2 5/2 3 3 1 3/2 2 2 19706.620 3 2 5/2 3 4 1 3/2 2 3 19706.889 2 2 5/2 3 2 1 3/2 2 1 19706.914 3 2 5/2 2 2 1 3/2 1 1 19709.044 −7 2 5/2 2 3 1 3/2 1 2 19709.080 5 2 3/2 2 3 1 1/2 1 2 19723.673 3 2 3/2 1 2 1 1/2 0 1 19726.516 −3 3 7/2 4 4 2 5/2 3 3 29564.843 0 3 7/2 4 3 2 5/2 3 2 29564.907 −3 3 7/2 4 5 2 5/2 3 4 29564.970 2 3 7/2 3 3 2 5/2 2 2 29566.119 −3 3 7/2 3 4 2 5/2 2 3 29566.174 5 3 7/2 3 2 2 5/2 2 2 29568.747 −6 3 5/2 3 3 2 3/2 2 3 29580.140 −2 3 5/2 2 2 2 3/2 1 2 29581.389 −3 3 5/2 3 3 2 3/2 2 2 29582.300 −8 3 5/2 3 2 2 3/2 2 1 29582.480 3 3 5/2 3 4 2 3/2 2 3 29582.569 1 3 5/2 2 2 2 3/2 1 1 29583.673 0 3 5/2 2 3 2 3/2 1 2 29583.913 5 3 5/2 2 1 2 3/2 1 1 29584.491 1 4 9/2 5 5 3 7/2 4 4 39422.912 −5 4 9/2 5 6 3 7/2 4 5 39422.990 1 4 7/2 4 4 3 5/2 3 3 39440.879 −1 4 7/2 4 5 3 5/2 3 4 39440.986 0 4 7/2 3 3 3 5/2 2 2 39441.702 −2 4 7/2 3 4 3 5/2 2 3 39441.785 6 a Estimated experimental uncertainties (1σ) are 5 kHz. b Calculated frequencies derived from the best fit constants in Tab. 7.3 111 7 Measurements and Analysis Theoretical Intensity 112 21,74 21,76 Frequency [GHz] 21,78 21,8 Theoretical Intensity 21,72 0 20 40 60 Frequency [GHz] Figure 7.8: C4 N stick spectrum calculated for Trot =3K. 7.2 C4N and C6N For the first time the linear cyano radicals C4 N and C6 N could be measured in the laboratory. The spectra were observed between 7.2 and 21.7 GHz for C4 N and 7.8 and 18.3 GHz for C6 N, i.e. at least four rotational transitions of each radical fall within the frequency range of the used spectrometer. The measured laboratory frequencies for these molecules are given in Table 7.11 and 7.12, and typical lines are shown in Fig. 7.9 and 7.10. Searches for the rotational spectra of C4 N and C6 N were guided by ab initio calculations of Pauzat et al. [144]. The rotational constants for both molecules were estimated by scaling these ab initio rotational constants by the ratio of the experimental 7.2 C4 N and C6 N 113 B value to that calculated at the same level of theory for C2 N, C3 N, and C5 N. Rotational transitions predicted in this way turned out to be quite accurate - to within 0.25% for C4 N and C6 N. The present identifications are extremely secure: (i) the two new molecules are almost certainly radicals because their rotational transitions are separated in frequency by halfinteger quantum numbers, and their lines exhibit the expected Zeeman effect (i.e., a fairly modest broadening owing to the small magnetic g factor of a 2 Π1/2 state) when a permanent magnet is brought near the molecular beam; (ii) the carriers of the observed lines are nitrogen-bearing molecules because the lines disappear when cyanoacetylene is replaced with diacetylene, and because characteristic hyperfine structure from the nitrogen nucleus was observed in all of the assigned spectra; (iii) impurities from contaminants in the gas samples and van der Waals complexes with the buffer gas can also be ruled out, because the lines were produced with acetylene plus cyanogen as the precursor gas, and when Ar replaced Ne as the buffer gas; and (iv) the identifications of C4 N and C6 N are also supported on spectroscopic grounds by the close agreement of B and D with those estimated by scaling from the ab initio geometries and the Cn H chains of similar size. In addition, the Λ-doubling constant p + 2q in the 2 Π1/2 ladder of C4 N and C6 N can be predicted to within a factor of two by scaling from CCN [142] on the assumption of free precession. The nitrogen hyperfine constants a − (b + c)/2, b, and d also smoothly decrease in magnitude from CCN to C6 N, as one might expect if the unpaired electron is delocalized along the chain; a similar decrease in the hydrogen hyperfine constant a−(b+c)/2 has been observed in the odd-numbered acetylenic chains up to C13 H [64]. Under optimized experimental conditions, the strongest rotational lines of C4 N were approximately 15 times less intense than those of C3 N, but were still observed with a signal to noise of approximately 25 after one minute of integration; the decrement in peak signal strength from C4 N to C6 N was only about a factor of three. An effective Hamiltonian for molecules in a 2 Π electronic state used in the present analysis is expressed as H = Hrot + HSO + HΛ + Hhf s were the terms on the right side of the equation are the rotational energy, the spin-orbit interaction, the Λ-type doubling interaction, and the hyperfine interaction, respectively, see Chapter 6.3.1. The hyperfine structure term includes the magnetic and the electricquadrupole interaction due to the 14 N nucleus. Unlike in the case of C2 N the spinrotation interaction was not needed to result in a good fit. Instead of the pure ASO spin orbit constant Aef f = ASO + γ was used for the fit, see Chapter 6.2. The lowest rotational transitions of both C4 N and C6 N in the ground 2 Π1/2 fine structure ladder are split into six components by Λ-doubling on the scale of 2-5 MHz and then by hyperfine structure from the nitrogen nucleus which is generally smaller by about an order of magnitude (i.e. on the scale of 0.2-0.5 MHz). Rotational transitions from the higher-lying X 2 Π3/2 ladder were not observed, because this level lies at least several tens of Kelvin above the X 2 Π1/2 level and is apparently not appreciably populated in the generally rotationally cold molecular beam (Trot = 0.2 - 3 K). The fits were done 114 7 Measurements and Analysis Figure 7.9: Measured C4 N transition. Shown are the ∆F=∆J transitions at 12094.384 MHz and 12094.480 MHz with F 5/2 → 3/2 and F 3/2 → 1/2 respectively. Figure 7.10: Measured transition of C6 N. Shown are the transitions F 17/2→15/2 and F 15/2→13/2 at 13086.043 MHz and F 13/2→11/2 at 13086.146 MHz. 7.2 C4 N and C6 N 115 14 I( N) Hund case a 1 nuclear spin F J R Λ L Σ quantization axis S Figure 7.11: Hund’s case a with one nuclear spin. In the case of C4 N is L=1, S=1/2, R=0,1,2,... , and I(14 N)=1/2. with the HUNDA program written by J. Brown [27] which is based on the effective Hamiltonian of Brown et al. [26, 24] in the strong Hund’s case (a) with 2JB |ΛASO |, see Fig. 7.11. The spin-orbit constant Aef f = ASO +γ was constrained at 40 cm−1 (∼ 1.2 THz) because no information of the Π3/2 states was available and it was assumed that ASO + γ from C4 N and C6 N is of the same order of magnitude than that from C2 N [142]. At most nine spectroscopic constants: the rotational constant B, centrifugal distortion D, two Λ-doubling constants p + 2q and (p + 2q)D , and five hyperfine constants, the diagonal term a− = a − (b + c)/2, the off-diagonal term b, the parity-dependent term d, the electric quadrupole term eQq0 , and the nonaxially symmetric quadrupole term eQq2 were required to reproduce the 29 measured lines of C4 N and the 42 of C6 N to better than 2 kHz. If eQq2 was constrained to zero in the C4 N fit, the rms increases by more than a factor of five. In Fig. 7.14 the energy level diagram of C4 N is shown. With the spectroscopic constants for C4 N and C6 N as listed in Table 7.9, precise frequencies for the astronomically interesting transitions can be calculated to very high precision. Fig. 7.8 shows a stick spectrum with a calculated intensity distribution corresponding to a rotational temperature of 3 K. For both radicals the strongest lines are found for ∆F = ∆J transitions (mainly with F’=J’+1). In the case of C4 N transitions with ∆F = 0 and ∆F = −1 could be observed but ∆F = 0 transitions of C6 N are very weak and only ∆F = −1 transitions could be measured. That resulted in a strong correlation between the a− and the eQq0 of ∼1 in the fit for the C6 N radical. The eQq0 is nearly constant for all known Cn N chains, with n<6, and was fixed to -4.38 MHz to avoid that correlation. The Cn N, n=2,4,6, are isoelectronic to Cn H, with n=3,5,7 and with the exception of C7 H there are already some data at hand concerning their magnetic and electronic quadrupole hyperfine structure. The Frosch & Foley hyperfine parameters a, bF , c, and d give information about the unpaired and total electron distribution h1/r3 i, hsin2 θ/r3 i, h3 cos2 θ − 1/r3 i, and hψ 2 (0)i. Unfortunately, it is impossible to obtain a complete set 116 7 Measurements and Analysis Table 7.9: Spectroscopic Constants of C4 N and C6 N in the X 2 Π State. Constanta C4 N Aef f 1200000b (40 cm−1 ) B 2422.6963(1) −6 D [× 10 ] 90(3) P ef f 4.5525(8) PDe f f [× 10−3 ] 5.23(2) a− 15.005(1) b(14 N ) 16.2(1) d 22.4254(9) eQq0 -4.389(1) eQq2 5.6(3) transitions ∆ F = 0,-1 no. of lines 29 exp. uncert. 2 kHz rms 1 kHz w-rms 0.614 C6 N 1200000b (40 cm−1 ) 873.11224(6) 11.5(7) 1.939(5) 0.066(2) 8.7(5) 7.4(10) 13.23(1) -4.38b ... ∆ F = -1 33 2 kHz 3.6 kHz 1.803 a Units are in MHz. The 1σ uncertainties (in parentheses) are in the units of the last significant digits. b fixed value of these parameters, owing to lack of data for the a+ = a + (b + c)/2 constant for which Ω3/2 transitions are needed. Ohshima et al. [142] showed how it is possible to use some simple assumptions concerning the missing constants and to estimate the molecular parameters. It is mainly the fact that the ratios a/d and −c/d should be in the ranges of 0.70-0.75 and 0.52-0.54 respectively for these kinds of radicals. For the C4 N and C6 N the molecular parameters could be calculated by scaling the C2 N value appropriate to the notation of [181, 142], i.e. using Eq. (6.35), (6.37), (6.38), and (6.40). The results for the Cn N chains are listed in Tab. 7.10 together with the values for the isoelectronic CH, C3 H, C5 H radicals; some of them are plotted in Fig 7.2. The d and the eQq2 constants are associated with the non-axial symmetry hsin2 θ/r3 i of the electron distribution in the radical. The difference between both constants is that d refers to the unpaired electrons (U) and eQq2 to the total electrons (T). In the case of C2 N hsin2 θ/r3 iU ≈ hsin2 θ/r3 iT which means that the non-axial distribution is mainly caused by the unpaired electron. For C4 N the situation is different because hsin2 θ/r3 iT − hsin2 θ/r3 iU = 6 · 1024 cm−3 and it is therefor not the unpaired electron that dominates the non-axial term. A comparison with other hsin2 θ/r3 iT -values from higher members of the Cn N (n even) chains would be desirable but for C6 N no eQq2 was determined and higher members are not detected so far. log (relative abundance) 7.2 C4 N and C6 N 117 C N radicals n n=3 0 4 -1 6 5 -2 2 3 4 5 6 7 number of carbon atoms Figure 7.12: Relative abundances of the Cn N radicals per gas pulse in the supersonic molecular beam as a function of chain length. Lines of C4 N and C6 N are readily observed in the supersonic molecular beam, even though both radicals are calculated ab initio to possess rather small dipole moments - 0.14 D for C4 N and 0.31 D for C6 N [144]. In FTM spectroscopy, line strengths are proportional to the first power of the dipole moment µ, not µ2 , as in classical absorption spectroscopy; rotational lines of small µ molecules are therefore relatively much more intense in FT spectroscopy than in conventional spectroscopy. Bauder and co-workers, for example, have detected a number of deuterated hydrocarbons [157] using the present technique; many with dipole moments in the range of 10−1 - 10−3 D. Relative abundances (see Fig. 7.2) of the nitrogen-bearing carbon chain radicals here to one another and to C3 N and C5 N were determined from intensity measurements on lines as close in frequency as possible to minimize variations in instrumental gain. These were converted to absolute abundances by comparing line intensities with those of the 118 7 Measurements and Analysis CnN, n=2,4,6 (positive) -3 0,4 4 3 0,2 0,1 2 0 4 n 6 6 2 3 4 n 5 6 CnN, n=2,4,6 -3 cm 6 24 0,3 2 24 Cn+1H, n=2,4 (scale x40) (x40) 0 Cn+1H, n=2,4 (scale x25) 5 4 3 3 〈 sin ϑ /r 〉U / 10 CnN, n=2,4,6 8 Cn+1H, n=2,4 (negative) 〈1/r 〉 / 10 cm 24 |〈 Ψ(0) 〉 | /10 cm -3 0,5 2 2 (x25) 1 0 2 4 n 6 Figure 7.13: Molecular constants h1/r3 i, hsin2 θ/r3 i, and hψ 2 (0)i of Cn N, n=2,4,6 and Cn+1 H, n=2,4 as estimated in Tab. 7.10 rare isotopic species of OCS in a supersonic beam of 1% OCS in Ar in the absence of a discharge, taking into account differences in the rotational partition functions and dipole moments. As Fig. 7.2 shows, the two new chains here are more than a factor of two more abundant than C5 N. The formation of carbon chain radicals in our molecular beam is apparently different for chains with odd and even numbers of carbon atoms. Although the abundance data for Cn N is much less complete than for Cn H, it is worth noting that the plot in Fig. 7.2 is similar to that previously derived for the acetylenic radicals (see Fig. 2 of Ref.[64]), implying similar, if not common, formation mechanisms in the discharge. Most of the C2n+1 H and C2n N chains are more abundant than the corresponding C2n H and C2n+1 N chains. If nonpolar carbon chains C2n+1 are more abundant than even-numbered chains C2n , for example, subsequent reactions involving the radicals C2 H or CN (produced directly via cleavage of the central C-C bond of either HC4 H or HC3 N) may produce the odd-even alternation that is observed. Evidence to support this formation mechanism is the mass distribution of a diacetylene discharge which exhibits an even-odd alternation in abundance for chains beyond C9 , with the odd chains being more abundant [156]. 7.2 C4 N and C6 N 2 Π 119 ∆ W =q D (J 2−1/4) (J+3/2) Π 7B 3/2 J=5/2 N=3 f e 5B f e J=3/2 6B S=+1/2 2B e transitions with ∆ F = 0 f transitions with ∆ F = −1 f transitions with ∆ F = 0 N=2 Aso = 40 cm −1 4B e transitions with ∆ F = −1 ∆ W = p eff (J+1/2) N=1 N=0 F=5/2 F=1/2 e J=5/2 F=3/2 f e f 3B Rotation Π Fine structure F=1/2 F=3/2 f F=1/2 e p eff = 4.5 MHz 1/2 Hund(b) F=1/2 e J=3/2 J=1/2 F=5/2 f 5B S=−1/2 F=3/2 F=3/2 14 MHz Λ −doubling Hyperfine structure (a +d)=37 MHz − (a −−d)=7 MHz Hund(a) Figure 7.14: Schematic C4 N energy level. The dominating splitting in the energy level diagram is due to the spin-orbital interaction which separates the Π3/2 and Π1/2 states. Transitions involving a Π3/2 state have not been measured and the energy separation is assumed to be 1.19 THz according to other similar species like C2 N, see [142]. In contrast to a Hund’s case (b) as it is plotted on the left with (2N)B spacings the fit was done in Hund’s case (a) and the rotational energy levels are now (2N+1)B separated. The Λ-doubling components are labeled by the e/f parity. The hyperfine structure is due to the nitrogen magnetic and quadrupole moment. 7 Measurements and Analysis 120 d Radical C2 N C4 N C6 N C3 H C5 H Table 7.10: Hyperfine and Molecular Constants of Carbon Chain Radicals bF -25b -12b -7b c 16.2 10.9 h 46.8 22.4 13.2 d - -9.2 5.6 - eQq2 0.1558 0.11 5.9 2.8 1.7 h1/r3 iU 0.2389 0.16 -2.9 -1.4 -0.8 h(3 cos2 θ − 1)/r3 iU 0.1368 0.09 5.5 2.6 1.5 hsin2 θ/r3 iU - 5.6 -3.4 c - hsin2 θ/r3 iT -0.0208 -0.03 0.24 0.25 0.04 hΨ2 (0)iU Molecular Constants (1024 cm−3 ) a 11.4 12.2 1.7 28.3 19.1 g Hyperfine Parameters (MHz) 34a 16.3a 9.6a -13.8 -21.8 f 14 12.3 8.3 e 14 0 µI ( N ) I (H) The calculations were done using Eq. (6.35), (6.37), (6.38), and (6.40) with 2µI( = 5.7208 and ( µI(H) )/( µI(I (14 NN)) ) ≈ 13.84. 14 N ) a calculated as in [142] with a/d = 0.70 − 0.75; b calculated as in [142] with −c/d = 0.52 − 0.54; c same -3e2 Q as in [142]; d constants taken from [67]; e estimation from a(H) with a/d = 0.76 like in the case of C3 H. f the b value is derived from the H C5 D b-value [88] with b(H)est = 2µ b(D) and b(C5 D) = −4.32M Hz. bF is than estimated with bF = b + c/3; g estimation µD from c(H) with c/d = 1.75 like in the case of C3 H. h constant taken from [35]; 7.2 C4 N and C6 N Table 7.11: Measured Rotational Transitions C4 N in the X 2 Π1/2 State. Transition Frequency a e/f b O−C c 0 0 J →J F →F (MHz) Λ Comp. (kHz) 1.5 → 0.5 1.5 → 1.5 7234.345 f 1 2.5 → 1.5 7247.840 e -1 1.5 → 0.5 7249.186 e 0 2.5 → 1.5 7255.402 f -1 1.5 → 1.5 7256.628 e -1 0.5 → 0.5 7257.124 e 0 0.5 → 0.5 7261.707 f 0 1.5 → 0.5 7271.774 f -1 2.5 → 1.5 2.5 → 2.5 12073.322 f -1 1.5 → 1.5 12084.411 f -1 3.5 → 2.5 12085.235 e 1 1.5 → 0.5 12086.839 e 0 3.5 → 2.5 12091.162 f 2 2.5 → 2.5 12094.262 e 2 2.5 → 1.5 12094.384 f 1 1.5 → 0.5 12094.480 f 0 1.5 → 1.5 12094.777 e 0 3.5 → 2.5 4.5 → 3.5 16921.368 e -1 3.5 → 2.5 16921.448 e -1 2.5 → 1.5 16922.195 e 0 4.5 → 3.5 16926.823 f 1 3.5 → 2.5 16928.186 f -1 2.5 → 1.5 16928.220 f 0 4.5 → 3.5 5.5 → 4.5 21757.146 e 2 4.5 → 3.5 21757.171 e -1 3.5 → 2.5 21757.636 e 0 5.5 → 4.5 21762.464 f -1 4.5 → 3.5 21763.210 f -1 3.5 → 2.5 21763.220 f 2 a Estimated experimental uncertainties (1σ) are 2 kHz. Designation of e and f levels is based on the assumption that the hyperfine constant d is positive. c Calculated frequencies derived from the best fit constants in Table 7.9. b 121 122 7 Measurements and Analysis Table 7.12: Measured Rotational Transitions C6 N, HUNDA-fit Transition Frequency a e/f b O−C c 0 0 J →J F →F (MHz) Λ Comp. (kHz) 4.5 → 3.5 5.5 → 4.5 7851.036 e -1 4.5 → 3.5 7851.047 e 0 3.5 → 2.5 7851.318 e 0 5.5 → 4.5 7853.245 f -3 4.5 → 3.5 7853.677 f 0 3.5 → 2.5 7853.677 f -4 5.5 → 4.5 6.5 → 5.5 9596.072 e 2 5.5 → 4.5 9596.072 e -2 4.5 → 3.5 9596.260 e 0 6.5 → 5.5 9598.199 f -1 5.5 → 4.5 9598.474 f 3 4.5 → 3.5 9598.474 f 2 6.5 → 5.5 7.5 → 6.5 11341.066 e -1 6.5 → 5.5 11341.066 e -2 5.5 → 4.5 11341.203 e 0 7.5 → 6.5 11343.149 f -1 6.5 → 5.5 11343.336 f 0 5.5 → 4.5 11343.336 f 0 7.5 → 6.5 8.5 → 7.5 13086.043 e -1 7.5 → 6.5 13086.043 e 0 6.5 → 5.5 13086.146 e 0 8.5 → 7.5 13088.099 f 1 7.5 → 6.5 13088.233 f 0 6.5 → 5.5 13088.233 f 2 8.5 → 7.5 9.5 → 8.5 14831.011 e 2 8.5 → 7.5 14831.011 e 4 7.5 → 6.5 14831.085 e -1 9.5 → 8.5 14833.047 f 3 8.5 → 7.5 14833.145 f -1 7.5 → 6.5 14833.145 f 2 9.5 → 8.5 10.5 → 9.5 16575.958 e -1 9.5 → 8.5 16575.968 e 1 8.5 → 7.5 16576.024 e -1 10.5 → 9.5 16577.985 f -2 9.5 → 8.5 16578.057 f 0 8.5 → 7.5 16578.068 f 3 10.5 → 9.5 11.5 → 10.5 18320.906 e -4 10.5 → 9.5 18320.916 e 3 9.5 → 8.5 18320.973 e ... 11.5 → 10.5 18322.929 f 1 10.5 → 9.5 18322.986 f -2 9.5 → 8.5 18322.990 f -2 a Estimated experimental uncertainties (1σ) are 2 kHz. b Designation of e and f levels is based on the assumption that the hyperfine constant d is positive. c Calculated frequencies derived from the best fit constants in Table 7.9. 7.3 The Search for C7 N 123 7.3 The Search for C7N Laboratory searches were also undertaken for C7 N using experimental conditions and gas mixtures that optimize the production of either C5 N or C6 N, or both molecules simultaneously. Searches were based on a high-level coupled cluster calculation by Botschwina [20], and covered frequency ranges that correspond to ±1% of the predicted rotational constants. Two searches, one assuming a 2 Σ ground state with rotational lines separated in frequency by integer quantum numbers, and one assuming a 2 Π ground state with rotational lines separated in frequency by half-integer quantum numbers, were performed, but no lines which could be attributed to C7 N were found in either survey. The absence of lines requires the C7 N line intensities to be at least 60 times less than those of C5 N, which is readily observed in the molecular beam with a signal to noise of better than 25 in one minute of integration. The failure to detect C7 N may indicate a 2 Π ground state for this molecule. Botschwina concluded on the basis of RCCSD(T)/cc-pVTZ calculations [20] that the ground state of C7 N is 2 Π (µ = 0.96 D), but that a 2 Σ+ state (µ = 3.86 D) is very close in energy, lying only 250 cm−1 above ground at the highest level of theory. If we assume the same abundance decrement from C5 N to C7 N as from C3 N to C5 N (a factor of 17) and a 2 Π ground state, the C7 N lines would be 180 times less intense than those of C5 N, i.e. three times below our present upper limit. If the ground state is 2 Σ+ instead, the expected decrease in line intensity is only a factor of 45. In either case, significant rovibronic interaction may occur between these two low-lying electronic states, a factor which could hinder spectral analysis and assignment regardless of the symmetry of the ground state. Detection of C7 N may still be possible: with further improvements in instrumentation and production efficiency a factor of five or more in sensitivity may be within reach. 7.4 Conclusions and Prospects Without a further improvement of the sensitivity of the FTMW spectrometer the detection of higher members of the Cn N chains, like C7 N or C8 N seems exceedingly difficult. For the same reason measurements on C5 N, C4 N, and C6 N isotopomers, i.e. to determine the exact bonding length or the spin density distribution in these radicals, appear less feasible. However, additional isotopic spectroscopy using 13 C-enriched samples of cyanogen, cyanoacetylene, methylcyanide, etc. may allow the distribution of carbon in the discharge source to be determined, providing clues to the chemical processes at work. This could turn out to be an important tool for a molecule production improvement. It is known that also non linear structures are build in a discharge jet [123]. Lowlying isomers of C4 N and C6 N may be amenable to laboratory detection with present techniques. Ding et al. [46] recently calculated the potential energy surface of C4 N, and concluded that 13 isomers, some with unusual ring, branched, and caged structures, probably exist. The isomer of the most immediate laboratory interest is the ring-chain analog to c-C5 H, with the CCH group replaced by a nitrile group. This isomer is 124 7 Measurements and Analysis predicted to have considerable kinetic stability towards isomerization and dissociation, and is calculated to lie only 2.8 kcal/mol above the linear chain. Because it is also predicted to possess a substantial dipole moment (µ = 0.63 D), and because c-C5 H has already been detected with the same discharge source [4], detection of c-C4 N may succeed with dedicated searches. Other lowlying polar isomers of C4 N such as CCCNC (23.4 kcal/mol; 1.38 D) may also be within reach. The electronic spectra of C4 N and C6 N are completely unknown. Both radicals probably possess strong 2 Π - X 2 Π electronic transitions at visible or near infrared wavelengths, like the shorter chain CCN [142] and most of the acetylenic chains Cn H up to C10 H. Many of these have now been studied by sensitive laser techniques (see e.g., Ref. [110]), including LIF, cavity ring-down laser absorption spectroscopy (CRLAS), and most recently, resonant two-color, two-photon ionization spectroscopy (R2C2PI) combined with time-of-flight mass detection. All of the species up to C6 N have abundances near the throat of the discharge nozzle of > 109 molecules per pulse - an adequate number density for all three techniques - but the best choice would appear to be R2C2PI because the optical spectrum of linear C3 H is significantly broadened owing to rapid internal conversion [45], and because REMPI is more sensitive than CRLAS and is mass selective as well. Up to now there is also a lack of ro-vibrational data even if one considers only linear Cn N radicals. In the case of C2 N - the first member of the linear Cn N chains - the Â2 ∆-X̂ 2 Π electronic transition of the CCN free radical has been observed by Oliphant et al. [143] in emission with a high-resolution Fourier transform spectrometer. Spectroscopic constants were derived including the ground-state vibrational frequencies, ν3 =1050.7636(6), 2ν3 =2094.8157(18), and ν1 =1923.2547(69) cm1 . But for the C3 N radical the vibrational frequencies have only been investigated by ab initio theory. Botschwina [121] using coupled cluster calculations RCCSD(T) determined the asymmetric stretching vibrations to be at ν3 = 2311.8 cm−1 and ν2 = 2116.6 cm−1 . There is no verification of these numbers by gas phase experiments so far, whether in the IR nor in the visible spectral range. For the C4 N radical the vibration frequencies are calculated to be at ν2 =1995 cm−1 and ν1 =2186 cm−1 by Ding et al.[46]. C3 N and C4 N are therefore good candidates for a IR gas phase detection, e.g. using the Cologne IR experiment. Part III Linear CnN Chains in Space 8 CnN Chains in Space Remember, too, That the whole sky is revolving With its constellations, its planets. I have to force my course against thatNot to be swept backwards as all else is. [...] Even if you were able to stick to the route You have to pass The horns of the Great Bull, the nasty arrows Of the Haemonian Archer, the gaping jaw Of the infuriated Lion,... Ted Hughes, ”Tales from Ovid”, Phaethon Presently there are more than 125 gas phase molecules found in space. The species range in size from 2-13 atoms. They are typically found in dense interstellar clouds with tremendous sizes of 1-100 light years, average gas densities of 102 -103 cm−3 , and temperatures in the range 10-60 K 1 . On the other hand many of the organic molecules are detected in extended circumstellar envelopes of cool, old stars that are carbon rich like CW Leo/IRC+10216 or in the somewhat ’special cases’ of molecular clouds TMC-1 and TMC-2. The large abundance of highly unsaturated carbon chains and radicals, most of which are linear, is a characteristic feature of IRC+10216 2 and except TMC-1 3 no other source in the sky shows such a wealth of long linear chains [34]. 1 Both, higher temperatures and higher gas densities are found in localized regions where star formation is occurring. 2 IRC+10216 belongs to the stars with initial masses in the range from about 1 to 8 M . When these stars “evolve off the main sequence, they go through several stages of evolution in which mass loss plays a crucial role [. . . ]. For such stars the mass loss usually occurs as a cool, low-velocity wind. During the asymptotic giant branch (AGB) evolutionary phase, mass loss rates may reach 1 × 10−4 M yr−1 . The resulting high density of such winds near the star and the low temperature of the stellar photosphere assure that most of the ejected material is in molecular form.“ [15]. 3 TMC-1 (Taurus Molecular Cloud 1) is a star forming region with low temperatures (T= 6 - 10 K). 128 8 Cn N Chains in Space Figure 8.1: The molecular clouds in the plane of the Milky Way, as seen in the 1-0 rotational transition of CO. So far, most of the over 120 astronomical molecules have been detected in a small number of locations, sometimes only one or two. IRC+10216 is one of the reaches molecular sources in space and CN, C3 N and C5 N have been found in this object. [179] So far, detection of Cn N radicals succeeded for the odd members C3 N and C5 N whereas non of the even membered cyano chains, e.g. C2 N, could be observed in the interstellar space. C3 N was first found in the circumstellar envelope of IRC+10216 by Gúelin & Thaddeus [78]. Since then it has been found in many other sources like IRAS 15194-5115, the proto planetary nebulae (PPN) CRL 618 and CRL 2688, and the molecular clouds (MC) TMC1, TMC-2 and HCL-2. A list of astronomically measured lines is given in the appendix (Tab. D.2). C3 N can best be seen in IRC+10216 were it has a rotational temperature of Trot ≈ 20 K. So far, only one survey has been undertaken in which isotopes of C3 N were detected (Cernicharo et al. [34], see appendix Tab. D.1). Unfortunately not all astronomically observed transitions match the predictions obtained from the newly presented laboratory data. The astronomical survey reveals that C3 N isotopomers have very weak intensities and are hardly to be recognized. The discrepancy in the line assignment could therefore be due to a misinterpretation of astronomical data. In Fig. 8.2 a detail of the line survey is shown. On the left side of the picture are three lines indicated by arrows. The left (L1) and middle (L2) line are assigned to 13 CCCN and the right one (L3) to the U152681 unidentified line. The problem is that non of the laboratory measured transitions corresponds to the left line (L1). If the lines L2 and L3 are assigned to 13 CCCN instead of L1 and L2 all disagreements between laboratory and astronomical data are solved. However, an other misassignment at 143104.0 MHz (which can not be seen in Fig. 8.2) remains4 . The next larger member of the odd Cn N chains C5 N was first detected in TMC-1 and IRC+10216 by Gúelin et al. [77]. C5 N appears to be two orders of magnitude less 4 Other unidentified lines in the survey at 142831.1 ± 10 MHz and 143575.8 ± 15 MHz do not match to laboratory transitions at 143129.47 MHz and 143136.22 MHz. 129 Figure 8.2: 13 CCCN detection in IRC+10216 by Cernicharo et al.[34] abundant than the related molecule HC5 N, i.e. N (HC5 N/C5 N)' 200. In comparison the HC3 N to C3 N abundance ratio is of the order of 20, i.e. N (HC3 N/C3 N)= 19. It is assumed that the rotational temperature of C5 N is the same as from HC5 N, i.e. Trot ≈ 29K. In 1991 Pauzat et al. [144] analyzed the feasibility of linear Cn N detection in space and concluded that due to the small dipole moments all even Cn N members in their 2 Π electronic ground state would be poor candidates for interstellar detection. Recently, in a theoretical work Mebel & Kaiser [127] examined the formation of interstellar C2 N isomers via neutral-neutral reactions in the interstellar medium like C(3 Pj ) + HCN → C2 N + H(2 S1/2 ) . (8.1) The formation of the C2 N isomers is calculated to proceed without any barrier, but reactions forming CCN, CNC, and c-C2 N are found to be strongly endothermic by 52.7, 59.0, and 99.6 kJ mol−1 , respectively. Considering this result it seems to be highly unlikely that C2 N can be synthesized in cold molecular clouds where average translation Fig. 2. continued temperatures of the reactants are only 10-15 K. The physical conditions in circumstellar envelopes of late type stars like IRC+10216, are different. Close to the photosphere of the central star temperatures can reach 4000 K and the elevated velocity of the reactants in the long tail of the Maxwell-Boltzmann distribution can overcome the reaction endothermicity to form CCN. These type of environments represent ideal targets for hitherto undetected C2 N in the infrared as well as in the microwave region. Unfortunately, a simple reaction (see Eq. 8.1) is not a realistic scenario for complex environments like the IRC+10216 envelope, PPN, molecular clouds or other interstellar molecule sources. In this situation, astrochemical models specified for the environment in question can be helpful for estimations of molecule abundances. Nowadays these models are of great complexity and it is often not easy to determine the most important synthesis of a species. Especially, if one considers the fact that reaction rates depend on the density, temperature, and radiation which in turn are functions of the position in the circumstellar envelope or cloud it is clear that the synthesis of a specific species is also a function of time and position. Because of the importance of IRC+10216 as a 130 8 Cn N Chains in Space C 2H 2 CN HC3 N C H 2 H 2 C 2 C 4H 2 CN HC N C3 N HCN 5 C H 2 H 2 C 2 C 6H 2 HCN CN HC7 N C5 N HCN C 2H N C 3H C 4H N C 5H C 6H C H 2 HC9 N Figure 8.3: Major pathways to cyanopolyyne formation in IRC+10216. Although C3 N is probably build by photo-destruction of HC3 N it seems to play an important role in the formation process of HC5 N and consequently for all higher cyanopolyynes. [131],[38] major molecule source with great diversity a short description of the Millar & Herbst [131, 132] astrochemical model of its circumstellar envelope is given here. The gas around IRC+10216 is assumed to expand in a spherically symmetric outflow 5 with a velocity of 14 kms−1 and a mass-loss rate of 3 × 10−5 M yr−1 . The model follows the chemistry from an inner radius ri = 1 × 1016 cm to an outer radius of 1018 cm. At ri parent species (H2 ,He,CO,C2 H2 ,CH4 ,HCN,NH3 ,N2 , and H2 S) are injected into the outward flow. The choise of an inner radius is governed by the onset of the photochemistry [132]. As the parent molecules flow outwards they are dissociated into reactive daughter products by the incident interstellar ultraviolet radiation field6 . For molecules, the average time to travel from the inner to the outer radius takes approximately 10 000 years. The number density of the gas follows an r−2 distribution, while the temperature profile T(r) in K is assumed to be T (r) = 100(r/ri )0.79 but never to be less than 10 K. In this way, the temperature never decreases below 10 K. The rates of photodissociation and photoionization as well as the strength of the un-extinguished radiation field is included in the model. The model network contains 407 species connected by 3851 reactions 7 . In this circumstellar model, as opposed to the interstellar models, neutralneutral reactions and photo-destruction are very important especially for the formation of larger hydrocarbons. In the distance of 3·1016 cm where much of the chemistry takes place the major formation 5 This smooth outflow is a oversimplification since observations [117] have revealed that the envelope consist of discrete, concentric shells. 6 that is by UV radiation coming from far outside IRC+10216 7 Hydrocarbons with more than 23 atoms are not included in the network. 131 mechanism for C3 N (cyanoacetylene) seems to be the photo-destruction of HC3 N. The latter is formed mainly via the two reactions CN + C2 H2 −→ HC3 N + H HCN + CCH −→ HC3 N + H (8.2) (8.3) On the other side the major depletion mechanism for C3 N is photo-destruction to form C2 and CN. C3 N is also an important constituent in the major pathways to cyanopolyyne formation in IRC+10216, as can be seen in Fig. 8.3. Similar to C3 N the next member of the Cn N (n odd) chains C5 N seems to be formed by photo-destruction of HC5 N which is produced similar to Eq. 8.2 and 8.3, i.e. by CN+C4 H2 → HC5 N + H and HCN + C4 H → HC5 N + H. The main question for Cn N (n even) astrochemistry is how likely the detection of C2 N is. Following this model a list of theoretical and observed column densities including the precursor molecules HCn N and Cn H is given in Tab. 8.1 and plotted in Fig. 8.4. C2 N has probably a two magnitudes smaller column density than C3 N so that it can hardly be seen in normal line surveys. The column density ratios 8 NL (HC3 N/C3 N) and NL (HC5 N/C5 N) are of the order of 10. If this is also valid for NL (HC2 N/C2 N) an astronomical search seems to be reasonable. The parent molecule HC2 N has already been observed towards IRC+10216 and the column density of C2 N is expected to be less than 3.6 · 1012 cm−2 which should hence be detectable as well. For comparison: When HC11 N was unambiguously detected in TMC-1 in 1997 by Bell et al. [8] 9 the signal corresponded to a column density of 2.8 · 1011 cm−2 at Trot =10 K. Assuming a rotational temperature similar to C3 N (Trot = 20 K) the best frequency range to search for C2 N is between 150 and 180 GHz with the transitions J 6.5→5.5 and 7.5→6.5. 8 9 Not to be confused with the abundance ratios mentioned earlier in this chapter. The 1982 detection in IRC+10216 by Bell et al. [9] was probably incorrect 132 8 Cn N Chains in Space molecule HCN HC2 N HC3 N HC5 N HC7 N HC9 N C2 H C3 H C4 H C5 H C6 H C7 H C8 H CN C2 N C3 N C4 N C5 N C7 N Trot [K] 20 35 25 35 35 52 20 35 theor. observed Tex Col.Dens. Col.Dens. Frac. abund. [K] [cm−2 ] [cm−2 ] N(X)/N(H2 ) 1.7 1016 2.8 1016 12 1.2-1.8 1013 26 1.8 1015 0.8-1.7 1015 25-35 7.1 1014 1.3-3.7 1014 26 2.2 1014 1.3 1014 12-23 5.8 1013 2.7-4.0 1013 16 5.7 1015 4.6-5.0 1015 7.1 10−6 8.5 1.4 1014 5.6 1013 15 1.0 1015 2.4-3.0 1015 4.3 10−6 13 14 27(Π1/2 ) / 39(Π3/2 ) 8.7 10 0.4-2.9 10 6.3 10−8 35(Π3/2 ) / 46(Π1/2 ) 5.8 1014 0.6-1.7 1014 7.8 10−8 13 12 4.5 10 2.2 10 3.1 10−9 1.1 1014 1.0 1013 1.4 10−8 8.7 1.0 1015 6.2 1014 12 3.6 10 15 3.2 1014 2.5-4.1 1014 3.5 10−7 8.2 1009 1.4 1014 6.3 1012 9.0 10−9 7.8 1012 Table 8.1: Molecular column densities in IRC+10216, observed values are taken from Cernicharo et al. [34] and Kawaguchi et al. [101]. Theoretical values are from Millar & Herbst [132]. column density [cm ] -2 column density [cm ] -2 1e+09 1e+10 2 4 2 n HC2N 6 8 HCnN observed theor. 4 10 column density [cm ] -2 n 1e+12 1e+13 1e+14 1e+15 1e+16 0 6 2 4 n 8 8 observed (n odd) theor. (n even) theor. 6 CnN observed (n odd) theor. (n even) theor. CnH Figure 8.4: Column densities of Cn N, Cn H, and HCn N in IRC+10216, see Tab. 8.1. 0 0 1e+11 1e+12 1e+13 1e+14 1e+15 1e+16 1e+13 1e+14 1e+15 1e+16 1e+17 10 10 133 134 8 Cn N Chains in Space . TA* (K) Figure 8.5: Possible C2 N transition towards IRC+10216 at 248 GHz. The left line appears at the C2 N 248 GHz transition frequency with µ2 S=1.2 - 1.9 D2 . The dotted line results from a line shape fit for molecules in a circumstellar envelope. The line at the right hand side could not be identified. The total integration time was 170 min. 8.1 The Search for Interstellar C2N In the previous section it has been shown that the detection of C2 N may be in reach with the present day radio telescope techniques. The astronomical search for this molecule was guided by the use of laboratory data. In 1995 Ohshima and Endo [142] did measurements on C2 N using a Fourier transform spectrometer in the Microwave region. Their molecular spectroscopic constants were used as reference as well as the constants derived by Kakimoto and Kasuya [98] to calculate the millimeter rotational spectrum of C2 N (see Tab. 8.3). 8.1.1 Observation In September 2002 a search was performed for the C2 N radical towards IRC+10216 employing the IRAM 30m telescope at Pico Veleta, Spain (see Fig. 8.8). The observations were done at position (Eq 1950) RA 09:45:14.8 Dec 13:30:40.0 and focused on four 500 MHz broad frequency bands with centers at 83 GHz, 154 GHz, 224 GHz and 248 GHz corresponding to the J=7/2 → 5/2, 13/2 → 11/2, 19/2 → 17/2 and 21/2 → 19/2 rotational transitions of C2 N, respectively. The integration time was 4 - 4 1/4 hours 8.1 The Search for Interstellar C2 N 135 . TA* (K) . TA* (K) Figure 8.6: Possible C2 N transitions towards IRC+10216. The C2 N transitions appear at 224.5 GHz (with µ2 S = 1.9 - 2.3 Debye2 and a total integration time of 265 min.) and 153.6 GHz (with µ2 S = 4.2 - 5.8 Debye2 and a total integration time of 250 min.). The dotted lines result from line shape fits for molecules in a circumstellar envelope. The right line in the top picture at 224.7 GHz is C17 O (v=0, J 2 → 1). The C2 N line at 224.5 GHz can also be due to a possible upper sideband image of SiC2 at 232534 MHz. 136 8 Cn N Chains in Space Table 8.2: Observational parameters, IRC+10216 molecule transition obs.freq. a Tsys J (MHz) (K) C2 N 21/2 → 19/2(e/f) [ 19/2 → 17/2(e) [ 19/2 → 17/2(f) 13/2 → 11/2(f) 248189.9(3) 224545.2(12) 224551.2(12) 153644.7(5) 800 650 650 400 U1d C17 O C2 S NaCN U2d HC5 N (v=0) 2 → 1 11,12 → 11,11 100,10 → 90,9 248291.22(8) 224714.18(15) 153449.21(5) 153557.04(7) 153842.29(9) 82538.835(3) 800 650 400 400 400 105 31 → 30 ρ vexp R T∗A dν [km m−1 ] (K.km s−1 ) 0.49 0.53 0.53 0.67 14.6 14.7 14.7 14.7 0.78(7) 0.12(9) ]c 0.12(9) ]c 0.04(1) 0.49 0.53 0.67 0.67 0.67 0.80 14.0 14.7 13.9 12.7 16.1 14.0 2.69(7) 1.78(3) 0.81(5) 0.84(5) 0.60(6) 0.73(3) Rest frequencies (assuming a source LSR velocity of -27 km s−1 ) in IRC+10216; The numbers in parenthesis are the estimated uncertainties (1σ); c poss. upper sideband image of SiC at 232534 MHz; 2 d unidentified line a b and for most of these bands the r.m.s. noise is 9 - 13 mK 10 , low enough to reveal lines as weak as 0.01 - 0.04 K. Optically thin lines from the outer shells of the IRC+10216 envelope have a U-shaped profile (see Fig. 8.5 and 8.6) where the two horns 11 arise from the blue-shifted (front), and red-shifted (rear) polar caps. The emission at the center originates from the meridian ring perpendicular to the line of sight. The horn-tocenter intensity ratio of emissions coming from spherical shells of constant thickness and constant radial velocity12 depend primarily on the shell diameter relative to the telescope beam [73], i.e. for a given frequency we have: the larger the shell, the larger the ratio. The lines at 224 and 248 GHz reveal the typical U-shape which suggests that the carrier of these lines is essentially present in the outer envelope. The measured intensities are given in TA∗ , the effective antenna temperature corrected for spillover losses and atmosphere attenuation. TA∗ is related to TM B the main beam-averaged source brightness temperature by Ta∗ = ρTM B , where ρ = ρ(ν) is the 30m telescope beam efficiency, see Table 8.2. scale is in TA∗ , the effective antenna temperature corrected for spillover losses and atmosphere attenuation. 11 This notion is taken from [73]. 12 see vexp in Tab. 8.2. 10 8.1 The Search for Interstellar C2 N 137 8.1.2 Data Analysis The Grenoble molecular line reduction software CLASS was used to fit the observed lines to an U-shaped line profile. Two lines are observed13 at frequencies of C2 N transitions with integrated line intensities of at least 0.78 and 0.24 K km s−1 , and one weaker line is found at 154 GHz with an intensity of 0.04 K km s−1 , see Table 8.2. The frequency scale is computed for a LSR source velocity of -27 kms−1 . The total molecular column density NT as well as the rotational temperature T are important for the unambiguous assignment of the observed transitions to a certain molecule. There to calculate this values. The integrated R are some formulas needed 2 line intensity Tl dν is proportional to µ S, with S the line strength of the transition. The dipole moment of C2 N has been calculated by Pd et al. [145] to be µ=0.425 Debye. The line strength can be calculated from the intensity I of a transition or vice versa with the following formula [154] 10Ip Q(T ) µS= l 4.16231 · 10−5 · ν exp −E − exp kT 2 −Eu kT (8.4) with Ip the logarithm of the intensity 14 ,ν in MHz, µ the relevant component of the dipole moment of the molecule in Debye, Q the partition function [152], El the lower state energy and Eu the upper state energy of the transition. For reasons of conformity the theoretical intensities of the C2 N transitions were calculated with the dpfit and dpcat program written by Herb Pickett. Here T is set to 300 K 15 . In case of small optical depth 16 the column density Nu in its upper level can be calculated by 17 R Nu 17 TM B dv = 1.669813 · 10 (8.5) gu νµ2 S R with gu the statistical weight of the level [196], TM B dv the line integral 18 in (K km s−1 ), ν in MHz and µ2 S from Eq.8.4. The line width in IRC+10216 of the newly measured transitions are of the order 30 km s−1 corresponding to 20-25 MHz for 248 GHz, see Fig. 8.5. If C2 N is the carrier of the absorption lines many hyperfine transitions remain unresolved, i.e. three lines at 153 GHz, six lines at 224GHz and also six at 248 GHz, see Fig. 8.5 13 The line at 224.5 GHz is a possible upper sideband image of the SiC2 line at 232.534 GHz. The Pickett program uses the logarithm Ip of the intensity I instead of the intensity itself in its catalog files, i.e. I = 10Ip 15 This is a necessary standard procedure because the Pickett program sums over a finite amount of transitions to calculate the partition function and not over all. Thus differences can occur if an other temperature is chosen. Q is given in the Pickett filename.out file. For C2 N Q(300K) = 5737.4682 16 and for TexR hν/k and Tex Tbg , see [196] Tl dv 3k 17 Nu gu = 8π 3 νµ2 S 18 Some times the line integral Ris given in K·MHz instead of K·km·s−1 . The conversion equation is R Tl dv [K km s−1 ]= 10−3 νc Tl dν [K MHz] with c the speed of light in [m/s] and ν the transition frequency in [MHz]. 14 138 8 Cn N Chains in Space Table 8.3: Line parameters, IRC+10216 molecule C2 N U1 C17 O C2 S NaCN U2 HC5 N transition obs.freq.a calc.freq. o-c Eupper J (MHz) (MHz) (MHz) (K) 21/2 → 19/2(e/f) [ 19/2 → 17/2(e) [ 19/2 → 17/2(f) 13/2 → 11/2(e) 13/2 → 11/2(f) [ 7/2 → 5/2(e) [ 7/2 → 5/2(f) 248189.9(3) 224545.2(12)c 224551.2(12)c 153644.70(8)c - 248189.1(1) 224545.2(1) 224551.2(1) 153617.9(1) 153644.70(8) 82701.63(1) 82743.70(1) 0.8 27.8 - 68.6 56.7 56.7 27.8 9.1 9.1 11.2d ] 11.2d ] (v=0) 2 → 1 11,12 → 11,11 100,10 → 90,9 31 → 30 248291.22(8) 224714.18(15) 153449.21(5) 153557.04(7) 153842.29(9) 82538.835(3) 224714.389(3) 153449.774(11) 153557.651e 82539.040 -0.21 -0.56 -0.61 -0.21 16.2 53.8 63.4 13.5 10.2 9.3 log(Nu ) (with Nu in cm−2 ) 11.0 10.6 ] 10.6 ] - Rest frequencies (assuming a source LSR velocity of -27 km s−1 ) and intensities in IRC+10216; The numbers in parenthesis are the estimated uncertainties (1σ). c In a free fit the rest frequencies have been 224549.64(188) MHz for the 19/2 → 17/2(e/f) transitions and 153642.61(10) for the 13/2 → 11/2(f) transition. Because of the close lying doublet components both transitions have been fitted with fixed rest frequencies. d integrated noise, 0.003K · 30 km s−1 e see [34] on-line data a b and 8.6. In this case a mean line strength S̄ of the unresolved transitions is computed and multiplied by the number n of these lines to give S̄s the summed line strength that replaces S in Eq. 8.5. If several resolvable transitions, e.g. rotational transitions, of a molecule are observed a rotational temperature can be assigned by using Nu NT Eu = exp − (8.6) gu Q(Trot ) kTrot where NT is the total molecular column density, Q the partition function, and Eu the energy of the upper level. Eq. 8.6 can be rewritten in the form Nu NT 1 log10 = log10 − log10 (e) Eu [in K] (8.7) gu Q(Trot ) Trot where Trot and NT 19 19 can be determined by a least-square fit [114]. The result can be The partition function for a temperature T has been calculated by Q(T ) = α · T β , where α and β are determined by the Q(Ti ) values given by the Pickett program. In the case of linear closed shell molecules β is always one but for the C2 N radical α=6.7 and β = 1.2 . 8.1 The Search for Interstellar C2 N 139 11,5 11,0 log10(Nu/gu) log10(Nu/gu) 11,0 10,5 10,5 10,0 10,0 9,5 0 20 40 Eu[K] 60 0 20 40 Eu[K] 60 Figure 8.7: Boltzmann plot for C2 N. The left plot is calculated for a non-diluted beam and the right plot is calculated for a diluted beam. (Left) A rotational temperture of 78±60 K and a column density of 1014 cm−2 could be estimated in the case of a diluted beam. plotted in a Boltzmann plot. The spacial extend of a molecular gas can influence the value of the integrated line intensities. This is because the beam size diameter θB of the telescope varies with frequency, i.e. θB ∼ 1/ν (8.8) The IRAM 30m telescope beam width is roughly 2400”/ν[GHz], i.e. 28” for 83 GHz, 15” for 154 GHz, 10” for 224 GHz, and 9” for 248 GHz. The range of possible line intensities can be estimated by considering two cases of telescope beam and source diameter ratios. In one case the area of emitting molecules is larger than the beam size so that the beam is completely ’filled’ independent of frequency. On the other hand, when the source is small the telescope beam can easily be larger than the diameter θs of the emitting area , i.e. the beam is diluted. Assuming that the intrinsic intensity distributions Ts of a source is Gaussian a beam filling factor can be introduced to calculate the measured intensity TL , i.e. θS2 TL = TS · (8.9) 2 θS2 + θB 2 If θs θB this equation reduces to TL = TS · ( θθ2 ) and applying Eq. 8.8 gives B TL (ν) ∼ TS · ν νnorm 2 . (8.10) with νnorm > ν. A factor (ν/νlowest )2 added on the right side of Eq. 8.5 modifies the column densities and thus also the rotational temperature Trot into T̃rot in the Boltzmann plot, see Fig. boltz, i.e. T̃rot represent the cases of a diluted beam. The distribution of the C2 N radicals around IRC+10216 is not known. However, if it is similar to the 140 8 Cn N Chains in Space C3 N intensity distribution (see [15]) which is shell like with some larger irregularities then the determination of the filling factor for each frequency can only be obtained by mapping the star envelope. This has not been done here. The result of the Boltzmann fit for a diluted beam is the temperatures T̃rot = 78±60 K and the total C2 N column density NT = 1014 cm−2 . The uncertainties of the temperature fit are very large and partly due to the fact that the integrated line intensity at 224 GHz is lower than the one at 248 GHz, i.e. it does not exactly follow a boltzmann distribution 20 . Calculations with a non-diluted beam result in a negative temperature. If the three lines at 154, 224 and 248 GHz belong to C2 N, it has to be shown that the non-detection of the 83 GHz transition is consistent with this assignment. To proof this the noise at the 83 GHz line positions were integrated according to the beam filling factor 1 or (83/154), see Tab. 8.3. The integrated noise intensities are above the fitted line (dotted) in Fig. 8.7 (left) which marks the level of consistent signal intensities. The expected C2 N signal at 83 GHz is smaller than the noise intensity and should therefore not be detectable. Hence, the non detection of the 83 GHz is consistent with a C2 N assignment of the other three lines. 8.1.3 Discussion If C2 N is produced via photo-destruction of HC2 N the physical conditions and the region in which they are detected should be similar or at least correlated to each other. In 1991 Guélin & Cernicharo [74] detected HCCN towards IRC+10216 using the IRAM 30m telescope. They found a source diameter ≤ 25” (i.e. little or no beam dilution). Comparison of the line profile of HC2 N with HC3 N and HC5 N indicate that this molecule is essentially present in the outer envelope. The rotational temperature of HCCN is Trot = 12 ± 4 K and C3 N has a rotational temperature of 20 K [34]. Guélin & Cernicharo detected a faint signal which coincides with the 13/2 → 11/2 of C2 N but no feature stronger than 0.02 K at the frequency of the 11/2 → 9/2 transition. Both the non-detection at 130 GHz by Guélin & Cernicharo as well as the discrepancy of the rotational temperatures in the case of a non-diluted beam between HC2 N and C2 N make it difficult to assign the newly observed transitions to the C2 N molecule. Guélin & Cernicharo who assumed a dipole moment of 1.3 Debye for C2 N estimated an upper limit of NT (1.3 D) = 5 · 1013 cm−2 for the column density of C2 N towards IRC+10216. The observation presented in this thesis together with the assumption that C2 N has a dipole moment of 0.425 Debye (Pd & Chandra [145], 2001) suggest that the column density of C2 N is NT (0.425 D)= 1014 cm−2 whereas astrophysical models by Millar & Herbst [132] predict a column density21 of 3.6 · 1012 cm−2 , see Table 8.1. 20 In this situation the uncertainties of the individual line intensities become particular important for the result of the fit. 21 Note: It should be NT (0.425 D) ≈ 10 · NT (1.3 D). 8.1 The Search for Interstellar C2 N 141 Figure 8.8: The IRAM 30m telescope at Pico Veleta, Granada 8.1.4 Conclusions and Prospects During the measurements at the IRAM 30m telescope (see Fig. 8.8) three lines at frequency positions of C2 N transitions could be detected and have been assigned to C2 N as a possible carrier. This assignment can only be tentative and more data is needed. However, an upper limit of the column densities of C2 N towards the envelope of IRC+10216 can be given. They are of the same order than those obtained from previous measurements by Guélin & Cernicharo [74] and are a factor of ten larger then predicted values by Millar & Herbst [132]. An unambiguous detection of C2 N may still be possible with dedicated searches using long integration times. In a free line fit procedure the observed line positions reveal deviations > 1σ with reference to the calculated frequencies. Up to now the predicted line positions in the mm range are mainly based on laboratory measurements at 35 GHz by Ohshima et al. [142]. Therefore new laboratory measurements of C2 N are necessary in the frequency region between 100 and 300 GHz. The measurements reported in Chapter 7 should serve as a guide for future astronomical observations of C4 N, C6 N, and the isotopic species of C3 N. The provided spectroscopic constants allow the astronomically most interesting radio lines of these to be predicted to an uncertainty of 0.30 km sec−1 or better up to 50 GHz. Astronomical detection of the carbon-13 species of C3 N in the cold molecular cloud TMC-1 is also likely, because they have already been detected by Cernicharo et al. [34] in the circumstellar shell of the evolved carbon star IRC+10216. 142 8 Cn N Chains in Space Part IV Appendix A Linear CnH Linear Cn H radicals are iso-electronic to Cn N and therefore of interest for this thesis. The following table is the analogue to Tab. 5.1 and 5.2 for Cn N radicals. Table A.1: Linear Cn H (n=1-8) radicals molecule ground state dipole moment [Debye] exp. B B value [MHz] first detection lab, vis (1920-25) astro, vis (1937) astro, radio (1973) lab, MW (1983) lab (1985) astron (1985) lab (1986) astron (1986) lab (1996) astron (1997) lab (1997) lab (1997) lab (1998) orig. contribution (see for reference) Cn H, n odd CH X2 Π 1.46 425472.8 C3 H X2 Π 3.29 11186.335 C5 H X2 Π 4.44 2395.131 C7 H X2 Π 5.29 875.484 C9 H C11 H C13 H X2 Π X2 Π X2 Π 413.258 226.900 137.710 (Herzberg [83]) Dunham, Swings [125, 174, 49, 83] Turner [185] Brazier, Brown [22] Gottlieb, Vrtilek [68] Thaddeus, Hjalmarson [178] Gottlieb, Thaddeus [65] Cernicharo, Kahane [36] Travers, McCarthy [182] Guelin et al. [75] McCarthy [122] McCarthy [122] Gottlieb [64] Cn H, n even C2 H X2 Σ 0.769 43674.53 C4 H X2 Σ 4.09 4758.657 C6 H X2 Π 5.05 1391.186 C8 H X2 Π 6.94 587.264 C10 H C12 H C14 H X2 Π X2 Π X2 Π 301.410 174.784 110.242 astron (1974) lab (1981) astron (1978) lab (1983) lab (1988) astron (1986) lab (1996) astron (1996) lab (1998) lab (1998) lab (1998) Tucker, Kutner, Thaddeus [183] Sastry, Helminger [163] Guelin, Green [79] Gottlieb [66] Pearson, Gottlieb [146] Suzuki, Ohishi [172] McCarthy [119] Cernicharo [32] Gottlieb [64] Gottlieb [64] Gottlieb [64] A general overview is given in Takahashi [176] and Pauzat [144]. Isotopomers of CCH were examined by McCarthy et al. [120] and Saleck et al. [162]. Isotopic CCCCH was measured by Chen et al. [37]. 146 A Linear Cn H B The HQ Matrix Elements The matrix elements of the electric quadrupole interaction HQ (Eq.6.48) can be written in the Hund’s case (a) using 3j- and 6j-Symbols 1 −1 1 F J I I 2 I 0 0 0 J+I+F hηΛ SΣJ Ω IF |HQ |ηΛSΣJΩIF i = (−) 2 I J0 4 −I 0 I 0 J 2 J 0 1/2 J 0 −Ω ×[(2J + 1)(2J + 1)] δΛ0 Λ δΩ0 Ω eQq0 (−) −Ω 0 Ω X 0 0 J0 2 J + δΛ0 ,Λ∓2 (6)1/2 eQq2 (−)J −Ω (B.1) −Ω0 −q Ω q=±2 This equation can be split in two matrices, M (eQq0 ) and M (eQq2 ) −1 1 I 2 I F J I J+I+F (−) M (eQq0 ) = 2 I J0 4 −I 0 I 0 J 2 J 0 1/2 J 0 −Ω ×[(2J + 1)(2J + 1)] δΛ0 Λ δΩ0 Ω eQq0 (−) (B.2) −Ω 0 Ω −1 1 I 2 I F J I J+I+F M (eQq2 ) = (−) [(2J 0 + 1)(2J + 1)]1/2 2 I J0 4 −I 0 I X 0 J 2 J 1/2 J 0 −Ω0 × δΛ0 ,Λ∓2 (6) eQq2 (−) (B.3) −Ω0 −q Ω q=±2 Π states have Ω or Ω0 values of 3/2 or 1/2 so that Ω’=Ω ± 1. For the energy the ∆J = 0 elements are calculated using the basis functions 2 2 |2 Π± |Ω| , Ji = |Λ, Σ, J, Ωi ± | − Λ, −Σ, J, −Ωi √ 2 (B.4) with ± referring to the e/f parity respectively. To calculate the 3j- and 6j-symbols the following equalities are useful: 4J ≡ [(2J + 3)(2J + 2)(2J + 1)(2J)(2J − 1)]1/2 I ≡ [(2I + 3)(2I + 2)(2I + 1)(2I)(2I − 1)]1/2 1 2 see Brown & Schubert [27], Eq. 2 The matrix elements are in a non parity conserving basis. 148 B The HQ Matrix Elements X = −R(F ) ≡ J(J + 1) + I(I + 1) − F (F + 1) K(F ) ≡ 3R(F )[R(F )+1]−4J(J+1)I(I+1) I(2I−1)(2J+3)(2J+2)(2J)(2J−1) s ≡ F +J +I The following formula are taken or derived from Edmonds [51]. 2[3I 2 − I(I I 2 I = −I 0 I I 2 2[3Ω − J(J J 2 J = (−1)J−Ω −Ω 0 Ω 4J 2[3X (X − 1) − 4J(J + 1)I(I F J I = (−1)s 2 I J 4J I J 2 J J J 2J+2 = (−1) Ω −2 Ω − 1 Ω Ω−1 s (J + Ω − 1)(J − Ω + 2) J J J 2 J = 3 Ω Ω − 1 −2 Ω Ω−2 (2J + 3)(2J − 1) s (J + Ω − 2)(J − Ω + 3) J J J 1 J =− Ω Ω − 2 −1 Ω Ω − 3 (2J)(2J + 1) For Eq. B.9 3 + 1)] + 1)] + 1)] 2 −2 1 −1 0 0 (B.5) (B.6) (B.7) (B.8) (B.9) (B.10) and Eq. B.10 ([51], Eq. 3.7.13) was used. The diagonal matrix elements M (eQq0 ) can be calculated directly using Eq. B.5 - B.7 with J = J 0 and Ω = 1/2 for the lower left matrix element of Tab. B.1 and Ω = 3/2 for the upper right matrix element. In Eq. B.3 the index q of the sum separates the upper right and lower left matrix element of M (eQq2 ). HQ is hermitian and it is thus sufficient to calculate only one addend. 0 J 2 J ≡ Ai (B.11) −Ω0 −q Ω is non zero if (−Ω0 + (−q) + Ω) = 0, but because |q| = 2 and Ω0 , Ω{±1/2, ±3/2} this is only possible for (Ω0 = −3/2, q = +2, Ω = 1/2)1 or (Ω0 = 3/2, q = −2, Ω = −1/2)2 . Because of the δΛ0 ,Λ∓2 only A1 applies and M (eQq2 ) can be derived by setting Ω0 = −3/2 and Ω = 1/2 using Eq. B.8 with (−Ω0 (= −3/2)=Ω b = 3/2 and Ω(= 3/2)=Ω b − 1 = 1/2). Eq. B.8 can be solved by Eq. B.9 and B.10. In the case Ω = 3/2 it is J J 0 (Ω=3/2)! J J 0 = = (−1)J−Ω (2J + 1)−1/2 (B.12) Ω Ω−3 0 Ω −Ω 0 The result is summarized in Tab. B.1. 3 With j1 = j2 ≡ J, j3 = 2, m1 = Ω, m2 = Ω − 1, m3 = −2 to terms vanish because j3 + m3 = 0 149 Table B.1: Matrix with electr. hf interaction |2 Π± 3/2 JIF i h2 Π± 3/2 JIF |...... h2 Π± 1/2 JIF |...... with K(F ) = eQq0 27 2 K(F )[ 4 − J(J + 1)] (Hermitian) |2 Π± 1/2 JIF i 1 1 3 1/2 2 2 ± eQq 4 K(F )[(J − 4 )(J + 2 )(J + 2 )] eQq0 3 2 K(F )[ 4 − J(J + 1)] 3R(F )[R(F )+1]−4J(J+1)I(I+1) I(2I−1)(2J+3)(2J+2)(2J)(2J−1) R(F ) = F (F + 1) − J(J + 1) − I(I + 1) Matrix elements in non-parity conserving basis derived by Tom C. Killian and Guido Fuchs from Brown & Schubert, [27] 150 B The HQ Matrix Elements C Molecular Constants of C13CCN and CC13CN In Chapter 7.1.2 the measurements and analysis of the 13 C-C3 N isotopomers have been described. Tab. 7.4 shows the results of the global fits including the mm-data from McCarthy et al. [121] and the newly measured MW-data. The fit of the data has been done in 3 steps. First the mm-data were fitted with Herb Pickett’s spfit/spcat program. With the B and D constants contraint to the values of the mm-data fit a new fit with only MW-data was done. A final fit including all available mm- and MW data was done to derive the ’recommended values’ of the C3 N isotopomers. Tab. 7.3 shows the result of the intermediat steps of the fit for 13 CCCN. The tables for the intermediate results for C13 CCN and CC13 CN are given here. Table C.1: Molecular Constants of C13 CCN (in MHz). Data reduction was done with the Pickett-program. a Constant this workb mm-data onlyc recommendedd values B 4920.7107(8) 4920.712(2) 4920.7095(2) D ×10−3 0.78(4) 0.749(2) 0.7453(4) γ −18.62(2) −18.9(2) −18.574(5) γD ×10−3 2.4(8) 0.2(2) ... 13 bF ( C) 188.6(2) 210.(30) 188.6(2) c(13 C) 52.4(2) −40.(200) 52.9(1) 14 bF ( N) −1.244(7) −8.(200) −1.234(6) 14 c( N) 2.86(4) −40.(100) 2.82(3) eQq0 −4.(1) −30.(200) −4.331(9) w-rmse 1.35 0.54 1.16 a Uncertainties (in parentheses) are (1σ) in the last significant digit. b 13 lines were used, see Tab.7.7. The uncertainties of the lines is estimated to be 2 kHz. c 28 lines from [121] were used. The uncertainties of the lines are estimated to be between 22-86 kHz. d Total fit with all measured 41 lines. e w-rms is normalized with uncertainties of measured lines. 152 C Molecular Constants of C13 CCN and CC13 CN Table C.2: Molecular Constants of CC13 CN (in MHz). Data reduction was done with the Pickett-program. a Constant this workb mm-data onlyc recommendedd values B 4929.0640(5) 4929.0639(2) 4929.0640(2) −3 D ×10 −0.76(2) 0.7496(2) −0.7497(3) γ −18.643(7) −18.60(2) −18.648(3) γD ×10−3 −0.2(2) −0.02(1) ... 13 bF ( C) 23.54(3) 21.(6) 23.55(2) c(13 C) 2.19(4) 40.(100) 2.17(3) 14 e bF ( N) −1.184(8) −1.193 −1.182(8) c(14 N) 2.87(2) 2.837e 2.88(2) e eQq0 −4.32(1) −4.321 −4.323(8) f w-rms 0.79 1.00 0.77 a Uncertainties (in parentheses) are (1σ) in the last significant digit. b 32 lines were used, see Tab.7.8. The uncertainties of the lines is estimated to be 5 kHz. c 12 lines from [121] were used. The uncertainties of the lines are estimated to be between 22-86 kHz. d Total fit with all measured 44 lines. e fixed value. f w-rms is normalized with uncertainties of measured lines. D Tables: Interstellar C3N ,C5N, and C3N Isotopomers C3 N was first detected in gas phase with a radio telescope by Guélin and Thaddeus [78, 79] in 1977 towards IRC+10216. Further C3 N sources are IRAS 15194-5115, IRC +10216, TMC-1, TMC-2, HCL 2, CRL 618, CRL 2688, and it has also been observed in direction of Cas A. Table D.2 summerizes the astronomical measured transitions of C3 N. During a line survey towards IRC+10216 Cernicharo et al. [34] detected three 13 C mono-substituted C3 N isotopomers. In the case of 13 CCCN the assignement of two of the astronomical observed lines is not in agreement with the data derived in this thesis, see Table D.1. In 1998 Guelin et al. [77] detected C5 N in the dark cloud TMC-1. Up to now TMC-1 and IRC+10216 are the only sources in which C5 N has been detected, see Table D.3. Table D.1: Transitions of isotopic C3 N in IRC+10216, Cernicharo et al. [34]. The transitions with a question mark do not agree with the laboratory data of this work and a correct assignment is not possible. R isotope transition frequency Tmb dν (N,J,F1 ,F)”→(N,J,F1 ,F)’ [MHz] [K km/s] 13 CCCN (15,,,)→(14,,,)a ? (143 104.0) ? 0.71 (15,31/2,15,)→(14,31/2,14,) 143 124.0 0.92 (16,,,)→(15,,,)a ? (152 640.0) ? 0.62 (16,33/2,17,)→(15,31/2,17,) 152 659.7 1.26 C 13 CCN (14,29/2,14,)→(13,27/2,14,) 137 763.2 1.17 (14,27/2,[13],)→(13,25/2,[12],) 137 778.3 0.55 (15,31/2,[16],)→(14,31/2,[15],) 147 602.2 0.45 (15,29/2,14,)→(14,27/2,14,) 147 617.6 0.38 13 CC CN (14,29/2,15,)→(13,27/2,14,) 137 996.2 1.30 (14,27/2,[14],)→(13,25/2,[13],) 138 014.7 1.17 154 D Tables: Interstellar C3 N ,C5 N, and C3 N Isotopomers Table D.2: Astronomical detections of C3 N source CSE1 IRAS 15194-5115 IRC +10216 frequency (N,J,F)”→(N,J,F)’ [MHz] (11,23/2,)→(10,21/2,) (17,33/2,)→(16,31/2,) (17,35/2,)→(16,33/2,) (16,31/2,)→(15,29/2,) (16,33/2,)→(15,31/2,) (15,29/2,)→(14,27/2,) (15,31/2,)→(14,29/2,) (14,27/2,)→(13,25/2,) (14,29/2,)→(13,27/2,) (11,21/2,)→(10,19/2,) (11,23/2,)→(10,21/2,) (10,19/2,)→( 9,17/2,) (10,21/2,)→( 9,19/2,) (9,17/2,)→(8,15/2,) (9,19/2,)→(8,17/2,) (5,9/2,)→(4,7/2,) (5,11/2,)→(4,9/2,) (4,7/2,)→(3,5/2,) (4,9/2,)→(3,7/2,) (3,5/2,)→(2,3/2,) (3,7/2,)→(2,5/2,) 108 834.27 168 213.1 168 194.4 158 321.1 158 302.3 148 427.8 148 409.1 138 534.6 138 515.7 108 853.0 108 834.3 98 939.9 98 958.6 89 045.7 89 064.4 49 485.2 49 466.5 39 590.2 39 571.3 29 695.1 29 676.1 2.14 2.14 0.18 0.13 0.13 0.14 0.226 0.270 0.172 0.205 0.043 0.058 29 695.13 29 695.13 29 694.99 29 676.28 29 676.14 29 676.14 19 800.121 19 799.951 19 781.094 19 780.826 19 780.800 9 885.89 0.15 0.15 0.04 0.12 0.11 0.11 0.055 0.022 0.094 0.05 0.058 0.02 MC2 TMC-1 (3,5/2,7/2 (3,5/2,5/2 (3,5/2,3/2 (3,7/2,9/2 (3,7/2,7/2 (3,7/2,5/2 (2,3/2,5/2 (2,3/2,3/2 (2,5/2,7/2 (2,5/2,3/2 (2,5/2,5/2 (1,3/2,5/2 (continued on next page) 1 2 T∗A (Tmb ) [K] transition circumstellar envelopes molecular cloud )→(2,3/2,5/2) )→(2,3/2,3/2) )→(2,3/2,1/2) )→(2,5/2,7/2) )→(2,5/2,5/2) )→(2,5/2,3/2) )→(1,1/2,3/2) )→(1,1/2,1/2) )→(1,3/2,5/2) )→(1,3/2,1/2) )→(1,3/2,3/2) )→(0,1/2,3/2) R RTA dν ( Tmb dν) [K km s−1 ] (0.4) (10.90) (10.08) (24.00) (22.47) (27.97) (26.37) (27.89) (26.27) 5.58 6.82 4.01 5.04 1.13 1.69 reference [140] [34] [34] [34] [34] [34] [34] [34] [34] [79], [15] [79], [15] [78] [78] [78] [78] [101] [101] [101] [101] [101] [101] [56], [76] [56], [76] [56]?, [76] [56], [76] [56], [76] [56], [76] [76] [76] [76] [76] [76] [76] 155 (continued from previous page) source transition [MHz] 29 676.14 29 676.28 19 781.094 T∗A (Tmb ) [K] 0.06 0.07 - (,21/2,)→(,19/2,) (,19/2,)→(,17/2,) (9,17/2,)→(8,15/2,) (9,19/2,)→(8,17/2,) 89 045.7 89 064.4 (0.1) (0.13) 0.2 0.2 [28] [28] [139] [139] (1,3/2,)→(0,1/2,) 9 885 ? 0.01 [10] (N,J,F)”→(N,J,F)’ (3,7/2,7/2 )→(2,5/2,5/2) (3,7/2,9/2 )→(2,5/2,7/2) (2,5/2,7/2 )→(1,3/2,5/2) TMC-2 HCL 2 PPN3 CRL 618 CRL 2688 other molecules in direction Cas A frequency R RTA dν ( Tmb dν) [K km s−1 ] Table D.3: Transitions of C5 N in IRC+10216 and TMC-1, see [77] R source transition frequency Tmb dν (N,J)”→(N,J)’ [MHz] [mK km/s] TMC-1 (9,19/2)→(8,17/2) 25 249.938 7.3 (9,17/2)→(8,15/2) 25 260.649 6.4 IRC+10216 (32,65/2)→(31,63/2) 89 785.6 95 (32,63/2)→(31,61/2) 89 797.0 105 3 proto planetary nebulae reference [56] [56] [33] 156 D Tables: Interstellar C3 N ,C5 N, and C3 N Isotopomers Bibliography [1] W.S. Adams. Some Results with the Coude Spectrograph of the Mount Wilson Observatory. The Astrophysical Journal, 93:11, 1941. [2] A. Adel. The Distribution of Energy in the Violet CN Bands in the Spectra of Comets 1914b (Zlatinsky) and 1915a (Mellish). Publ.Astron.Soc.Pacific, 49:254, October 1937. [3] Michael D. Allen, Kenneth M. Evenson, David A. Gillet, and John M. Brown. Far-Infrared Laser Magnetic Resonance Spectroscopic Study of the ν2 Bending Fundamental of the CCN Radical in Its X̃ 2 Πr State . Journal of Molecular Spectroscopy, 201:18 – 29, 2000. [4] A.J. Apponi, M.E. Sanz, C.A. Gottlieb, M.C. McCarthy, and P. Thaddeus. The Cyclic C5 H Radical. The Astrophysical Journal, 547:L65, 2001. [5] P.W. Atkins. Physical Chemistry. Oxford University Press, 5th edition, 1994. [6] T.J. Balle and W.H. Flygare. Fabry-Perot cavity pulsed Fourier transform microwave spectrometer with a pulsed nozzle particle source. Rev.Sci.Instrum., 52(1):33, September 1981. [7] P.S. Bechtold and M. Neeb. B4: Struktur und elektronische Eigenschaften von Clustern, Vortrag zum 28. IFF Ferienkurs 1997 über die ”Dynamik und Strukturbildung in kondensierter Materie”. Script, Institut für Festkörperforschung Forschungszentrum Jülich GmbH, 1997. [8] M. B. Bell, P. A. Feldman, M. J. Travers, M. C. McCarthy, C. A. Gottlieb, and P. Thaddeus. Detection of HC11 N in the Cold Dust Cloud TMC-1. The Astrophysical Journal, 483:L61, July 1997. [9] M.B. Bell, P.A. Feldman, S. Kwok, and H.E. Matthews. Detection of HC11 N in IRC+10◦ 216. Nature, 295:389 – 391, 1982. [10] M.B. Bell and H.E. Matthews. Detection of C3 N in the Spiral Arm Gas Clouds in the Direction of Cassiopeia A. The Astrophysical Journal, 438:223 – 225, January 1995. 158 Bibliography [11] J. Berkowitz and W.A. Chupka. Mass spectrometric study of vapor ejected from graphite and other solids by focused laser beams. Journal of Chem. Phys., 40:2735, 1964. [12] Peter F. Bernath. Spectra of Atoms and Molecules. Oxford University Press, 1st edition, 1995. [13] Peter F. Bernath, Kenneth W. Hinkle, and John J. Keady. Detection of C5 in the Circumstellar Shell of IRC+10216. Science, 244:562–564, 1989. [14] Ute Berndt. Infrarot-Spektroskopie an kleinen Kohlenstoff-Clustern. Cuvillier Verlag, Göttingen, 2000. [15] John H. Bieging and Mario Tafalla. The Distribution of Molecules in the circumstellar Envelope of IRC+10216: HC3 N, C3 N, and SiS. Astron.Journal, 105:576 – 594, February 1993. [16] M. Bogey, C. Demuynck, and J.L. Destombes. The millimeter wave spectrum of the 13 C14 N radical in its ground state. Can.J.Phys., 62:1248 – 1253, 1984. [17] M. Bogey, C. Demuynck, and J.L. Destombes. The millimeter wave spectrum of 13 CN in the excited vibrational states ν ≤ 9. Chem.Phys., 102:141, 1986. [18] A. Borghesi and G. Guizetti. (c) Graphite, Handbook of Optical Constants of Solids II. Academic Press Inc., 1991. Ed.: Palik, E.D. [19] P. Botschwina. The two lowest electronic states of C5 N - results of coupled cluster calculations. Chem.Phys.Lett., 259 N5-6:627 – 634, Sept 1996. [20] P. Botschwina, M. Horn, K. Markey, and R. Oswald. Coupled cluster calculations for HC7 N, HC7 N+ , C7 N. Molecular Physics, 92 N3:381 – 392, Oct 1997. [21] G.D. Boyd and J.P. Gordon. Confocal multimode resonator for millimeter through optical wavelength masers. Bell Sys. Tech. J., 40:489 – 508, March 1961. [22] C. R. Brazier and J. M. Brown. The microwave spectrum of the CH free radical. Journal of Chem. Phys., 78:1608 – 1610, February 1983. [23] C.R. Brazier, L.C. O’Brien, and P.F. Bernath. Fourier transform detection of laser-induced fluorescence from the CCN free radical. Journal of Chem. Phys., 86:3078 – 3081, 1987. [24] J.M. Brown, E.A. Colbourn, J.K.G Watson, and F.D. Wayne. An Effective Hamiltonian for Diatomic Molecules. Journal of Molecular Spectroscopy, 74:294, 1979. [25] J.M. Brown, J.T. Hougen, K.-P. Huber, J.W.C. Johns, I. Kopp, H. LefebvreBrion, A.J. Merer, D.A. Ramsay, J. Rostas, and R.N. Zare. The Labeling of Parity Doublet Levels in Linear Molecules. Journal of Molecular Spectroscopy, 55:500 – 503, 1975. Bibliography 159 [26] J.M. Brown, M. Kaise, C.M.L. Kerr, and D.J. Milton. A determination of the fundamental Zeeman parameters for the OH radical. Molecular Physics, 36:553, 1978. [27] J.M. Brown and J.E. Schubert. The EPR Spectrum of the OD Radical: A Determination of the Molecular Parameters for the Ground State. Journal of Molecular Spectroscopy, 95:194, 1982. [28] V. Bujarrabal, J. Gómez-Gonzáles, R. Bachiller, and J. Martı́n-Pintado. Protoplanetary nebulae: the case of CRL 618. Astronomy and Astrophysics, 204:242– 252, March 1988. [29] E.J. Campbell, L.W. Buxton, T.J. Balle, M.R. Keenan, and W.H. Flygare. The gas dynamics of a pulsed supersonic nozzle molecular source as observed with a Fabry-Perot cavity microwave spectrometer. Journal of Chem. Phys., 74(2):829 – 840, January 1981. [30] A. Carrington and A.D. McLachlan. Introduction to Magnetic Resonance with Applications to Chemistry and Chemical Physics. Harper and Row, New York, 1967. [31] The cologne database for molecular spectroscopy, cdms, 2001. world wide web at http://www.cdms.de. [32] J. Cernicharo and M. Guelin. Discovery of the C8 H radical. Astronomy and Astrophysics, 309:L27–L30, May 1996. [33] J. Cernicharo, M. Guélin, and J. Askne. TMC1-like cloudlets in HCL2. Astronomy and Astrophysics, 138:371, 1984. [34] J. Cernicharo, M. Guélin, and C. Kahane. A λ2 mm molecular line survey of the C-star envelope IRC+10216. Astron.Astrophys.Suppl.Ser., 142:181 – 215, 2000. [35] J. Cernicharo, M. Guélin, and C.M. Walmsley. Detection of the hyperfine structure of the C5 H radical. Astronomy and Astrophysics, 172:L5 – L6, 1987. [36] J. Cernicharo, C. Kahane, J. Gómez-González, and M. Guélin. Tentative detection of the C5 H radical. Astronomy and Astrophysics, 164:L1 – L4, 1986. [37] W. Chen, S.E. Novick, M.C. McCarthy, C.A. Gottlieb, and P. Thaddeus. Carbon13 hyperfine structure of the CCCCH radical. Journal of Chem. Phys., 103:7828 – 7833, 1995. [38] Isabelle Cherchneff, Alfred E. Classgold, and Gary A. Mamon. The Formation of Cyanopolyyne Molecules in IRC+10216. The Astrophysical Journal, 410:188, June 1993. 160 Bibliography [39] A.C. Cheung, D.M. Rank, C.H. Townes, D.D. Thornton, and W.J. Welch. Detection of NH3 molecules in the interstellar medium by their microwave emission. Phys.Rev.Lett., 21:1701 – 1705, 1968. [40] A.C. Cheung, D.M. Rank, C.H. Townes, and W.J. Welch. Further microwave emission lines and clouds of ammonia in our galaxy. Nature, 221:917 – 919, 1969. [41] R.F. Curl and R.E. Smalley. Fullerenes. Scientific American (Int. Ed.), 265:32 – 41, October 1991. [42] P.B. Davies and I.H. Davis. Far infrared L.M.R. of X̃ 2 Π NCO. Molecular Physics, 69:175 – 191, 1990. [43] W. Demtröder and H.-J. Foth. Molekülspektroskopie in kalten Düsenstrahlen. Physikalische Blätter, 43:7 – 13, 1987. [44] R. H. Dicke, P. J. E. Peebles, P. G. Roll, and D. T. Wilkinson. Cosmic Black-Body Radiation. The Astrophysical Journal, 142:414 – 419, July 1965. [45] H. Ding, T. Pino, F. Güthe, and J.P. Maier. Gas phase electronic spectrum of C3 H in the visible. Journal of Chem. Phys., 115(15):6913 – 6919, October 2001. [46] Y. Ding, J. Liu, X. Huang, Z. Li, and C. Sun. C4 N: The first Cn N radical with stable cyclic isomers. Journal of Chem. Phys., 114:5172, 2001. [47] T. Dixon and R. Woods. The Laboratory Microwave Spectrum of the Cyanide Radical. Journal of Chem. Phys., 67:3956 – 3964, 1977. [48] Steven D. Doty and Chun Ming Leung. Detailed Chemical Modeling of the Circumstellar Envelopes of Carbon Stars: Application to IRC+10216. The Astrophysical Journal, 502:898 – 908, August 1998. [49] T. Dunham. Interstellar Neutral Potassium Publ.Astron.Soc.Pacific, 49:26 – 28, February 1937. and Neutral Calcium. [50] Geoffrey Duxbury. Infrared Vibrations - Rotation Spectroscopy. From Free Radicals to the Infrared Sky. Wiley-VCH, 2000. [51] A.R. Edmonds. Angular Momentum in Quantum Mechanics. Princeton Academic Press, 4th edition, 1996. [52] Friedrich Engelke. Aufbau der Moleküle. Teubner, Stuttgart, 3rd edition, 1996. [53] J. Drowart et al. Mass spectrometric study of carbon vapor. Journal of Chem. Phys., 31:1131, 1959. [54] K.L. Ramakumar et al. Carbon cluster formation in an rf-spark source . Int.J.Mass Spectrom.Ion.Phys., 75:171, 1987. Bibliography 161 [55] M. Feher, C. Salud, and J.P. Maier. The infrared laser spectrum of the ν1 band of CCN. Journal of Molecular Spectroscopy, 145:246 – 250, 1991. [56] P Friberg, Å Hjalmarson, W.M. Irvine, and M. Guélin. Interstellar C3 N: Detection in Taurus Dark Clouds. The Astrophysical Journal, 241:L99 – L103, October 1980. [57] R.A. Frosch and H.M. Foley. Magnetic Hyperfine Structure in Diatomic Molecules. Phys.Rev, 88:1337, 1952. [58] Guido W. Fuchs. Diplomarbeit. Charakterisierung einer Kohlenstoff-Cluster-Quelle, 1999. [59] T. F. Giesen, A. O. Van Orden, J. D. Cruzan, R. A. Provencal, R. J. Saykally, R. Gendriesch, F. Lewen, and G. Winnewisser. Interstellar Detection of CCC and High-Precision Laboratory Measurements near 2 THZ. The Astrophysical Journal, 551:L181–L184, April 2001. [60] Thomas Giesen. High Resolution Spectroscopy on Small Carbon Clusters, 2001. Habiltitationsschrift. [61] Thomas F. Giesen, Ute Berndt, M.T. Yamada, Guido Fuchs, Rudolf Schieder, Gisbert Winnewisser, Robert A. Provencal, Frank N. Keutsch, Alan Van Orden, and Richard J. Saykally. Detection of the Linear Carbon Cluster C10 : Rotationally Resolved Diode-Laser Spectroscopy. Chem.Phys.Chem, 4:242 – 247, April 2001. [62] Thomas F. Giesen, Alan Van Orden, H.J. Hwang, R.S. Fellers, Robert A. Provencal, and Richard J. Saykally. Infrared laser spectroscopy of the linear C13 carbon cluster. Science, 265:756 – 759, 1994. [63] Walter Gordy and Robert L. Cook. Microwave Molecular Spectra. John Wiley & Sons, 1984. [64] C. A. Gottlieb, M. C. McCarthy, M. J. Travers, J.-U. Grabow, and P. Thaddeus. Rotational spectra of the carbon chain free radicals C10 H, C12 H, C13 H, and C14 H. Journal of Chem. Phys., 109:5433–5438, October 1998. [65] C.A. Gottlieb, E.W. Gottlieb, and P. Thaddeus. Laboratory Detection of The C5 H Radical . Astronomy and Astrophysics, 164:L5 – L6, 1986. [66] C.A. Gottlieb, E.W. Gottlieb, P. Thaddeus, and H. Kawamura. Laboratory Detection of The C3 N and C4 H Free Radicals . The Astrophysical Journal, 275:916 – 921, December 1983. [67] C.A. Gottlieb, E.W. Gottlieb, P. Thaddeus, and J.M. Vrtilek. The Rotational Spectrum of the C3 H Radical. The Astrophysical Journal, 303:446, 1986. [68] C.A. Gottlieb, J.M. Vrtilek, E.W. Gottlieb, and P. Thaddeus. Laboratory Detection of The C3 H Radical . The Astrophysical Journal, 294:L55 – L58, July 1985. 162 Bibliography [69] J.-U. Grabow, N. Heineking, and W. Stahl. A Molecular Beam Microwave Fourier Transform (MB-MWFT) Spectrometer with an Electric Discharge Nozzle. Z.Naturforsch., 46a:914 – 916, 1991. [70] C.P. Grigoropoulos. Lasers, Optics and Thermal Considerations in Ablation Experiments, Experimental Methods in the Physical Sciences 30. Academis Press, San Diego, 1998. Ed.: J.C. Miller and R.F.Haglund. [71] Particle Data Group. Particle Physics Booklet. Springer, 2000. [72] M. Grutter, M. Wyss, and J.P. Maier. Electronic absorption spectra of C2n H − ,C2n−1 N − (n=4-7), and C2n−1 N (n=3-7) chains in neon matrices. Journal of Chem. Phys., 110:1492 – 1496, January 1999. [73] M. Guélin, S. Muller, J. Cernicharo, A. J. Apponi, M. C. McCarthy, C. A. Gottlieb, and P. Thaddeus. Astronomical detection of the free radical SiCN. Astronomy and Astrophysics, 363:L9–LL12, November 2000. [74] M. Guelin and J. Cernicharo. Astronomical detection of the HCCN radical Toward a new family of carbon-chain molecules? Astronomy and Astrophysics, 244:L21–L24, April 1991. [75] M. Guelin, J. Cernicharo, M. J. Travers, M. C. McCarthy, C. A. Gottlieb, P. Thaddeus, M. Ohishi, S. Saito, and S. Yamamoto. Detection of a new linear carbon chain radical: C7 H. Astronomy and Astrophysics, 317:L1–L4, 1997. [76] M. Guélin, P. Friberg, and A. Mezaoui. Astronomical Study of the C3 N and C4 H Radicals: Hyperfine Interactions and Rho-type Doubling. Astronomy and Astrophysics, 109:23 –31, 1982. [77] M. Guélin, N. Neininger, and J. Cernicharo. Astronomical Detection of the cyanobutadiynyl radical C5 N. Astronomy and Astrophysics, 335:L1 – L4, 1998. [78] M. Guélin and P. Thaddeus. Tentative Detection of The C3 N Radical . The Astrophysical Journal, 212:L81 – L85, March 1977. [79] M Guélin and P. Thaddeus. Detection of The C4 H Radical Toward IRC+10216 . The Astrophysical Journal, 224:L27 – L30, August 1978. [80] H. Haberland. Clusters of Atoms and Molecules 1. Springer, Berlin, Heidelberg, 1994. [81] T. I. Hasegawa and E. Herbst. Three-Phase Chemical Models of Dense Interstellar Clouds - Gas Dust Particle Mantles and Dust Particle Surfaces. Monthly Notice of the Royal Astronomical Society, 263:589, August 1993. [82] Gerhard Herzberg. Molecular Spectra and Molecular Structure II: Infrared and Raman Spectra of Polyatomic Molecules. Van Nostrand, Princeton, N.J., 1945. Bibliography 163 [83] Gerhard Herzberg. Molecular Spectra and Molecular Structure I: Spectra of Diatomic Molecules. Van Nostrand, Princeton, N.J., 2nd edition, 1950. [84] Gerhard Herzberg. The Spectra and Structure of Simple Free Radicals. Dover, New York, 1st edition, 1971. [85] Gerhard Herzberg. Einführung in die Molekülspektroskopie. Wissenschaftliche Forschungsberichte, Darmstadt, 1973. [86] Gerd Herziger. Lasertechnik. Script, RWTH Aachen, 1994. [87] Kenneth W. Hinkle, John J. Keady, and Peter F. Bernath. Detection of C3 in the Circumstellar Shell of IRC+10216. Science, 241:1319–1322, 1988. [88] T. Hirota, H. Ozawa, Y. Sekimoto, and S. Yamamoto. Microwave Spectrum of the Linear C5 D Radical. Journal of Molecular Spectroscopy, 174:196, 1995. [89] R.E. Honig. Mass spectrometric study of the molecular sublimation of graphite. Journal of Chem. Phys., 22:126, 1954. [90] K. Hoshina, H. Kohguchi, Y. Ohshima, and Y. Endo. Laser-induced fluorescence spectroscopy of the C4 H and C4 D radicals in a supersonic jet. Journal of Chem. Phys., 108:3465–3478, March 1998. [91] F. Hund. Zur Deutung einiger Erscheinungen in den Molekelspektren. Z.Physik, 36:657 – 674, 1926. [92] Minoru Iida, Yasuhiro Oshima, and Yasuki Endo. Laboratory detection of HC9 N using a fourier transform microwave spectrometer. The Astrophysical Journal, 371:L45 – L46, 1991. [93] H. Ito, K. Kuchitsu, S. Yamamoto, and S. Saito. Microwave Spectroscopy of the ν=3-10 levels of CN (X 2 Σ+ ). Chem.Phys.Lett., 186:539 – 546, 1991. [94] K. B. Jefferts, A. A. Penzias, and R. W. Wilson. Observation of the CN Radical in the Orion Nebula and W51. The Astrophysical Journal, 161:L87, August 1970. [95] K. B. Jefferts, A. A. Penzias, and R. W. Wilson. Observation of the CN Radical in the Orion Nebula and W51. The Astrophysical Journal, 161:L87, August 1970. [96] M.A. Johnson, N.L. Alexander, I. Hertel, and W.C. Lineberger. Improved flexibility in MODR using a supersonic jet source: applications to CO+ and CN. Chem.Phys.Lett., 105:374 – 379, 1984. [97] Judson. Detector Offerings, Judson Technologies, LLC, 221 Commerce Drive, Montgomeryville, PA 18936 USA. (2002), www.judtech.com. 164 Bibliography [98] Masao Kakimoto and Takahiro Kasuya. Doppler-Limited Dye Laser Excitation Spectroscopy of the CCN Radical. Journal of Molecular Spectroscopy, 94:380–392, 1982. [99] Y. Kasai, Y. Sumiyoshi, Y. Endo, and K. Kawaguchi. Laboratory Detection of the C5 N Radical by Fourier Transform Microwave Spectroscopy. The Astrophysical Journal, 477:L65 – L67, March 1997. [100] K Kawaguchi, T. Suzuki, S. Saito, E. Hirota, and T. Kasuya. Dye laser excitation spectroscopy of the CCN radical: the Ã2 ∆i - X̃Πr (0,1,0)-(0,1,0) and (0,2,0)-(0,2,0) bands. Journal of Molecular Spectroscopy, 106:320 – 329, 1984. [101] Kentarou Kawaguchi, Yasuko Kasai, Shin-ichi Ishikawa, and Norio Kaifu. A Spectral-Line Survey Observation of IRC+10216 between 28 and 50 GHz. Publ.Astron.Soc.Japan, 47:853 – 876, 1995. [102] Kentarou Kawaguchi, Shuji Saito, and Eizi Hirota. Microwave spectroscopy of the NCO radical in the ν2 =0 2 Π, ν2 =1 2 ∆, and ν2 =2 2 Φ vibronic states. Molecular Physics, 55(2):341 – 350, 1985. [103] E. Klisch, T. Klaus, S. P. Belov, G. Winnewisser, and E. Herbst. Laboratory rotational spectrum of CN in the 1 THz region. Astronomy and Astrophysics, 304:L5, December 1995. [104] Egbert Klisch. Rotationsspektroskopie freier Radikale. GCA-Verlag, Herdecke, 1999. [105] H Kogelnik and T. Li. Laser Beams and Resonators. Proc.IEEE, 54:1312 – 1329, 1966. [106] W.H. Kohl. Handbook of Materials and Techniques for Vacuum Devices. Reinhold Publisher, New York, 1967. [107] D. J. Krajnovich. Laser sputtering of highly oriented pyrolytic graphite at 248 nm. Journal of Chem. Phys., 102:726–743, January 1995. [108] H.W. Kroto, J.R. Heath, S.C. O’Brien, R.F. Curl, and R.E. Smalley. C60 : Buckminsterfullerene. Nature, 318:162 – 163, 1985. [109] L.D. Landau and E.M. Lifschitz. Quantenmechanik. Akademie Verlag, Berlin, 9th edition, 1979. [110] H. Linnartz, T. Motylewski, and J. P. Maier. The 2 Π ← X 2 Π electronic spectra of C8 H and C10 H in the gas phase. Journal of Chem. Phys., 109:3819–3823, September 1998. [111] H. Linnartz, T. Motylewski, F. Maiwald, D.A. Roth, F. Lewen, I. Pak, and G. Winnewisser. Millimeter wave spectroscopy in a pulsed supersonic slit nozzle discharge. Chem.Phys.Lett., 292:188 – 192, 1998. Bibliography 165 [112] H. Linnartz, D. Pfluger, O. Vaizert, P. Cias, P. Birza, D. Khoroshev, and J. P. Maier. Rotationally resolved A2 Πu ← X 2 Πg electronic transition of NC6 N+ . Journal of Chem. Phys., 116:924–927, January 2002. 3 − [113] H. Linnartz, O. Vaizert, T. Motylewski, and J. P. Maier. The 3 Σ− u ← X Σg electronic spectrum of linear C4 in the gas phase. Journal of Chem. Phys., 112:9777– 9779, June 2000. [114] G. H. MacDonald, A. G. Gibb, R. J. Habing, and T. J. Millar. A 330360 GHz spectral survey of G 34.3+0.15. I. Data and physical analysis. Astron.Astrophys.Suppl.Ser., 119:333–367, October 1996. [115] J.M.L. Martin, J.P. Francois, and R. Gijbels. Ab initio study of the infrared spectra of linear Cn clusters (n=6-9). Journal of Chem. Phys., 93:8850–8861, 1990. [116] J.M.L. Martin, Peter R. Taylor, J.P. Francois, and R. Gijbels. Ab initio study of the spectroscopy and thermochemistry of the C2 N and CN2 molecules. Chem.Phys.Lett., 226:475–483, August 1994. [117] N. Mauron and P.J. Huggins. Multiple shells in the circumstellar envelope of IRC+10216. Astronomy and Astrophysics, 349:203, 1999. [118] M. C. McCarthy, W. Chen, M. J. Travers, and P. Thaddeus. Microwave Spectra of 11 Polyyne Carbon Chains. Astrophys.J.Suppl.Ser., 129:611–623, August 2000. [119] M. C. McCarthy, M. J. Travers, A. Kovacs, C. A. Gottlieb, and P. Thaddeus. Laboratory detection of the C8 H radical. Astronomy and Astrophysics, 309:L31 – L33, May 1996. [120] M.C. McCarthy, C.A. Gottlieb, and P. Thaddeus. Rotational spectrum and hyperfine structure of 13 CCH and C13 CH. Journal of Molecular Spectroscopy, 173:303, 1995. [121] M.C. McCarthy, C.A. Gottlieb, P. Thaddeus, M. Horn, and P. Botschwina. Structure of the CCCN and CCCCH radicals: Isotopic substitution and ab initio theory. Journal of Chem. Phys., 103:7820 – 7827, 1995. [122] M.C. McCarthy, M.J. Travers, A. Kovács, C.A. Gottlieb, and P. Thaddeus. Eight new carbon chain molecules. Astrophys.J.Suppl.Ser., 113:105, 1997. [123] Michael C. McCarthy and Patrick Thaddeus. Microwave and laser spectroscopy of carbon chains and rings. Chemical Society Reviews, 30:177 – 185, 2001. [124] G.M. McClelland, K.L. Saenger, J.J. Valentini, and D.R. Herschbach. Vibrational and Rotational Relaxation of Iodine in Seeded Supersonic Beams. J.Phys.Chem., 83:947 – 959, 1979. 166 Bibliography [125] A. McKellar. Evidence for the Molecular Origin of Some Hitherto Unidentified Interstellar Lines. Publ.Astron.Soc.Pacific, 52:187 – 192, 1940. [126] A. McKellar. The Structure of the λ3883 CN Band in the Spectrum of Comet Cunningham (1940 c). Publ.Astron.Soc.Pacific, 53:235, August 1941. [127] A. M. Mebel and R. I. Kaiser. The Formation of Interstellar C2 N Isomers in Circumstellar Envelopes of Carbon Stars: An Ab Initio Study. The Astrophysical Journal, 564:787–791, January 2002. [128] A.J. Merer and D.N. Travis. Absorption Spectrum of the CCN Radical. Can.J.Phys., 43:1795 – 1830, 1965. [129] A.J. Merer and D.N. Travis. The Absorption Spectrum of CNC. Can.J.Phys., 44:353, 1966. [130] H. Mikami, S. Yamamoto, S. Saito, and M. Guélin. Laboratory microwave spectroscopy of the C3 N radical in the vibrationally excited state ν5 . Astronomy and Astrophysics, 217:L5, 1989. [131] T.J. Millar and Eric Herbst. A new chemical model of the circumstellar envelope surrounding IRC+10216. Astronomy and Astrophysics, 288:561 – 571, 1994. [132] T.J. Millar, Eric Herbst, and R.P.A. Bettens. Large molecules in the envelope surrounding IRC+10216. Monthly Notice of the Royal Astronomical Society, 316:195 – 203, 2000. [133] G. Monninger. Optische Spektroskopie von matrix-isolierten Kohlenstoff-Clustern. PhD thesis, Ruprecht-Karls-Universität, Heidelberg, 1995. [134] Andreas Moravec. Anharmonische Resonanzen von Cyanacetylen. PhD thesis, I.Physik, Universität zu Köln, 1994. [135] T. Motylewski and H. Linnartz. Cavity ring down spectroscopy on radicals in a supersonic slit nozzle discharge . Rev.Sci.Instrum., 70:1305–1312, 1999. [136] T. Motylewski, O. Vaizert, T. F. Giesen, H. Linnartz, and J. P. Maier. The 1 Πu ← X 1 Σ+ g electronic spectrum of C5 in the gas phase. Journal of Chem. Phys., 111:6161–6163, October 1999. [137] S. Murahashi, T. Takizawa, S. Kurioka, and S. Maekawa. Nippon Kagaku Zasshi, 77:1689, 1956. [138] P.T. Murray and D.T. Peeler. Pulsed Laser Deposition of Carbon Films: Dependence of Film Properties on Laser Wavelength. Journal of Electronic Materials, 23:855 – 859, 1994. Bibliography 167 [139] Nguyen-Quang-Rieu, S. Deguchi, H. Izumiura, N. Kaifu, M. Ohishi, H. Suzuki, and Nobuharu Ukita. A Sensitive Line Search in Circumstellar Envelopes. The Astrophysical Journal, 330:374 – 384, July 1988. [140] L.-Å. Nyman, H. Olofsson, L.E.B. Johansson, R.S. Booth, U. Carlström, and R. Wolstencroft. A molecular radio line survey of the carbon star IRAS151945115. Astronomy and Astrophysics, 269:377 – 389, 1993. [141] Y. Ohshima and Y. Endo. Structure of C3 S studied by pulsed-discharge-nozzle fourier-transform microwave spectroscopy. Journal of Molecular Spectroscopy, 153:627 – 634, 1992. [142] Y. Ohshima and Y. Endo. Fourier-Transform Microwsave Spectroscopy of CCN (X2 Π1/2 ). Journal of Molecular Spectroscopy, 172:225 – 232, 1995. [143] N. Oliphant, A. Lee, P.F. Bernath, and C.R. Brazier. Fourier transform emission spectroscopy of the jet-cooled CCN free radical. Journal of Chem. Phys., 92:2244 – 2247, 1990. [144] F. Pauzat, Ellinger Y., and A.D. McLean. Is Interstellar Detection of Higher Members of the Linear Radicals Cn CH and Cn N Feasible? The Astrophysical Journal, 369:L13 – L16, March 1991. [145] R. Pd and P. Chandra. Ground and valence excited states of C2 N and CN2 transients: Ab initio geometries, electronic structures, and molecular properties. Journal of Chem. Phys., 114:1589–1600, January 2001. [146] J.C. Pearson, C.A. Gottlieb, D.R. Woodward, and P. Thaddeus. Laboratory detection of the C6 H radical. Astronomy and Astrophysics, 189:L13 – L15, 1988. [147] A. A. Penzias, K. B. Jefferts, and R. W. Wilson. Interstellar CN Excitation at 2.64 mm. Phys.Rev.Lett., 28:772–775, March 1972. [148] A. A. Penzias and R. W. Wilson. A Measurement of Excess Antenna Temperature at 4080 Mc/s. The Astrophysical Journal, 142:419–421, July 1965. [149] A. A. Penzias, R. W. Wilson, and K. B. Jefferts. Hyperfine Structure of the CN Radical Determined from Astronomical Observations. Phys.Rev.Lett., 32:701–703, April 1974. [150] M. Perı́c, B. Engels, and S.D. Peyerimhoff. Ab initio investigation of the vibronic structure of the C2 H spectrum: calculation of the hyperfine coupling constants for the three lowest-lying electronic states. Journal of Molecular Spectroscopy, 150:56 – 69, 1991. [151] M. Perı́c, B. Engels, and S.D. Peyerimhoff. Ab initio investigation of the vibronic structure of the C2 H spectrum: computation of the vibronically avaraged values for the hyperfine coupling constants. Journal of Molecular Spectroscopy, 150:70 – 85, November 1991. 168 Bibliography [152] Herb Pickett. Cataloge. world wide web at http://spec.jpl.nasa.gov. [153] H.M. Pickett. The Fitting and Prediction of Vibration-Rotation Spectra with Spin Interactions. Journal of Molecular Spectroscopy, 148:371 – 377, 1991. [154] H.M. Pickett, R.L. Poynter, E.A. Cohen, Delitsky M.L., J.C. Pearson, and H.S.P. Müller. Submillimeter, Millimeter, and Microwave Spectral Line Catalog. J.Quant.Spectrosc.Radiat.Transfer, 60:883 – 890, 1998. [155] Hugh .O. Pierson. Handbook of Carbon, Graphite, Diamond and Fullerenes. Noyes Publications, Park Ridge, New Jersey, U.S.A., 1993. Ed.: H.O. Pierson. [156] T. Pino, H. Ding, F. Güthe, and J. P. Maier. Electronic spectra of the chains HC2n H (n = 8-13) in the gas phase. Journal of Chem. Phys., 114:2208–2212, feb 2001. [157] J. Pliva, A. Valentin, L. Henry, F. Muller, and A. Bauder. Cyclopropane-d6 : High Resolution Study of the Infrared Bands ν9 , ν10 , ν11 , and the Pure Rotational Spectrum Measured by Fourier Transform Microwave Spectroscopy. Journal of Molecular Spectroscopy, 168:442–454, 1994. [158] Pd. Rajendra and P. Chandra. Ground and valence excited states of C2 N and CN2 transients: Ab intio geometries, electronic structures, and molecular properties. Journal of Chem. Phys., 114:1589 – 1600, January 2001. [159] Joanna Sadlej and Björn O. Roos. A CASSCF-MRCI study of the ground and lower excited states of the CCCN radical. Chem.Phys.Lett., 180:81 – 87, 1991. [160] A. H. Saleck, R. Simon, N. Schneider, and G. Winnewisser. Detection of interstellar 12 15 C N. The Astrophysical Journal Letter, 414:L133–L136, September 1993. [161] A. H. Saleck, R. Simon, and G. Winnewisser. Interstellar CN rotational spectra: (12)C(15)N. The Astrophysical Journal, 436:176–182, November 1994. [162] A. H. Saleck, R. Simon, G. Winnewisser, and J.G.A. Wouterloot. Detection of interstellar 13 CCH and C13 CH. Can.J.Phys., 72:747, 1994. [163] K.V.L.N Sastry, Paul Helminger, Arthur Charo, Eric Herbst, and Frank C. Lucia. Laboratory Millimeter and Submillimeter Spectrum of CCH. The Astrophysical Journal, 251:L119 – L120, 1981. [164] R. Schlachta, G. Lask, S.H. Tsay, and V.E. Bondybey. Pulsed discharge source of supersonically cooled transient species. Chem.Phys., 155:267, 1991. [165] Giacinto Scoles. Atomic and Molecular Beam Methods, Vol. 1. Oxford Universtity Press, 1st edition, 1988. [166] L.N. Shen and W.R.M. Graham. Observation of an infrared frequency of the C4 molecule. Journal of Chem. Phys., 91:5115 – 5116, 1989. Bibliography 169 [167] D.D. Skatrud, F.C. deLucia, G.A. Blake, and K.V.L.N Sastry. The millimeter and submillimeter spectrum of CN in its first four vibrational states. Journal of Molecular Spectroscopy, 99:35 – 46, 1983. [168] P.S. Skell, J.J. Havel, and M.J. McGlinchey. Chemistry and the carbon arc. Acc.Chem.Res., 6:97, 1973. [169] R.E. Smalley. Self-assembly of the fullerenes. Acc.Chem.Res., 25:98 – 105, March 1992. [170] T.C. Steimle, D.R. Woodward, and J.M. Brown. The lambda-doubling spectrum of 13 CH, studied by microwave optical double resonance. Journal of Chem. Phys., 85:1276, 1986. [171] V. Storm, H. Dreizler, D. Consalvo, J.-U. Grabow, and I. Merke. A newly designed molecular beam Fourier transform microwave spectrometer in the range 1-4 GHz. Rev.Sci.Instrum., 67(8):2714 – 2719, August 1996. [172] H. Suzuki, M. Ohishi, and N. Kaifu. Detection of the Interstellar C6 H Radical. Publ.Astron.Soc.Japan, 38:911 – 917, 1986. [173] T. Suzuki, S. Saito, and E. Hirota. Hyperfine coupling constants of the CCN radical in the Ã2 ∆(000) state by microwave-optical double resonance spectroscopy. Journal of Chem. Phys., 83:6154 – 6157, 1985. [174] P. Swings and L. Rosenfeld. Considerations Regarding Interstellar Molecules. The Astrophysical Journal, 86:483–486, November 1937. [175] J. Szcezepanski, R. Pellow, and M. Vala. The kinetics and formation of small carbon clusters in an argon matrix. Z.Naturforsch., 47a:595, 1992. [176] Junko Takahashi. Ab Initio Calculations for Detectable New Isomers of Interstellar Carbon-Chain Radicals Cn H (n=2-8). Publ.Astron.Soc.Japan, 52:401 – 407, 2000. [177] P. Thaddeus and J. F. Clauser. Cosmic Microwave Radiation at 2.63 mm from Observations of Interstellar CN . Phys.Rev.Lett., 16:819–822, May 1966. [178] P. Thaddeus, C.A. Gottlieb, Å. Hjalmarson, L.E.B. Johansson, W.M. Irvine, P. Friberg, and R.A. Linke. Astronomical Indentification of the C3 H Radical. The Astrophysical Journal, 294:L49 – L53, 1985. [179] P. Thaddeus and M.C. McCarthy. Carbon chains and rings in the laboratory and in space. Spectro.Chem.Acta Part A, 57:757, 2001. [180] R. Thomson and F.W. Dalby. Experimental determination of the dipole moments of the X(2 Σ+ ) and B(2 Σ+ ) states of the CN molecule. Can.J.Phys., 46:2815 – 2820, 1968. 170 Bibliography [181] C.H. Townes and A.L. Schawlow. Microwave Spectroscopy. Dover, 1975. [182] M.J. Travers, M.C. McCarthy, C.A. Gottlieb, and P. Thaddeus. Laboratory Detection of the C7 H Radical. The Astrophysical Journal, 465:L77 – L80, 1996. [183] K.D. Tucker, M.L. Kutner, and P. Thaddeus. The Ethynyl Radical C2 H- A new interstellar molecule. The Astrophysical Journal, 193:L115 – L119, 1974. [184] B. E. Turner and R. H. Gammon. Interstellar CN at radio wavelengths. The Astrophysical Journal, 198:71–89, May 1975. [185] B. E. Turner and B. Zuckerman. Microwave Detection of Interstellar CH. BAAS, 5:420, September 1973. [186] A. Van Orden. Direct Infrared Laser Absorbtion Spectroscopy of Jet-Cooled Carbon and Silicon-Carbon Clusters. PhD thesis, Berkeley, CA, U.S.A., 1996. [187] A. van Orden, T. F. Giesen, R. A. Provencal, H. J. Hwang, and R. J. Saykally. Characterization of silicon-carbon clusters by infrared laser spectroscopy: The ν3 (σu ) band of linear Si2 C3 . Journal of Chem. Phys., 101:10237–10241, December 1994. [188] A. van Orden, R. A. Provencal, T. F. Giesen, and R. J. Saykally. Characterization of silicon-carbon clusters by infrared laser spectroscopy: The ν1 band of SiC4 . Chem.Phys.Lett., 237:77 – 80, 1995. [189] A. Van Orden and R. J. Saykally. Small Carbon Clusters: Spectroscopy, Structure, and Energetics. Chem.Rev., 98:2313–2358, 1998. [190] S. L. Wang, C. M. L. Rittby, and W. R. M. Graham. Detection of cyclic carbon clusters. II. Isotopic study of the ν12 (eu ) mode of cyclic C8 in solid Ar. Journal of Chem. Phys., 107:7025–7033, November 1997. [191] S.L. Wang, C.M.L. Rittby, and W.R.M. Graham. Detection of cyclic carbon clusters. I. Isotopic study of the ν4 (el ) mode of cyclic C6 in solid Ar. Journal of Chem. Phys., 107:6032–6037, 1997. [192] S. Weinreb, A.H. Barrett, M.L. Meeks, and J.C. Henry. Radio Observations of the OH in the interstellar Medium. Nature, 200: no. 4909, 829 – 831, November 1963. [193] W. Weltner, P.N. Walsh, and C.L. Angell. Spectroscopy of carbon vapor condensed in rare-gas matrices at 4◦ and 20◦ K. Journal of Chem. Phys., 40:1299, 1964. [194] G. Winnewisser, T. Drascher, T. Giesen, I. Pak, F. Schmülling, and R. Schieder. The tunable diode laser: A versatile spectroscopic tool. Spectro.Chem.Acta Part A, 55:2121–2142, 1999. [195] Gisbert Winnewisser and Eric Herbst. Interstellar molecules. Rep.Prog.Phys., 56:1209 – 1273, 1993. Bibliography 171 [196] Friedrich Wyrowski. Radio observations of high mass star forming regions: hot cores and photon dominated regions. PhD thesis, I.Physikalisches Institut, Köln, 1997. [197] Fulin Xiong, Y.Y. Wang, V. Leppert, and R.P.H. Chang. Pulsed laser deposition of amorphous diamond-like carbon films with ArF (193 nm) excimer laser. J.Mater.Res., 8:2265 – 2272, 1993. [198] P.D. Zavitsanos and G.A. Carlson. Experimental study of the sublimation at high temperatures. Journal of Chem. Phys., 59:2966, 1973. 172 Bibliography List of Figures 2.1 2.14 Spectra of C2 at 516 nm obtained by using two different types of molecular sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C60 mass spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pressure-temperature diagram of graphite . . . . . . . . . . . . . . . . . The Cologne Laser Ablation Source . . . . . . . . . . . . . . . . . . . . . Jet produced by excimer laser ablation technique. . . . . . . . . . . . . . QMS experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . . Mass spectrum of Nd:YAG laser ablated graphite rod . . . . . . . . . . . Excimer laser ablation on different kinds of material. . . . . . . . . . . . The effect of an Nd:YAG laser beam on different kinds of material. . . . Discharge slit nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Jets produced by a discharge slit nozzle. . . . . . . . . . . . . . . . . . . Experimental setup for mass spectrometry on a molecular beam . . . . . Energy distribution of a He+ produced in the ion source of the plasma monitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mass spectra of a molecular beam . . . . . . . . . . . . . . . . . . . . . 3.1 3.2 3.3 3.4 The Cologne carbon cluster experiment . . . Sensitivity of IR detectors . . . . . . . . . . Frequency calibration . . . . . . . . . . . . . Rovibrational transition of C3 at 2067 cm−1 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 Block diagram of the FTMW low band system, 5 - 25 GHz . . . . . . . . The Production of Cn N radicals . . . . . . . . . . . . . . . . . . . . . . . Nozzle for the production of molecules and radicals . . . . . . . . . . . . Test nozzle to optimize the geometry of the discharge nozzle . . . . . . . The radical nozzle during a discharge . . . . . . . . . . . . . . . . . . . . The supersonic jet expansion . . . . . . . . . . . . . . . . . . . . . . . . . The long nozzles during a discharge . . . . . . . . . . . . . . . . . . . . . Continuum free-jet expansion . . . . . . . . . . . . . . . . . . . . . . . . Mach number and Temperature along the centerline axis of a free expansion Model of flow development . . . . . . . . . . . . . . . . . . . . . . . . . Theoretical intensities of the C4 N rotational transitions . . . . . . . . . . Cavity mode Lorentzian line shape . . . . . . . . . . . . . . . . . . . . . 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11 2.12 2.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 11 13 14 15 17 20 23 24 25 26 27 28 29 34 35 36 38 43 46 47 48 50 50 51 53 56 57 62 65 174 List of Figures 4.13 Measured Qef f and calculated Qth for the Fabry-Perot cavity . . . . . . . 4.14 Time and Frequency Domains . . . . . . . . . . . . . . . . . . . . . . . . 66 68 5.1 5.2 5.3 Schematic diagram of electron configuration of CN and CCN . . . . . . . Calculated geometries of C2 N . . . . . . . . . . . . . . . . . . . . . . . . Theoretical geometries of C4 N . . . . . . . . . . . . . . . . . . . . . . . . 71 74 75 6.1 Hypothetical energy level diagram of a 2 Π radical . . . . . . . . . . . . . 6.2 Vector diagram of Hund’s coupling cases a) and b) . . . . . . . . . . . . 6.3 Unpaired electron distribution for a 2 Π state in Hund’s case (b). . . . . . 78 84 90 Measured CCC15 N transition with Zeeman splitting . . . Energy level diagram of CCC15 N . . . . . . . . . . . . . Measured 13 CCCN transitions. . . . . . . . . . . . . . . . Hund case bβs and bβJ with 2 nuclear spins . . . . . . . 13 CCCN energy level scheme . . . . . . . . . . . . . . . . Resonance structure of CCCN . . . . . . . . . . . . . . . bF and c-values of different isoelectronic carbon chains . C4 N stick spectrum . . . . . . . . . . . . . . . . . . . . . Measured C4 N transition . . . . . . . . . . . . . . . . . . Measured transition of C6 N . . . . . . . . . . . . . . . . Hund’s case a with one nuclear spin . . . . . . . . . . . . Relative abundances of the Cn N radicals per gas pulse in molecular beam as a function of chain length. . . . . . . 7.13 Molecular constants h1/r3 i, hsin2 θ/r3 i, and hψ 2 (0)i . . . 7.14 Schematic C4 N energy level. . . . . . . . . . . . . . . . . 7.1 7.2 7.3 7.4 7.5 7.6 7.7 7.8 7.9 7.10 7.11 7.12 8.1 8.2 8.3 8.4 8.5 8.6 8.7 8.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . the supersonic . . . . . . . . . . . . . . . . . . . . . . . . . . . The molecular clouds in the plane of the Milky Way . . . . . . . . 13 CCCN detection in IRC+10216 . . . . . . . . . . . . . . . . . . Major pathways to cyanopolyyne formation in IRC+10216 . . . . Column densities of Cn N, Cn H, and HCn N in IRC+10216 . . . . . Possible C2 N transition towards IRC+10216 at 248 GHz . . . . . Possible C2 N transitions towards IRC+10216 at 154 and 224 GHz Boltzmann plot for C2 N . . . . . . . . . . . . . . . . . . . . . . . The IRAM 30m telescope at Pico Veleta, Granada . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 102 103 104 105 107 109 112 114 114 115 117 118 119 128 129 130 133 134 135 139 141 List of Tables 1.1 Known Interstellar and Circumstellar Molecules (Dec 2002) . . . . . . . . 5 2.1 2.2 2.3 Particle numbers of C3 , C9 and C13 using excimer laser ablation . . . . . Technical data of applied ablation lasers . . . . . . . . . . . . . . . . . . Relative C-cluster concentration of different production techniques . . . . 16 19 21 4.1 Free jet flow properties for Ne . . . . . . . . . . . . . . . . . . . . . . . . 59 5.1 5.2 Cn N, n odd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cn N, n even . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 76 6.1 6.2 6.3 6.4 6.5 6.6 6.7 Selection of important interactions and their constants . Angular momenta and their projections . . . . . . . . . . Theoretical Λ-type doubling for 2 Π state . . . . . . . . . Matrix with Spin-Orbit, Rotation and Λ-doubling . . . . Matrix with electr. hf interaction . . . . . . . . . . . . . Matrix with magnetic hyperfine interaction . . . . . . . . Transformed matrix with magnetic hyperfine interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 84 87 94 95 96 96 7.1 7.2 7.3 7.4 7.5 7.6 7.7 7.8 7.9 7.10 7.11 7.12 Measured Rotational Transitions of CCC15 N in the X 2 Σ+ State Molecular Constants of CCC15 N . . . . . . . . . . . . . . . . . . Molecular Constants of 13 CCCN . . . . . . . . . . . . . . . . . . Spectroscopic constants of the 13 C isotopic CCCN species. . . . bF (13 C) and c(13 C) values . . . . . . . . . . . . . . . . . . . . . Measured Rotational Transitions of 13 CCCN in the X 2 Σ+ State. Measured Rotational Transitions of C13 CCN in the X 2 Σ+ State. Measured Rotational Transitions of CC13 CN in the X 2 Σ+ State. Spectroscopic Constants of C4 N and C6 N in the X 2 Π State. . . Hyperfine and Molecular Constants of Carbon Chain Radicals . Measured Rotational Transitions C4 N in the X 2 Π1/2 State. . . . Measured Rotational Transitions C6 N, HUNDA-fit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 100 106 106 108 110 110 111 116 120 121 122 8.1 8.2 8.3 Molecular column densities in IRC+10216 . . . . . . . . . . . . . . . . . 132 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 . . . . . . . . . . . . . . . . . . . . . 176 List of Tables A.1 Linear Cn H (n=1-8) radicals . . . . . . . . . . . . . . . . . . . . . . . . . 145 B.1 Matrix with electr. hf interaction . . . . . . . . . . . . . . . . . . . . . . 149 C.1 Molecular Constants of C13 CCN . . . . . . . . . . . . . . . . . . . . . . . 151 C.2 Molecular Constants of CC13 CN . . . . . . . . . . . . . . . . . . . . . . . 152 D.1 Transitions of isotopic C3 N in IRC+10216 . . . . . . . . . . . . . . . . . 153 D.2 Astronomical detections of C3 N . . . . . . . . . . . . . . . . . . . . . . . 154 D.3 Transitions of C5 N in IRC+10216 and TMC-1 . . . . . . . . . . . . . . . 155 Acknowledgments There are many people who have made my graduate career at the Universität zu Köln and the Harvard-Smithsonian Institution a successful and rewarding time for me. Surely the first to mention here is Prof. Gisbert Winnewisser. It has been a pleasure to have him as a thesis advisor. He was always interested in the progress of my work and he cared about my questions and problems. His connection to Prof. Patrick Thaddeus gave me the possibility for an important and memorable time as a researcher at the Harvard Smithsonian Institution. And back at home I felt that his door was always open for me. Thanks for that! Prof. Patrick Thaddeus gave me the opportunity to work in his research group, and I learned a lot from him since he is a brilliant scientist. As my research advisor at the Harvard Smithsonian Institution, he really took excellent care of me which I appreciated so much and for which I want to thank him a lot. Especially his charismatic TuesdayWednesday meetings were always very motivating and inspiring for me. I thank Dr. Thomas Giesen for being my lab advisor. Even before the beginning of my Diplom thesis he provided me with the day to day help, advised me and got me on the road towards molecular spectroscopy. I felt really at home in his work group and never felt any lack of support. His patience in answering clever and sometimes not at all clever questions is remarkable. I wish him all the best for his future career as a first grade scientist. Dr. Michael C. McCarthy did a really great job as my lab advisor in Harvard. Not only is he an excellent scientist but he also is capable of making the work in his group very pleasant. I just loved following the daily lunch conversations with Carl which were always interesting, informative, and delightful. Also his first day crash course in the lovely FTMW spectrometer will always keep in my mind. Carl Gottlieb was one of my big oracles - whenever I had a question he had an answer. It was very nice to work together with him and I still see him in front of me writing formulas on the white board and make me think that we were working on really important things. Complementary to that his wife Elaine Gottlieb gave me support in Fortran to get some numbers out of Carl’s theories. Both of them gave me a warm welcome and good travel advises, Acadia is great! I also profited a lot by the experience of Sam Palmer whose knowledge about the experiments and especially the radio techniques is just incredible and who is a person who always seemed to be in a relaxed mood. His interest in teaching and the rest of the outside world brought up a lot of interesting discussions. 178 Acknowledgments I also want to thank the rest of the Thaddeus group for creating such a good atmosphere in which to work. First of all I want to mention Maria E. Sanz who is an excellent merengue dancer and who’s Spanish is just to lovely to listen to. She helped me a lot in getting started with the experiment. Need a good tip for a restaurant or pub in Cambridge or Boston? Call John Dudeck. But don’t make the mistake and bet a beer for who wins the next Tour de France! He supported me with his self-made HC3 N precursors and consequently was indispensable for the success of this project. Jane Kucera is one of the most diligent first year students I know and it was a pleasure to work with her. I hope that her future career will continue in that successful manner. Tom Dame was helpful in many ways, e.g. working on data to optimize the discharge nozzle dimensions and supporting me with figures and maps of the Milky Way. Dr. Rolf Berger provided the quadrupole mass spectrometer and was invaluable in getting it started. Thanks to Dr. Ute Berndt, Petra Neubauer-Guenther, and Michael Caris which were great lab-mates. Especially Petra helped me whenever she could in terms of getting the lab supply organized or by advising ’our’ molecular physics students. I hope that for both of the doctores-to-be, Michael and Petra, the work in Lab 320 will be interesting and that they will have a rich scientific harvest. I also want to thank Frank Schlöder, Dr. Frank Schmülling, and Michael Olbrich for their computer support. Especially Frank Schlöder was very helpful with his patience concerning my Linux, LATEX and dada problems. Friedrich Wyrowski helped me when ever I struggled with astrophysical problems. Thanks! Sandra Brünken, Patrick Pütz, Jörg Stodolka, Guido Sonnabend, and Daniel Wirtz are great comrades and I think we had a lot of fun working together in the I. Institut. The machine shop did a great job and never ran out of solutions. Thanks! No doubt, Thomas Giesen and Katja Roth did a great job in correcting this thesis. Thanks! My family has always given me their total support. I would like to thank my parents, Agnes and Werner Fuchs, my brother Tobias and sister Sabine, as well as Christel and Josef Feldt who have always provided me with every opportunity. My aunt Elisabeth and uncle Wilhelm guided me through all steps of my education - thanks. Special thanks go to my wife Uli. Both of us studied at the same time, same place and nearly on the same topic: The pitfalls of physics and molecules. There is no other person to whom I owe such high esteem than her who during many years of scientific battle never stopped hoping that it all will have an happy end. This work was supported by the Deutsche Forschungsgemeinschaft and the Smithsonian Institution. Ich versichere, daß ich die von mir vorgelegte Dissertation selbstständig angefertigt, die benutzten Quellen und Hilfsmittel vollständig angegeben und die Stellen der Arbeit - einschließlich Tabellen, Karten und Abbildungen -, die anderen Werken im Wortlaut oder dem Sinn nach entnommen sind, in jedem Einzelfall als Entlehnung kenntlich gemacht habe; daß dieser Dissertation noch keiner anderen Fakultät oder Universiät zur Prüfung vorgelegen hat; daß sie abgesehen von unten angegebenen Teilpublikationen noch nicht veröffentlicht worden ist, sowie daß ich eine solche Veröffentlichung vor Abschluß des Promotionsverfahrens nicht vornehmen werde. Die Bestimmung der Promotionsordnung sind mir bekannt. Die von mir vorgelegte Dissertation ist von Herrn Prof. Dr. G. Winnewisser betreut worden. (Guido W. Fuchs) Parts of this thesis are published in: 1. M.C. McCarthy, G.W. Fuchs, J. Kucera, G. Winnewisser, and P. Thaddeus, Rotational Spectra of C4 N, C6 N, and the Isotopic Species of C3 N, Journal of Chemical Physics, 118, 3549 - 3557 (2003) Publication List: 1. T.F. Giesen, U. Berndt, K.M.T. Yamada, G. Fuchs, R. Schieder, G. Winnewisser, R.A. Provencal, F.N. Keutsch, A. Van Orden, and R.J. Saykally, Detection of the Linear Carbon Cluster C10 : Rotationally Resolved Diode-Laser Spectroscopy, ChemPhysChem 2, 242-247 (2001) 2. P. Neubauer-Guenther, T. F. Giesen, U. Berndt, G. Fuchs, and G. Winnewisser The Cologne Carbon Cluster Experiment: Ro-Vibrational spectroscopy on C8 and other small carbon clusters Spectro.Chem.Acta Part A, 59/3, 431 - 441 (2003) Curriculum Vitae Personal representation: Name: Date of birth: Marital status: Fuchs, Guido Wilhelm 14th December, 1971, Polch Germany married Education: 1978 - 1982 1982 - 1988 1988 - 1991 Grundschule Kehrig Realschule Mayen Kurfürst-Balduin-Gymnasium Münstermaifeld Community service: 03/1992 - 04/1993 Arbeiter-Samariter-Bund (ASB) LV Köln Studies: 10/91 - now Universität zu Köln: Subject: Physics Vordiplom Diplom: ”Charakterisierung einer Kohlenstoff-Cluster-Quelle” 10/94 01/1999 Studies in foreign countries: 02/1996 - 11/1996 03/2001 - 9/2001 Activities: 03/1995 - 04/1995 03/1996 - 11/1996 05/1998 - 02/1999 02/1999 - 02/2001 12/1999 - 06/2000 9/2001 - now Köln, 15th May 2003 University of Cape Town theor. physics Termination: B.Sc. (Honours) Harvard-Smithsonian Institution, Center for Astrophysics (FTMW research on reactive molecules) ”Miniforschung” at the Universität zu Köln (I.PI) section for receiver technology tutor for ”first year students” at the University of Cape Town student assistant at the I.PI scientific assistant at the I.IP lecturer at the IFBM Cologne (Institut für Biologie und Medizin) scientific co-worker at the I. IP
© Copyright 2025 Paperzz