You are on page 1of 992

Acquisitions

Editor: Brian Brown Product Development Editor: Nicole Dernoski Editorial Assistant:
Lindsay Burgess Production Project Manager: Priscilla Crater Design Coordinator: Holly McLaughlin
Manufacturing Coordinator: Beth Welsh Marketing Manager: Daniel Dressler Prepress Vendor: Absolute
Service, Inc.

2nd Edition
Copyright © 2015 Wolters Kluwer Health

1st Edition Copyright © 2010 Lippincott Williams & Wilkins, a Wolters Kluwer business. All rights
reserved. This book is protected by copyright. No part of this book may be reproduced or transmitted in any
form or by any means, including as photocopies or scanned-in or other electronic copies, or utilized by any
information storage and retrieval system without written permission from the copyright owner, except for
brief quotations embodied in critical articles and reviews. Materials appearing in this book prepared by
individuals as part of their official duties as U.S. government employees are not covered by the above-
mentioned copyright. To request permission, please contact Wolters Kluwer Health at Two Commerce
Square, 2001 Market Street, Philadelphia, PA 19103, via email at permissions@lww.com, or via our website
at lww.com (products and services).

9 8 7 6 5 4 3 2 1

Printed in China

Library of Congress Cataloging-in-Publication Data Ultrasound-guided regional anesthesia and pain


medicine / editors, Paul E. Bigeleisen, Michael Gofeld, Steven L. Orebaugh ; associate editors, Annelot
Krediet, Gerbrand J. Groen, Stephen Breneman, Bassem Asaad, and Jacques E. Chelly. — 2nd edition.
p. ; cm.
Includes bibliographical references and index.
ISBN 978-1-4511-7333-8 (hardback : alk. paper) I. Bigeleisen, Paul E., editor.
[DNLM: 1. Anesthesia, Conduction—methods—Atlases. 2. Pain Management—Atlases. 3.
Ultrasonography—methods—Atlases. WO 517]
RD84
617.9′64—dc23

2014047977
This work is provided “as is,” and the publisher disclaims any and all warranties, express or implied,
including any warranties as to accuracy, comprehensiveness, or currency of the content of this work.

This work is no substitute for individual patient assessment based upon health care professionals’
examination of each patient and consideration of, among other things, age, weight, gender, current or prior
medical conditions, medication history, laboratory data, and other factors unique to the patient. The
publisher does not provide medical advice or guidance and this work is merely a reference tool. Health care
professionals, and not the publisher, are solely responsible for the use of this work including all medical
judgments and for any resulting diagnosis and treatments.

Given continuous, rapid advances in medical science and health information, independent professional
verification of medical diagnoses, indications, appropriate pharmaceutical selections and dosages, and
treatment options should be made and health care professionals should consult a variety of sources. When
prescribing medication, health care professionals are advised to consult the product information sheet (the
manufacturer’s package insert) accompanying each drug to verify, among other things, conditions of use,
warnings and side effects, and identify any changes in dosage schedule or contradictions, particularly if the
medication to be administered is new, infrequently used, or has a narrow therapeutic range. To the
maximum extent permitted under applicable law, no responsibility is assumed by the publisher for any
injury and/or damage to persons or property, as a matter of products liability, negligence law or otherwise,
or from any reference to or use by any person of this work.

LWW.com
Contributors

Bassem Asaad, MB ChB, ABA, ABAPM, ABPM, FIPP


Assistant Professor of Clinical Anesthesiology Department of Anesthesiology
Stony Brook University
Stony Brook, New York

Manasi Badve, MD
Consultant Anesthesiologist Department of Anesthesiology and Pain Medicine P.
D. Hinduja National Hospital and Medical Research Center Mumbai, India

Vikram Bansal, MD
Fellow
Department of Anesthesiology University of Pittsburgh
Pittsburgh, Pennsylvania

Todd Beery, DO
Physiatrist
Interventional Spine and Sports Medicine Middlebury, Connecticut

David Belavy, FANZCA Senior Lecturer


Burns, Trauma and Critical Care Research Centre The University of Queensland
Staff Anaesthetist
Department of Anesthesiology and Perioperative Medicine Royal Brisbane and
Women’s Hospital Brisbane, Australia

Ralph Beltran, MD
Pediatric Anesthesiologist Department of Anesthesiology and Pain Medicine
Nationwide Children’s Hospital Clinical Assistant Professor The Ohio State
University Medical Center Columbus, Ohio

Alon Ben-Ari, MD
Visiting Instructor
Department of Anesthesiology University of Pittsburgh
Attending Staff Anesthesiologist Department of Anesthesiology UPMC-
Presbyterian
Pittsburgh, Pennsylvania

Diana L. Besleaga, MD
Anesthesia Resident
Department of Anesthesiology Stony Brook University
Stony Brook, New York

Tarun Bhalla, MD
Associate Professor
Department of Anesthesiology and Pain Medicine The Ohio State University
Wexner Medical Center Pediatric Anesthesiologist Director of Acute Pain and
Regional Anesthesia Department of Anesthesiology and Pain Medicine
Nationwide Children’s Hospital Columbus, Ohio

Rafael Blanco, MD
Consultant Anaesthesiologist Anaesthetic Department
University Hospital of Lewisham London, United Kingdom

Andre P. Boezaart, MD, PhD


Professor of Anesthesiology and Orthopaedic Surgery Departments of
Anesthesiology and Orthopaedic Surgery Division of Acute and Perioperative
Pain Medicine University of Florida College of Medicine Gainesville, Florida

Karen Boretsky, MD
Assistant Professor of Anesthesiology University of Pittsburgh School of
Medicine Director
Acute Perioperative Pain Service Director
Resident International Rotation Children’s Hospital of Pittsburgh of UPMC
Pittsburgh, Pennsylvania

Stephen M. Breneman, MD, PhD


Associate Professor
Department of Anesthesiology University of Rochester
Chief of Acute Pain Service Department of Anesthesiology Strong Memorial
Hospital
Rochester, New York

Michael N. Brown, MD
Pain Medicine Fellow
Department of Anesthesiology and Pain Medicine University of Washington
University of Washington Medical Center Seattle, Washington

Richard Brull, MD, FRCPC


Associate Professor
Department of Anesthesia
University of Toronto
Site Chief
Department of Anesthesia
Women’s College Hospital
Toronto, Ontario, Canada

David Burns, MD
Division Chief
Regional Anesthesia, Acute Pain Providence Sacred Heart Medical Center
Spokane, Washington

Arvind Chandrakantan, MD, FAAP


Assistant Professor of Anesthesiology Anesthesiology Department Stony Brook
University Medical Center Stony Brook, New York

Phillip C. Cory, MD
Cory Biomedical Consulting Bozeman, Montana

Laurent Delaunay, MD
Département d’anesthésiologie Clinique Générale
Annecy, France

Brian Durkin, DO
Director
Center for Pain Management Associate Professor of Anesthesiology Stony
Brook Medicine
Stony Brook, New York

Urs Eichenberger, MD
Privatdozent
Department of Anesthesiology and Pain Therapy University of Bern
Attending Physician
Department of Anesthesiology and Pain Therapy University Hospital of Bern
Bern, Switzerland

Patrik Filip, MD
Attending/Staff Anesthesiologist Department of Anesthesiology West Penn
Allegheny—Allegheny General Hospital Pittsburgh, Pennsylvania

Andrea Fanelli, MD
Visiting Professor
Department of Anesthesiology University of Pittsburgh
Fellow in Acute Pain Service Department of Anesthesiology UPMC Shadyside
Hospital
Pittsburgh, Pennsylvania

Yoshihiro Fujiwara, MD, PhD


Professor
Department of Anesthesiology Aichi Medical University
Nagakute, Aichi, Japan

Samuel M. Galvagno Jr., DO, PhD


Associate Professor
Department of Anesthesia
University of Maryland School of Medicine Chief
Division of Critical Care Medicine Department of Anesthesia
University of Maryland Medical Center/R Adams Cowley Shock Trauma Center
Baltimore, Maryland

Daniela Elena Francesca Ghisi, MD


Visiting Professor
Department of Anesthesiology University of Pittsburgh
Fellow in Acute Pain Service Department of Anesthesiology UPMC Shadyside
Hospital
Pittsburgh, Pennsylvania

Michael Gofeld, MD, DEAA, FIPP


Assistant Professor
Department of Anesthesiology and Pain Medicine Assistant Professor
Department of Neurological Surgery Center for Pain Relief
University of Washington Medical Center Seattle, Washington

Gerbrand J. Groen, MD, PhD


Associate Professor
Department of Anesthesiology University Medical Center Groningen Groningen,
The Netherlands Ryan Guffey, MD
Fellow
Department of Anesthesia
University of Pittsburgh Medical Center Pittsburgh, Pennsylvania

Jooyeon Ha, MFA 2D/3D Artist


Pittsford, New York
www.jooyeonha.com

Jason D. Hanks, MD
Department of Anesthesiology University of Pittsburgh Medical Center Regional
Fellow
Department of Anesthesiology University of Pittsburgh Medical Center
Pittsburgh, Pennsylvania

Kaoru Hara, MD
Anesthesia Deparment
Matsue Seikyo General Hospital Matsue, Japan

Jean-Louis Horn, MD
Professor—Med Center Line Department of Anesthesia
Stanford University
Chief
Division of Regional Anesthesia Department of Anesthesia
The Stanford University Medical Center Stanford, California

Joseph C. Hung, MD
Clinical Fellow
Anesthesia, Critical Care, and Pain Medicine Harvard Medical School
Massachusetts General Hospital Boston, Massachusetts

Jeremy Kaplowitz, MD
Assistant Professor
Department of Anesthesiology University of Maryland School of Medicine
Baltimore, Maryland

Kevin King, DO
Clinical Assistant Professor Department of Anesthesiology University of
Pittsburgh School of Medicine Staff Anesthesiologist
Department of Anesthesiology University of Pittsburgh Medical Center
Pittsburgh, Pennsylvania

Lavinia Kolarczyk, MD
Assistant Professor
Department of Anesthesiology University of North Carolina Chapel Hill, North
Carolina Toru Komatsu, MD, PhD
Professor Emeritus
Department of Anesthesiology Aichi Medical University
Aichi Medical University Hospital Nagakute, Aichi, Japan

Daniel L. Krashin, MD
Pain Fellow
Department of Anesthesiology and Pain Medicine University of Washington
Seattle, Washington

Annelot Krediet, MD, MSc PhD Student


Anesthesiology Department University Medical Center Utrecht Utrecht, The
Netherlands

Robert Scott Lang, MD


Clinical Assistant Professor Department of Anesthesiology University of
Pittsburgh Medical Center Pittsburgh, Pennsylvania

Aaron Lange, MD
Assistant Professor
Department of Anesthesiology University of Maryland
University of Maryland Medical Center Baltimore, Maryland

Imanuel R. Lerman, MD, MS


Assistant Professor
Anesthesiology Center for Pain Medicine University of California at San Diego
Assistant Professor
Anesthesiology Department Thornton Hospital
San Diego, California

Stephen Lucas, MD
Clinical Assistant Professor Department of Anesthesiology State University of
New York at Buffalo Buffalo, New York
Managing Partner
Department of Anesthesiology SCI Anesthesia, PLLC
Victor, New York

Alan J. R. Macfarlane, BSc (Hons),


MBChB (Hons), MRCP, FRCA
Hospital Honorary Clinical Senior Lecturer Anaesthesia, Pain Management and
Critical Care Department University of Glasgow
Consultant Anaesthetist
Department of Anaesthesia Glasgow Royal Infirmary
Glasgow, Scotland, United Kingdom David P. Martin, MD
Anesthesiologist
Department of Anesthesia and Pain Medicine Nationwide Children’s Hospital
Columbus, Ohio

Nizar Moayeri, MD, PhD


Assistant Professor
Neurosurgical Resident
Department of Neurosurgery University Medical Center Utrecht Utrecht, The
Netherlands

Hanni E. Monroe, MD
Assistant Professor
Department of Anesthesiology University of Maryland School of Medicine
Baltimore, Maryland

Kacey A. Montgomery, MD
Fellow, Pain Medicine
Department of Anesthesiology and Pain Medicine University of Washington
Seattle, Washington

Elilary Montilla Medrano, MD


Acute Pain, Regional and Orthopedics Anesthesiology Fellow
Department of Anesthesiology University of Pittsburgh Medical Center
Pittsburgh, Pennsylvania

Milena Moreno, MD
Instructor in Anesthesiology Department of Anesthesiology Pontificia Javeriana
University Anesthesiologist
Department of Anesthesiology San Ignacio Hospital
Bogota, Columbia, South America Hadi S. Moten, MD, MS
Resident Physician
Department of Anesthesiology Stony Brook University Medical Center Stony
Brook, New York

Arvind Murthy, MD
Director
Regional Anesthesia
Department of Anesthesiology Holy Family Hospital
Methuen, Massachusetts

James J. Nicholson, MD, MHS


Assistant Professor
Orthopedic Surgery
Stony Brook Medical Center Stony Brook, New York

Kristin Odenko Ligda, MD


Assistant Professor of Anesthesiology Department of Anesthesiology University
of Pittsburgh Medical Center Pittsburgh, Pennsylvania

Steven L. Orebaugh, MD
Associate Professor
Anesthesiology and Critical Care Medicine University of Pittsburgh School of
Medicine Pittsburgh, Pennsylvania

Philip Wenn-Hsin Peng, MBBS, FRCPC


Professor
Department of Anesthesia and Pain Management University of Toronto
Director
Anesthesia Chronic Pain Program University Health Network Toronto, Ontario,
Canada

Karl D. Reiling, PhD


Assistant Clinical Professor Cellular and Developmental Biology University of
Colorado School of Medicine Director
Simulation Department
Touch of Life Technologies Aurora, Colorado

Meg A. Rosenblatt, MD
Professor
Anesthesiology and Orthopaedics Mount Sinai School of Medicine New York,
New York

Alex Rosioreanu, MD
Musculoskeletal Radiologist Zwanger-Pesiri Radiology
Lindenhurst, New York

Kim Russon, MRCP, FRCA Consultant Anaesthetist


Anaesthetic Department
Rotherham Foundation Trust Hospital Rotherham, England, United Kingdom
Shinichi Sakura, MD
Associate Professor
Department of Anesthesiology Shimane University
Director
Surgical Center
Shimane University Hospital Izumo City, Japan

Ron Samet, MD
Assistant Professor
Department of Anesthesiology Director
Acute Pain and Regional Anesthesia Service Staff
Division of Trauma Anesthesiology R. Adams Cowley Shock Trauma Center
University of Maryland School of Medicine Baltimore, Maryland

Joshua Sappenfield, MD
Assistant Professor
Department of Anesthesiology University of Florida
Gainesville, Florida

Rachael Seib, MD, MA Assistant Professor


Department of Anesthesiology Queens University
Staff Anesthesiologist
Department of Anesthesiology Kingston General Hospital Kingston, Ontario,
Canada Vihang D. Shah, MD
Acute Pain and Regional Anesthesiology Fellow Department of Anesthesiology
University of Pittsburgh Medical Center Pittsburgh, Pennsylvania

Roger E. Shere-Wolfe, MD, MA, JD


Department of Anesthesiology University of Maryland School of Medicine
Baltimore, Maryland

Nigam Sheth, MD
Assistant Professor
Department of Anesthesiology University of Maryland Medical Center Director
of Resident and Student Education Department of Anesthesiology Department
of Veterans Affairs Baltimore, Maryland

Yasuyuki Shibata, MD
Assistant Professor
Department of Anesthesiology Aichi Medical University
Aichi-gun, Aichi
Assistant Professor
Department of Anesthesiology Nagoya University Hospital Nagoya, Aichi,
Japan

Daniella Smith, MD
Assistant Professor
Department of Anesthesia
University of Maryland School of Medicine University of Maryland Medical
Center Baltimore, Maryland

Iwan Sofjan, MD
Resident
Department of Anesthesiology University of Pittsburgh Medical Center
Pittsburgh, Pennsylvania

Dmitri Souzdalnitski, MD, PhD


Medical Director, Pain Management South Pointe Hospital, Cleveland Clinic
Cleveland, Ohio

Sunathenam Suresh, MD
Professor and Chief
Department of Pediatric Anesthesiology Ann & Robert H. Lurie Children’s
Hospital of Chicago Chicago, Illinois

Ahmad Mouhammad Taha, MD, MSc Lecturer


Department of Anesthesia
Ain Shams University
Ain Shams University (Demerdash) Hospitals Cairo, Egypt

Jonathan M. Tan, MD, MPH


Resident Physician
Department of Anesthesiology Stony Brook University Medical Center Stony
Brook, New York

Shruthima Thangada, MD
Instructor
Department of Anesthesiology New York University Langone Medical Center
New York, New York

Joseph D. Tobias, MD
Professor
Departments of Anesthesiology and Pediatrics The Ohio State University
Chairman
Department of Anesthesiology and Pain Medicine Nationwide Children’s
Hospital Columbus, Ohio

Manuel C. Vallejo, MD, DMD


Professor and Chair
Department of Anesthesiology West Virginia University
Morgantown, West Virginia Mihaela Visoiu, MD
Assistant Professor of Anesthesiology Department of Anesthesiology University
of Pittsburgh
Pediatric Anesthesiologist Pediatric Anesthesiology
Children’s Hospital of Pittsburgh Pittsburgh, Pennsylvania

Brian A. Williams, MD, MBA Professor


Department of Anesthesiology University of Pittsburgh School of Medicine
Director of Ambulatory and Regional Anesthesia Anesthesia Service
VA Pittsburgh Healthcare System Pittsburgh, Pennsylvania

Sylvia Wilson, MD
Assistant Professor
Chief
Regional Anesthesia Pain Service Rotation Director
Orthopedic and Regional Anesthesia Department of Anesthesiology and
Perioperative Medicine Medical University of South Carolina Charleston,
South Carolina Richard Zhu, MD
Resident
Department of Anesthesiology Yale University School of Medicine Yale-New
Haven Hospital
New Haven, Connecticut
Preface to the First Edition

began this atlas in Pittsburgh in 1988 with cadaver dissections, which I


I preserved in formalin tanks. In 1992, I began some simple ultrasound studies
of the brachial plexus. By 1997, I was plastinating additional cadaver dissections
at the University of Rochester and making simple line drawings of my
dissections. Jooyeon Ha, Sara Bednarz, and Chris Bickel joined me from the
departments of animation and medical illustration at Rochester Institute of
Technology. Together, we created professional illustrations, animations, and
three-dimensional virtual reality reconstructions. In 2002, I began ultrasound-
guided nerve blocks. In 2006, I returned to Pittsburgh and was joined by Nizar
Moayeri, who, along with Gerbrand Groen (University of Utrecht), had created a
large archive of microanatomical dissections of the brachial, thoracic, and
lumbar plexuses. In Pittsburgh, Jaques Chelly assigned some outstanding fellows
(Alon Ben Ari, Melina Moreno, and David Burns) to work with me on new
techniques in ultrasound-guided blocks. At the same time, Dr. Chelly introduced
me to the editors at LWW and the formal foundation for this atlas was begun. C.
C. Li, James Chien, Steven Damelin, and Jungha An contributed new ideas in
image processing. Phil Cory and Stuart Grant added new ideas in radiofrequency
imaging and ultrasound. Assad Oberei, Robert Borcala, and Yun Jing added new
insights into needle design. Steve Orebaugh wrote many of the chapters. Other
collaborators from around the world graciously joined us. Steven Breneman and
Meg Stanbury edited many of the chapters and images. All together, my
colleagues and I have spent more than 8,000 hours working on these dissections,
drawings, animations, and researches. The Fulbright Foundation, the National
Institute of Health, the National Science Foundation, Sleeping Gorilla, the
University of Pittsburgh, SonoSite, B Braun, Life Tech, and the Manipulative
Therapy Foundation of Holland contributed more than 1 million dollars in
support. I hope you enjoy the atlas.

Paul E. Bigeleisen
Summer 2009, Rochester, NY and Pittsburgh, PA
Preface to the Second Edition

he first edition of Ultrasound-Guided Regional Anesthesia and Pain


T Medicine was a success when it was released in 2010. In general, the
graphics and ultrasound images were better than in similar atlases produced at
that time. Nonetheless, our Atlas was out of date within 2 years of its publication
because of the rapid expansion of ultrasound into the diagnosis and performance
of procedures for chronic pain conditions. In addition, the diagnosis and
treatment of many other medical conditions with ultrasound guidance has found
its way into the practice of anesthesiologists. For these reasons, I asked Michael
Gofeld to join me as an editor on this edition for his expertise in the practice of
ultrasound-guided musculoskeletal blocks. I also asked Sam Galvagno to write
an introduction to other uses of ultrasound in the treatment of shock and trauma.
I would also like to thank the many new authors who have contributed 40 new
chapters to this edition. I would especially like to thank: Annelot Krediet, who
was finally able to illustrate and elucidate the anatomy of the paravertebral
space.
Michelle Odonkor and Bassem Asaad, who contributed many editorial
suggestions.
Karen Boretsky, who really made the chapter on pediatric epidural and
paravertebral blocks come to life.
Jooyoen Ha, who continued to provide extraordinary graphic support.
Steph Sadler and Susie Hancock, who created and edited many images.
Finally, both of my parents have passed away since the publication of the first
edition. I am dedicating this edition to Grace and Jacob Bigeleisen, both of
whom provided me with my spark of curiosity and wonder in life.
Contents

PREFACE TO THE FIRST EDITION

PREFACE TO THE SECOND EDITION

INTRODUCTION Ultrasound in Trauma and Critical Care


Samuel M. Galvagno Jr and Joshua Sappenfield

SECTION I FUNDAMENTALS

1. Dermatomes, Sclerotomes, and Planes


Paul E. Bigeleisen and Jooyeon Ha
2. Pharmacology of Perineural Analgesia
Brian A. Williams, Nigam Sheth, Sylvia Wilson, and Lavinia Kolarczyk
3. The Physics of Nerve Stimulation and Nerve Impedance
Philip C. Cory
4. Microanatomy of the Peripheral Nervous System and Stimulation
ThresholdsInside and Outside the Epineurium
Part 1. Microanatomy
Andre P. Boezaart, Paul E. Bigeleisen, and Nizar Moayeri

Part 2. Nerve Stimulation Thresholds


Ron Samet and Paul E. Bigeleisen

5. Ultrasound Equipment
Steven L. Orebaugh and Paul E. Bigeleisen
6. Principles of Sonography
Paul E. Bigeleisen, Daniella Smith, and Steven L. Orebaugh
7. Three-and Four-Dimensional Ultrasound in Neuraxial Anesthesia
David Belavy
8. Anatomical Anomalies in Ultrasound Simulation
Jeremy Kaplowitz and Paul E. Bigeleisen
9. Ultrasound Simulator Training
Paul E. Bigeleisen, Eric Stavnitsky, and Karl Reiling
10. Advanced Ultrasound-Guided Needle Technology
Steven L. Orebaugh, Arvind Murthy, Kristin Odenko Ligda, Richard Zhu,
Paul E. Bigeleisen, and Iwan Sofjan
11. From Paresthesia to Neurostimulation and Ultrasound-Guided Regional
Anesthesia
Sylvia Wilson and Jacques E. Chelly
12. Understanding Needle-to-Nerve Proximity in Peripheral Nerve Blocks
Alan J. R. Macfarlane and Richard Brull
13. Ultrasound-Guided Intraneural Injection—The Human Data
Meg A. Rosenblatt and Paul E. Bigeleisen
14. Use of Ultrasound for Placement of Peripheral Nerve Block Catheters
Robert Scott Lang, Paul E. Bigeleisen, and Steven L. Orebaugh

SECTION II UPPER EXTREMITY

15. Ultrasound-Guided Interscalene Block Using the Posterior Approach


David Byrnes and Patrik Filip
16. Ultrasound-Guided Interscalene Block
Steven L. Orebaugh, Ryan Guffey, Nizar Moayeri, and Paul E. Bigeleisen
17. Ultrasound-Guided Supraclavicular Block
Steven L. Orebaugh and Paul E. Bigeleisen
18. Ultrasound-Guided Suprascapular and Axillary Nerve Block
Part 1. Suprascapular Nerve Block
Bassem Asaad, Jonathan M. Tan, and Paul E. Bigeleisen

Part 2. Ultrasound-Guided Selective Axillary Nerve Block


Aaron Lange and Paul E. Bigeleisen

19. Ultrasound-Guided Infraclavicular Block


Steven L. Orebaugh, Gerbrand J. Groen, Elilary Montilla Medrano, and
Paul E. Bigeleisen
20. Ultrasound-Guided Axillary Block
Steven L. Orebaugh, Jason D. Hanks, and Paul E. Bigeleisen
21. Ultrasound-Guided Blocks at the Elbow and Forearm
Steven L. Orebaugh, Shruthima Thangada, and Paul E. Bigeleisen

SECTION LOWER EXTREMITY


III

22. Ultrasound-Guided Lumbar Plexus Block (Transverse Approach)


Shinichi Sakura, Kaoru Hara, and Jean-Louis Horn
23. Ultrasound-Guided Femoral Nerve Block
Stephen M. Breneman
24. Ultrasound-Guided Lateral Femoral Cutaneous Block
Paul E. Bigeleisen, Milena Moreno, and Steven L. Orebaugh
25. Ultrasound-Guided Saphenous Nerve Block of the Thigh, Knee, and
Ankle
Part 1. Saphenous Nerve Block of the Thigh
Steven L. Orebaugh, Vihang D. Shah, Milena Moreno, and Paul E. Bigeleisen

Part 2. Ultrasound-Guided Saphenous Nerve of the Knee (Infrapatellar)


and Ankle
Diana L. Besleaga, Bassem Asaad, Alex Rosioreanu, Stephen M. Breneman, and James J.
Nicholson

26. Ultrasound-Guided Proximal Obturator Block: A Proximal Interfascial


Technique
Ahmaad Muhammd Taha
27. Ultrasound-Guided Distal Obturator Nerve Block
Yoshihiro Fujiwara and Toru Komatsu
28. Ultrasound-Guided Proximal Parasacral Block
Kevin King and Jacques E. Chelly
29. Ultrasound-Guided Distal Parasacral Block
Ahmad Mouhammad Taha
30. Ultrasound-Guided Anterior Sciatic Nerve Block
Yasuyuki Shibata, Toru Komatsu, and Laurent Delaunay
31. Ultrasound-Guided Subgluteal Sciatic Block
Steven L. Orebaugh and Paul E. Bigeleisen
32. Ultrasound-Guided Popliteal and Lateral Sciatic Block
Steven L. Orebaugh, Vihang D. Shah, and Paul E. Bigeleisen
33. Ultrasound-Guided Ankle Block
Stephen M. Breneman

SECTION IV TRUNCAL BLOCKS

34. Ultrasound-Guided Ilioinguinal and Iliohypogastric Nerve Blocks


Michael Gofeld and Urs Eichenberger
35. Ultrasound-Guided Transversus Abdominis Plane Block and Quadratus
Lumborum Block
Part 1. Transversus Abdominis Plane Block
Kim Russon and Rafael Blanco

Part 2. Quadratus Lumborum Block


Roger F. Shere-Wolfe, Vikram Bansal, Karen Boretsky, Mihaela Visoiu, and Paul E. Bigeleisen

36. Anatomy of the Paravertebral Space


Annelot Krediet, Nizar Maoyeri, Gerbrand J. Groen, and Paul E.
Bigeleisen
37. Ultrasound-Guided Medial Paravertebral Block, Lateral Paravertebral
Block, Pectoral Block, and Serratus Anterior Block
Part 1. Medial and Lateral Paravertebral Block
Paul E. Bigeleisen, Alon Ben-Ari, Andrea Fanelli, and Daniela Elena Francesca Ghisi Part 2.
Lateral Thoracic Paravertebral Block
Paul E. Bigeleisen, Alon Ben-Ari, Andrea Fanelli, and Daniela Elena Francesca Ghisi Part 3.
Pectoral and Serratus Anterior Block
Hanni Monroe, Ron Samet, and Paul Bigeleisen

38. Ultrasound-Guided Caudal, Lumbar, and Epidural Injection


Part 1. Caudal Epidural Steroid Injection: A Primer on the Combined Use
of Ultrasound and Fluoroscopy for Pain Management
Imanuel R. Lerman, Joseph C. Hung, and Dmitri Souzdalnitski

Part 2. Ultrasound for Labor Epidural Placement


Manuel C. Vallejo and Manasi Badve
SECTION V PEDIATRIC REGIONAL ANESTHESIA

39. Fundamentals of Ultrasound-Guided Pediatric Regional Anesthesia


Tarun Bhalla and Joseph D. Tobias
40. Ultrasound-Guided Brachial Plexus Block in Infants and Children
David P. Martin, Joseph D. Tobias, Stephen Lucas, Sunathenam Suresh,
and Paul E. Bigeleisen
41. Ultrasound-Guided Lower Extremity Blockade in Children
Ralph Beltran, Joseph D. Tobias, Stephen Lucas, and Paul E. Bigeleisen
42. Ultrasound-Guided Epidural Block (Caudal, Lumbar, Thoracic), Truncal
and Paravertebral Blocks in Children
Part 1. Introduction
Karen Boretsky, Paul E. Bigeleisen, Arvind Chandrakantan, Hadi S. Moten, and Mihaela Visoiu
Part 2. Single-Injection Caudal Approach to the Epidural Space
Karen Boretsky, Paul E. Bigeleisen, Arvind Chandrakantan, Hadi S. Moten, and Mihaela Visoiu
Part 3. Lumbar and Thoracic Epidural Catheters via the Caudal Approach
Karen Boretsky, Paul E. Bigeleisen, Arvind Chandrakantan, Hadi S. Moten, and Mihaela Visoiu
Part 4. Lumbar and Thoracic Continuous Epidural Analgesia via the Direct, Intervertebral
Approach
Karen Boretsky, Paul E. Bigeleisen, Arvind Chandrakantan, Hadi S. Moten, and Mihaela Visoiu
Part 5. Ultrasound-Guided Regional Anesthesia of the Thorax, Trunk, and Abdomen in Infants and
Children
Tarun Bhalla, Ralph Beltran, David P. Martin, Stephen Lucas, Paul E. Bigeleisen, and Joseph D.
Tobias

SECTION VI PAIN BLOCKS


SECTION VI PAIN BLOCKS
43. Ultrasound-Guided Maxillary and Mandibular Block
Part 1. Maxillary Nerve Block
Paul E. Bigeleisen and Milena Moreno
Part 2. Mandibular Nerve Block
Paul E. Bigeleisen and Milena Moreno

44. Ultrasound-Guided Stellate Ganglion Block


Yasuyuki Shibata, Toru Komatsu, Nizar Moayeri, and Gerbrand J. Groen
45. Ultrasound-Guided Cervical Sympathetic Block
Michael Gofeld
46. Celiac Ganglion Block: Endoscopic and Transabdominal Approaches
Paul E. Bigeleisen
47. Ultrasound-Guided Superior Hypogastric Plexus Block
Paul E. Bigeleisen
48. Ultrasound-Guided Genitofemoral Nerve Block
Michael Gofeld
49. Ultrasound-Guided Pudendal Nerve Block
Rachael Seib and Philip Peng
50. Neuraxial Anatomy Relevant to the Two-Dimensional Ultrasound
Examination
David Belavy
51. Ultrasound-Guided Third Occipital Nerve Block and Cervical Medial
Branch Block
Daniel L. Krashin and Michael Gofeld
52. Ultrasound-Guided Cervical Nerve Root Injections
Daniel L. Krashin and Michael Gofeld
53. Ultrasound-Guided Lumbar Transforaminal Epidural Steroid Injection
Kacey A. Montgomery and Michael Gofeld
54. Ultrasound-Guided Lumbar Facet Medial Branch Block and Intra-
articular Facet Joint Injection
Kacey A. Montgomery and Michael Gofeld
55. Ultrasound-Guided Botulinum for Spasticity and Applications in
Physiatry
Todd Beery
56. Ultrasound-Guided Sacroiliac Joint Injection
Kacey A. Montgomery and Michael Gofeld
57. Ultrasound-Guided Piriformis Injection
Kacey A. Montgomery and Michael Gofeld
58. Ultrasound-Guided Subacromial Bursa Injection
Michael N. Brown and Michael Gofeld
59. Ultrasound-Guided Bicipital Tendon Sheath Injection
Michael N. Brown and Michael Gofeld
60. Ultrasound-Guided Glenohumeral Joint Injection
Michael N. Brown and Michael Gofeld
61. Ultrasound-Guided Supraspinatus Tendon Injection
Michael N. Brown and Michael Gofeld
62. Ultrasound-Guided Acromioclavicular Joint Injection
Michael N. Brown and Michael Gofeld
63. Ultrasound-Guided Sternoclavicular Joint Injection
Michael N. Brown and Michael Gofeld
64. Ultrasound-Guided Intra-articular Elbow Injection
Michael N. Brown and Michael Gofeld
65. Ultrasound-Guided Wrist Injections
Michael N. Brown and Michael Gofeld
66. Ultrasound-Guided Ulnar Triquetral Injection
Michael N. Brown and Michael Gofeld
67. Ultrasound-Guided Carpal Tunnel Injection
Michael N. Brown and Michael Gofeld
68. Ultrasound-Guided Intra-articular Hip Injection
Michael N. Brown and Michael Gofeld
69. Ultrasound-Guided Patellar Tendon Injection
Michael N. Brown and Michael Gofeld
70. Ultrasound-Guided Intra-articular Ankle Injection
Michael N. Brown and Michael Gofeld
71. Ultrasound-Guided Tarsal Tunnel Injection
Michael N. Brown and Michael Gofeld
72. Ultrasound-Guided Peripheral Nerve Stimulation
Michael Gofeld
73. Ultrasound-Guided Intrathecal Pump Management
Hadi S. Moten and Brian Durkin
Index
Introduction

Ultrasound in
Trauma and
Critical Care
SAMUEL M. GALVAGNO JR AND JOSHUA

SAPPENFIELD

Introduction and Indications: Ultrasound has become an indispensable


tool for trauma and critical care practitioners. Ultrasound imaging
provides an additional clinical tool that can be used to answer specific
clinical questions and guide appropriate care. Due to advances in
ultrasound technology, including improved visualization and portability,
this technology has led to rapidly expanding applications. In this chapter,
the focused cardiac ultrasound study (FOCUS) and the focused
assessment with sonography (FAST) are introduced. In both of these
exams, simple binary questions are posed. For the FOCUS exam, the
operator systematically assesses left and right ventricular size and
function, the pericardial space for fluid and tamponade, and a color
Doppler assessment of valves to assess gross pathology. The inferior
vena cava may also be imaged to assess dynamic fluid responsiveness
parameters. In the FAST exam, four quadrants of the abdomen are
imaged to determine the presence or absence of free fluid.
Anatomy: For the FOCUS exam, four main views are obtained. The first
view, the parasternal long-axis view, involves a long-axis plane line
through the left ventricular apex, capturing the left atrium, mitral valve,
part of the left ventricle, and the left ventricular outflow tract. The right
ventricle is poorly visualized with this view. The descending thoracic
aorta is also visualized. The second view, the parasternal short-axis view,
is perpendicular to the long-axis view and is obtained by rotating the
ultrasound probe 90 degrees clockwise. The parasternal short-axis view
provides visualization of the aortic valve, the mitral valve, papillary
muscles, and the apex. The apical four-chamber view is perpendicular to
both the long-axis and short-axis parasternal views. The ultrasound
plane bisects the left ventricular apex and intersects both atria and both
ventricles. The subcostal view also provides a four chamber view of
both atria and the ventricles.
In the basic FAST exam, four ultrasound views are obtained. In the first
view, the right upper quadrant is imaged to evaluate the presence of free
fluid in the hepatorenal fossa (Morrison’s pouch). In the second view,
the probe is placed under the xiphoid process (subxiphoid view) to
detect pericardial fluid. If this view is difficult to obtain (e.g., due to
obesity, upper abdominal trauma), additional views from the FOCUS
exam can be used to evaluate the pericardium. The left upper quadrant is
imaged in the third view to detect perisplenic fluid. The final view
involves placement of the ultrasound probe in the suprapubic area to
evaluate the presence of free fluid in the rectouterine pouch (pouch of
Douglas) and rectovesical pouch (Fig. I.1).
Patient position: For the FOCUS exam, the patient is positioned supine.
The left lateral decubitus position may be used when possible to help the
heart fall laterally to the left, beyond the sternum. Additionally, the left
upper extremity may be abducted to open the intercostal spaces. In
critically ill patients who cannot tolerate position changes, the only
obtainable view may be the subcostal view. For the FAST exam, the
patient is positioned supine.

Transducer: 1 to 5 MHz phased array probe Approach and


Technique: (Fig. I.2)
FOCUS exam: 1. Parasternal long-axis view The transducer is
positioned in the 2nd to 4th intercostal space, parasternal. The index
marker is angled toward the patient’s right shoulder. Gentle
movements, including tilting, angulation, and rotation, are employed
to obtain optimal visualization of structures. Depth is typically 14 to
16 cm. The heart should be visualized in the most horizontal
orientation possible. Mechanically ventilated patients may have a
more vertical heart. This view does not provide an adequate view of
the right ventricle; only the right ventricular outflow tract is usually
visualized (Fig. I.3).
2. Parasternal short-axis view The transducer is maintained at the
same position as the parasternal long-axis view but rotated 90
degrees with the index marker pointing toward the patient’s left
shoulder. Gradual adjustments are made to visualize the left
ventricle. Depth is typically the same as for the parasternal long-axis
view (14 to 16 cm). Tilting the transducer cephalad helps obtain
views of the aortic valve and pulmonary arteries. Tilting the
transducer caudad brings the mitral valve and papillary muscles into
view. This view is the ideal view for assessing left ventricular
function (Figs. I.4 and I.5).
3. Apical four-chamber view The transducer is placed laterally at
medium depth (14 to 20 cm). The index marker on the transducer
should be oriented in the 2 o’clock or 3 o’clock position, facing the
patient’s left shoulder. In spontaneously breathing patients, the
transducer may be placed more medially to acquire the best view.
Lateral decubitus positioning may help improve the view. Color
Doppler can be applied to assess gross valvular function (i.e., severe
regurgitation) (Fig. I.6).
4. Subcostal four-chamber view The patient should be positioned
supine. The transducer is introduced under the xiphoid process, with
the index marker oriented to the patient’s left side. Because this is
another four-chamber view, this view is useful for assessing
pericardial fluid and right ventricular size and function; however,
off-axis views can lead to under-or overestimation (Fig. I.7).

5. Inferior vena cava Estimation of the size and collapsibility of the


inferior vena cava (IVC) may be useful in the evaluation of cardiac
preload and circulating fluid status. The use of ultrasound-guided
IVC measurements has been validated primarily in spontaneously
breathing patients. In mechanically ventilated patients, IVC
measurements are less reliable due to a high prevalence of IVC
dilation. The IVC diameter should normally be greater than 1 cm (10
mm) but less than 3 cm (30 mm). In conscious patients, the patient is
asked to take a brief inspiration or “sniff.” A normal response to the
sniff test is an inspiratory decrease in the diameter (>50%) of the
IVC. An abnormal response to the sniff is a dilated IVC (>20 mm)
without an inspiratory decrease in IVC diameter (<50%). An
abnormal sniff test may indicate elevated right atrial pressure and
fluid overload. Alternatively, an IVC less than 1 cm (10 mm) may
indicate hypovolemia (Fig. I.8).

FAST exam: The basic FAST exam is a four-view abbreviated


ultrasound technique directed at identifying the presence of free
intraperitoneal or pericardial fluid, usually due to hemorrhage.
Results are interpreted as positive, negative, or indeterminate.
Sensitivity varies between 42% and 96%, although specificity is
often higher. The FAST may reduce the need for diagnostic
peritoneal lavage (DPL), may reduce the requirement to obtain a
computed tomography (CT) scan, and may reduce the time from
initial diagnosis and time in the emergency department to the
operating room (Figs. I.9 to I.11).
Tips: 1. The three principles of the FOCUS exam are patient
positioning, image acquisition, and image interpretation.
2. Basic echocardiography is best learned with a “hands-on” approach.
Excellent introductory courses are available from the Society of
Critical Care Medicine (www.sccm.org) and the American College
of Chest Physicians (www.accp.org).
3. The FOCUS exam is designed to answer binary questions. For
example, the assessment includes left and right ventricular size and
function (normal versus abnormal), pericardial space (presence or
absence of fluid or tamponade), and color Doppler to assess the
aortic, mitral, and tricuspid valves for gross abnormalities (i.e.,
severe regurgitation, stenosis).
4. The FAST exam is a “rule-in” triage tool for patients with blunt
abdominal trauma; CT remains the gold standard for detecting
injuries in patients with blunt trauma.
5. Free fluid in the pelvis can be missed when the bladder is empty.
Hence, whenever possible, the bladder should be scanned before a
urinary catheter is placed.
6. Retroperitoneal hemorrhage is missed with the FAST, although some
newer protocols are designed to improve detection.
7. Both FOCUS and FAST exams may be limited in morbidly obese
patients or in patients with subcutaneous emphysema.
Section I
Fundamentals
1 Dermatomes, Sclerotomes, and Planes

2 Pharmacology of Perineural Analgesia

3 The Physics of Nerve Stimulation and Nerve


Impedance

4 Microanatomy of the Peripheral Nervous System


and Stimulation Thresholds Inside and Outside the
Epineurium

5 Ultrasound Equipment

6 Principles of Sonography

7 Three-and Four-Dimensional Ultrasound in


Neuraxial Anesthesia

8 Anatomical Anomalies in Ultrasound Simulation

9 Ultrasound Simulator Training

10 Advanced Ultrasound-Guided Needle Technology

11 From Paresthesia to Neurostimulation and


Ultrasound-Guided Regional Anesthesia

12 Understanding Needle-to-Nerve Proximity in


Peripheral Nerve Blocks

13 Ultrasound-Guided Intraneural Injection—The


Human Data
14 Use of Ultrasound for Placement of Peripheral
Nerve Block Catheters


1
Dermatomes,
Sclerotomes, and
Planes
PAUL E. BIGELEISEN AND JOOYEON HA

Anesthesiologists and algologists must understand the somatic and


visceral pathways of the body in order to perform efficient and accurate
local anesthetic injections. There is considerable variation in the
pathways of both visceral and somatic nerves. Nonetheless, it is useful
for the practitioner to have some idea of these pathways. The sensory
innervation of the skin is called dermatomes (Fig. 1.1). The sensory
innervation of the bones is called sclerotomes (Fig. 1.2). As a rule, if a
nerve crosses a joint, it usually provides innervation to the capsule and
bones of that joint. The sensory innervation of the muscles is called
myotomes. In general, the sensory innervation of the muscles is the same
as the motor innervation of the muscles. In most cases, ultrasound for
nerve block provides the practitioner a two-dimensional image of the
region of interest. In many cases, these cross sections can be described
with traditional terminology (Fig. 1.3).
2
Pharmacology of
Perineural Analgesia
BRIAN A. WILLIAMS, NIGAM SHETH, SYLVIA

WILSON, AND LAVINIA KOLARCZYK

Background and indications


Koller is credited with the use of cocaine for topical ophthalmic
anesthesia (1884),1 Bier is credited for the use of cocaine for spinal
anesthesia (1898),1 and Braun is credited for the use of epinephrine as an
adjuvant to procaine (1900).2 In the past 20 years, novel perineural
adjuvants (e.g., clonidine, buprenorphine, dexamethasone), have been
combined with local anesthetics. These appear to be safe in preclinical
laboratory models when combined with local anesthetics. The exception
is midazolam which significantly worsens in vitro neurotoxicity. The
local anesthetics discussed throughout this text book (bupivacaine,
ropivacaine, levo-bupivacaine, mepivacaine, lidocaine, and
chloroprocaine) can be more extensively reviewed in other current
textbooks.
Physiology and pharmacology
Local anesthetics (LAs) are weak bases (nonionized). Only the
nonionized isoform diffuses across the lipid nerve sheath of the
peripheral nerve. The onset of action depends on the amount of drug in
the nonionized form. Because of this, sodium bicarbonate (NaHCO3)
speeds block onset by increasing LA pH and local tissue pH closer to the
pKa of the LA. Conversely, acidosis or infection delays block onset by
decreasing the pH both of LA and of local tissue. The ionized isoform
acts intracellularly to block voltage-gated sodium (NaV) channels,
thereby inhibiting impulse conduction and slowing depolarization. The
transmembrane electrical threshold potential is not reached, and
therefore the action potential is not propagated (Fig. 2.1).

In perineural use, LAs bind NaV channels in the axolemma, preventing


NaV channels from opening and thus preventing action potentials.
Different NaV channel types have differing affinities for LAs, which
becomes important when considering local anesthetic systemic toxicity.
Specifically, cardiac NaV channel subtype 1.5 has remarkable affinity
for bupivacaine, and this accounts for the potential for cardiac arrest
with unwanted intravascular injection or absorption of bupivacaine.
Peripheral nerves (NaV 1.7, 1.8, 1.9), dorsal root ganglia (NaV 1.7), and
skeletal muscle (NaV 1.4) do not have significant, if any, presence of the
NaV 1.5 channel.
It should be clearly understood that the conduction block mechanism of
LA (at NaV channels) is not likely the only analgesic mechanism
associated with LA. Likewise, LA activity at NaV channels does not
contribute to the neurotoxic (i.e., nerve damage) risks of LA. Nerve
damage risks can occur at generally accepted clinical doses via
mechanisms unrelated to NaV channel activity.3 Axonal damage from
LA can occur due to a variety of mechanisms including: (1)
intrafascicular injection (which is theoretically preventable with
ultrasound-guided regional anesthesia by a well-trained practitioner), (2)
with LA in high concentration, or (3) in situations of prolonged exposure
(such as with a perineural catheter). More relevant is the potential for
disruption of numerous cellular functions; primarily, intracellular
calcium levels may play a central role.4 Specifically, there is disruption
of cytoplasmic calcium signaling, leading to release of calcium from
intracellular stores and leakage of calcium into the cell across the cell
membrane. Ultimately, excess cytoplasmic calcium leads to neuronal
death from the activation of intracellular kinases and the disruption of
electron transport in the mitochondria. Apoptosis (programmed cell
death) is closely linked to these intracellular calcium alterations.3 It is
critical to acknowledge that there is no evidence that ultrasound-guided
regional anesthesia protects patients from potentially toxic alterations of
intracellular calcium after peripheral nerve block.
The mechanisms of action for clonidine, buprenorphine, dexamethasone,
and midazolam have not been as well described. Research is needed to
determine the extent to which perineural adjuvants can reduce the
overall dose-volume-concentration of the LA. Reducing total LA dose
should be a primary objective in our subspecialty’s clinical and research
agenda due to the known neurotoxicity of LA.
Clonidine-mediated perineural analgesia does not occur via agonist
activity at the level of the alpha-2 adrenoreceptor; rather, it is mediated
via the hyperpolarization-activated cation current.5 The same appears to
be true for dexmedetomidine (based on in vivo preclinical models).6,7 It
should be noted that the alpha-2 adrenoreceptor does mediate
inflammatory responses of various white blood cells and appears to
decrease perineural inflammation at the injection site of local anesthetic
when alpha-2 agonists are co-administered with LA.8
Buprenorphine is presumably active after transaxonal influx at axon-
level opioid receptors and adjacent axoplasmic G-protein coupling
mechanisms, typical of opioid receptors, although this is not verified.
A corticosteroid (methylprednisolone) has been described to suppress
the transmission in thin unmyelinated C-fibers but not in myelinated A-
beta fibers.9 Whether this same mechanism, or other mechanisms, holds
true for perineural dexamethasone remains to be determined.
Midazolam is profoundly cytotoxic when combined with ropivacaine
and exposed to primary sensory neurons (rat) in culture.10 Perineural
analgesia from bupivacaine enhanced by midazolam has not been
mechanistically determined; possibilities include peripheral
benzodiazepine receptors or peripheral GABAA receptors. The authors
of this textbook chapter do not express or imply the endorsement of a
combination of midazolam and LA given as a peripheral nerve block to
patients. Restated, the authors do not recommend any patient study or
off-label treatment strategy that involves the use of combined LA and
midazolam for peripheral nerve blocks. In vitro combinations of
clonidine, buprenorphine, dexamethasone, and midazolam are not
neurotoxic in cell culture of primary sensory neurons, but combining
midazolam and ropivacaine is profoundly and synergistically
neurotoxic.10
Local Anesthetic and Adjuvant Technique
Drugs are freshly prepared, and all drugs should be preservative free.
Readers are referred to the recommended local anesthetic doses for each
block described in this textbook. Generally, the adjuvant technique
described in this chapter synthesizes clinical experience, clinical
research to date, and most relevant, recent in vitro research addressing
the relative neurotoxicity of LA versus adjuvants. Attention will be
directed toward the strategic reduction of LA concentrations when
perineural adjuvants are used. The multiple adjuvants described in the
following text are based on the in vitro findings that clonidine-
buprenorphine did not increase LA neurotoxicity more than did the use
of LA alone.10 Also considered is the in vitro dose-response evidence of
dexamethasone worsening LA neurotoxicity (Williams et al., manuscript
in review, 2010–2011, described further in the section “Summary of
Evidence”).
The use of clonidine, buprenorphine, or both as perineural analgesic
adjuvants is unequivocally considered off-label, and all associated
precautions in patient education and documentation are suggested.
Clonidine is added in doses as low as 0.25 µg/kg or as high as 1.5 µg/kg.
Patients assuming the sitting position for surgery (e.g., after interscalene
block; shoulder surgery in beach chair position) should be dosed at the
lowest dose range in this spectrum. Doses of 1 µg/kg clonidine or
greater can lead to more likely potential concerns of hypotension,
bradycardia, and sedation (especially in surgical positioning other than
supine).
Buprenorphine is added in doses from 1.5 to 3.0 µg/kg. Doses of 2.5 to
3.0 µg/kg are typically reserved for opioid-tolerant patients, and
buprenorphine used perineurally in these doses does not appear to create
opioid withdrawal symptoms or reverse systemic opioid analgesia.
Doses of 1.5 µg/kg are typically used in patients receiving interscalene
nerve blocks; otherwise, 2 µg/kg is a dose that can typically avoid
opioid-induced side effects (nausea, vomiting, pruritus, etc.).
Multimodal antiemetic prophylaxis,11 which has been recommended for
both general and regional anesthesia,12,13 is especially useful when
perineural buprenorphine use is planned (e.g., oral perphenazine, 8 mg,
and intravenous ondansetron-dexamethasone, 4 mg each).
In the absence of perineural dexamethasone dose studies addressing
doses less than 8 mg, these authors recommend that if the decision is
made to study dexamethasone or to use dexamethasone off-label in
routine clinical care, the dexamethasone dose per nerve does not exceed
1 to 2 mg until dose-response neurotoxicity is further elucidated.
Dexamethasone in 1 to 2 mg perineural doses extrapolated to cell culture
(and combined with ropivacaine 2.5 mg/mL, clonidine 1 µg/mL, and
buprenorphine 3 µg/mL) is likely neither cytoprotective nor cytotoxic
when compared with plain ropivacaine.
The use of combined LA and midazolam for peripheral nerve blocks is
absolutely not recommended by these authors.
Summary of evidence
For this textbook chapter, the authors consider clonidine and
buprenorphine to be textbook drugs,14,15 with dexamethasone being
more controversial at the time of this writing based on in vitro data
summarized below (Table 2.1).
Clonidine16 and buprenorphine17–19 are documented to be efficacious as
perineural analgesic adjuvants based on studies involving both placebo
controls and systemic controls.
There is no evidence of clonidine efficacy in perineural infusions with
ropivacaine in otherwise healthy patients. There are no reports of
buprenorphine and/or dexamethasone perineural infusions, with or
without local anesthetics. There are two cases reported describing the
perineural combination of clonidine-buprenorphine rendering motor-
sparing analgesia in the absence of LA.20 The emphasis hereafter will be
placed on LA combined with one or more perineural adjuvants.
Dexamethasone has been reported as added to LA in perineural doses of
8 mg.21,22 This textbook does not express or imply an endorsement for
this dose of dexamethasone for perineural use. In bench research,
ropivacaine 2.5 mg/mL and dexamethasone in this dose range
(extrapolated to cell culture) is not cytoprotective. In cell culture,
dexamethasone 66.6 µg/mL does not significantly alter ropivacaine
neurotoxicity, but dexamethasone 133.3 µg/mL does significantly
worsen ropivacaine-induced neurotoxicity (when compared with plain
ropivacaine).
In other words, there is an in vitro dose-response curve with respect to
higher concentrations of dexamethasone worsening ropivacaine
neurotoxicity. It is difficult to extrapolate drug concentrations in cell
culture to single-injection perineural dosing. However, in the absence of
perineural dexamethasone dose studies addressing doses less than 8 mg,
these authors recommend that if the decision is made to study
dexamethasone or use off-label in routine clinical care, the dose per
nerve not exceed 1 to 2 mg until dose-response neurotoxicity is further
elucidated. Dexamethasone in 1 to 2 mg perineural doses extrapolated to
cell culture (and combined with ropivacaine 2.5 mg/mL, clonidine 1
µg/mL, and buprenorphine 3 µg/mL) is neither cytoprotective nor
cytotoxic when compared with plain ropivacaine.
Liposomal bupivacaine
The U.S. Food and Drug Administration (FDA) recently approved a
liposomal bupivacaine for postoperative analgesia by wound infiltration.
The pharmacokinetics of liposomal bupivacaine cause a prolonged
duration of bupivacaine release. The suspension of bupivacaine in
multivesicular liposomes results in increased drug stability, slow release
of bupivacaine, increased time to reach peak plasma concentrations, and
an increased elimination half-life.23 Contact should be avoided with
antiseptics (e.g., chlorhexidine or iodine), as these may alter the lipid
layer, resulting in uncontrolled local anesthetic release. Compared with
bupivacaine with and without epinephrine, intravenous injection of
liposomal bupivacaine in animal studies has demonstrated improved
cardiac and central nervous system toxicity profiles.24 Patients receiving
liposomal bupivacaine have not experienced significant
electrocardiogram (PR, QRS, or QTc) changes, heart rate changes,
increased incidence of cardiovascular adverse events,25–26 or delayed
wound and bone healing.27 Although investigations have been published
in animal models investigating repeated injection,28 administration after
other local anesthetics29 and perineural injection,30 liposomal
bupivacaine is not clinically approved for perineural, epidural, or
intrathecal use and should not be administered with other local
anesthetics. Notably, granulomatous inflammation, a foreign body
reaction documented to occur with liposomal injection, was the only
pathologic change noted histologically in adipose tissue around the
brachial plexus after perineural use.30
The FDA approval was largely based on two multicenter, randomized,
placebo-controlled trials. These involved patients undergoing
hemorrhoidectomy31 or bunionectomy32 receiving surgical infiltration
with saline or liposomal bupivacaine at the end of surgery (Table 2.2).
The primary outcome for both was cumulative numeric pain scores,
measured as area under the curve (AUC). This was significantly lower in
the liposomal bupivacaine group compared with placebo. Opiate
consumption was also significantly lower in the liposomal bupivacaine
group, and patients were more likely to be opioid free compared with
placebo. In addition, patients in the liposomal group experienced fewer
adverse events than did those receiving placebo in the bunionectomy
study.

Conversely, two randomized, blinded, controlled trials comparing


liposomal bupivacaine to bupivacaine HCl have not observed the same
efficacy (Table 2.2). In patients undergoing mammoplasty33 or total knee
arthroplasty,34 liposomal bupivacaine or bupivacaine HCl with
epinephrine was infiltrated at the end of surgery. Mean cumulative pain
scores with activity, measured as AUC, was the primary outcome in both
studies and did not differ significantly (Table 2.2). In the breast
augmentation study, total opiate consumption was lower in the liposomal
bupivacaine group at 24 and 48 hours postoperatively. However, after
total knee arthroplasty, a difference was not found between mean
numeric pain scores, opiate consumption, or time to return to normal
activities between the treatment groups. The lack of a significant
difference was attributed to inadequate statistical power. Although a
meta-analysis by Dasta et al.35 did find liposomal bupivacaine superior
to bupivacaine HCl, it included nine studies from a range of surgeries
with a range of dosages and some with only a placebo comparison.
In summary, liposomal bupivacaine appears to be a safe choice for
wound infiltration to decrease opiate consumption and improve
postoperative analgesia compared with placebo. It continues to be
investigated for its effectiveness compared to bupivacaine HCl and
alternative modes of delivery.
References
1. Wulf HF. The centennial of spinal anesthesia. Anesthesiology.
1998;89:500–506.
2. Stahnke J, Stahnke J. The history of local anesthesia (author’s
transl) [in German]. MMW Munch Med Wochenschr.
1980;122:1236–1238.
3. Hogan QH. Pathophysiology of peripheral nerve injury during
regional anesthesia. Reg Anesth Pain Med. 2008;33:435–441.
4. Gold MS, Reichling DB, Hampl KF, et al. Lidocaine toxicity in
primary afferent neurons from the rat. J Pharmacol Exp Ther.
1998;285:413–421.
5. Kroin JS, Buvanendran A, Beck DR, et al. Clonidine prolongation
of lidocaine analgesia after sciatic nerve block in rats is mediated
via the hyperpolarization-activated cation current, not by alpha-
adrenoreceptors. Anesthesiology. 2004;101:488–494.
6. Brummett CM, Norat MA, Palmisano JM, et al. Perineural
administration of dexmedetomidine in combination with
bupivacaine enhances sensory and motor blockade in sciatic nerve
block without inducing neurotoxicity in rat. Anesthesiology.
2008;109:502–511.
7. Brummett CM, Padda AK, Amodeo FS, et al. Perineural
dexmedetomidine added to ropivacaine causes a dose-dependent
increase in the duration of thermal antinociception in sciatic nerve
block in rat. Anesthesiology. 2009;111:1111–1119.
8. Lavand’homme PM, Eisenach JC. Perioperative administration of
the alpha2-adrenoceptor agonist clonidine at the site of nerve injury
reduces the development of mechanical hypersensitivity and
modulates local cytokine expression. Pain. 2003;105:247–254.
9. Johansson A, Hao J, Sjolund B. Local corticosteroid application
blocks transmission in normal nociceptive C-fibres. Acta
Anaesthesiol Scand. 1990;34:335–338.
10. Williams BA, Hough KA, Tsui BY, et al. Neurotoxicity of adjuvants
used in perineural anesthesia and analgesia in comparison with
ropivacaine. Reg Anesth Pain Med 2011;36:225–230.
11. Glass PS, White PF. Practice guidelines for the management of
postoperative nausea and vomiting: past, present, and future. Anesth
Analg. 2007;105:1528–1529.
12. Williams BA, Kentor ML, Skledar SJ, et al. Routine multimodal
antiemesis including low-dose perphenazine in an ambulatory
surgery unit of a university hospital: a 10-year history. Supplement
to: Eliminating postoperative nausea and vomiting in outpatient
surgery with multimodal strategies including low doses of
nonsedating, off-patent antiemetics: is “zero tolerance” achievable?
ScientificWorldJournal. 2007;7:978–986.
13. Skledar SJ, Williams BA, Vallejo MC, et al. Eliminating
postoperative nausea and vomiting in outpatient surgery with
multimodal strategies including low doses of nonsedating, off-
patent antiemetics: is “zero tolerance” achievable?
ScientificWorldJournal. 2007;7:959–977.
14. Benzon HT, Rathmell JR, Wu CL, et al, eds. Raj’s Practical
Management of Pain. 4th ed. Philadelphia, PA: Elsevier; 2008.
15. Hadzic A, ed. Textbook of Regional Anesthesia and Acute Pain
Management. New York, NY: McGraw-Hill; 2007.
16. Eisenach JC, De Kock M, Klimscha W. Alpha(2)-adrenergic
agonists for regional anesthesia. A clinical review of clonidine
(1984–1995). Anesthesiology. 1996;85:655–674.
17. Candido KD, Franco CD, Khan MA, et al. Buprenorphine added to
the local anesthetic for brachial plexus block to provide
postoperative analgesia in outpatients. Reg Anesth Pain Med.
2001;26:352–356.
18. Candido KD, Winnie AP, Ghaleb AH, et al. Buprenorphine added to
the local anesthetic for axillary brachial plexus block prolongs
postoperative analgesia. Reg Anesth Pain Med. 2002;27:162–167.
19. Candido KD, Hennes J, Gonzalez S, et al. Buprenorphine enhances
and prolongs the postoperative analgesic effect of bupivacaine in
patients receiving infragluteal sciatic nerve block. Anesthesiology.
2010;113:1419–1426.
20. Whiting DJ, Williams BA, Orebaugh SL, et al. Case report:
postoperative analgesia and preserved motor function with
clonidine and buprenorphine via a sciatic perineural catheter. J Clin
Anesth. 2009;21:297–299.
21. Vieira PA, Pulai I, Tsao GC, et al. Dexamethasone with bupivacaine
increases duration of analgesia in ultrasound-guided interscalene
brachial plexus blockade. Eur J Anaesthesiol. 2010;27:285–288.
22. Parrington SJ, O’Donnell D, Chan VW, et al. Dexamethasone added
to mepivacaine prolongs the duration of analgesia after
supraclavicular brachial plexus blockade. Reg Anesth Pain Med.
2010;35:422–426.
23. Davidson EM, Barenholz Y, Cohen R, et al. High-dose bupivacaine
remotely loaded into multivesicular liposomes demonstrates slow
drug release without systemic toxic plasma concentrations after
subcutaneous administration in humans. Anesth Analg.
2010;110:1018–1023.
24. Boogaerts J, Declercq A, Lafont N, et al. Toxicity of bupivacaine
encapsulated into liposomes and injected intravenously:
comparison with plain solutions. Anesth Analg. 1993;76:553–555.
25. Bergese SD, Onel E, Morren M, et al. Bupivacaine extended-release
liposome injection exhibits a favorable cardiac safety profile. Reg
Anesth Pain Med. 2012;37:145–151.
26. Naseem A, Harada T, Wang D, et al. Bupivacaine extended release
liposome injection does not prolong QTc interval in a thorough
QT/QTc study in healthy volunteers. J Clin Pharmacol.
2012;52:1441–1447.
27. Baxter R, Bramlett K, Onel E, et al. Impact of local administration
of liposome bupivacaine for postsurgical analgesia on wound
healing: a review of data from ten prospective, controlled clinical
studies. Clin Ther. 2013;35:312–320.
28. Richard BM, Rickert DE, Newton PE, et al. Safety evaluation of
EXPAREL (DepoFoam Bupivacaine) administered by repeated
subcutaneous injection in rabbits and dogs: species comparison. J
Drug Deliv. 2011;2011:467429.
29. Richard BM, Rickert DE, Doolittle D, et al. Pharmacokinetic
compatibility study of lidocaine with EXPAREL in Yucatan
miniature pigs. ISRN Pharm. 2011;2011:582351.
30. Richard BM, Newton P, Ott LR, et al. The Safety of EXPAREL ®
(Bupivacaine liposome injectable suspension) administered by
peripheral nerve block in rabbits and dogs. J Drug Deliv.
2012;2012:962101.
31. Gorfine SR, Onel E, Patou G, et al. Bupivacaine extended-release
liposome injection for prolonged postsurgical analgesia in patients
undergoing hemorrhoidectomy: a multicenter, randomized, double-
blind, placebo-controlled trial. Dis Colon Rectum. 2011;54:1552–
1559.
32. Golf M, Daniels SE, Onel E. A phase 3, randomized, placebo-
controlled trial of DepoFoam® bupivacaine (extended-release
bupivacaine local analgesic) in bunionectomy. Adv Ther.
2011;28:776–788.
33. Smoot JD, Bergese SD, Onel E, et al. The efficacy and safety of
DepoFoam bupivacaine in patients undergoing bilateral, cosmetic,
submuscular augmentation mammaplasty: a randomized, double-
blind, active-control study. Aesthet Surg J. 2012;32:69–76.
34. Bramlett K, Onel E, Viscusi ER, et al. A randomized, double-blind,
dose-ranging study comparing wound infiltration of DepoFoam
bupivacaine, an extended-release liposomal bupivacaine, to
bupivacaine HCl for postsurgical analgesia in total knee
arthroplasty. Knee. 2012;19:530–536.
35. Dasta J, Ramamoorthy S, Patou G, et al. Bupivacaine liposome
injectable suspension compared with bupivacaine HCl for the
reduction of opioid burden in the postsurgical setting. Curr Med
Res Opin. 2012;28:1609–1615.
3
The Physics of Nerve
Stimulation and
Nerve Impedance
PHILIP C. CORY

For decades, anesthesiologists have been using equipment for nerve


stimulation that is nonoptimal. The following is a discussion of the
physics and electronics involved in nerve stimulation.
Requirements for nerve depolarization
Neuronal depolarization by externally applied electrical fields requires a
cathodal transmembrane voltage gradient in the range of 6 to 7 mV.1 In
his authoritative chapter, Hille2 states, “The conductance changes
apparently depend only on voltage and not on the direction or magnitude
of current flow [italics added].” An electric field strength of 100 mV/cm
is required to achieve such a transmembrane gradient for neurons longer
than five times their length constant (approximately 5 × 1 mm).3 This
cathodal gradient distorts the voltage-sensing fourth region (S4) of
sodium channel membrane–spanning segments to effect pore opening.
As positive charges on the S4 region move toward the extracellular side
of the membrane, an outwardly directed gating current occurs followed
by the inwardly directed current associated with sodium ions moving
from extracellular to intracellular.4 This gating current is not bulk ion
flow associated with the voltage gradient. It is the movement in space of
the charged moieties on the S4 region.
Coulomb’s Law
Coulomb’s law of electrostatics (eq. 1) defines forces between two
immobile charges in space, describing neither current flow nor current
effects at a distance.

This relationship states that the force (F, or electric intensity in V/m)
between two static charges (q1 and q2) is inversely related to the distance
(r) between the charges and the permittivity of free space (ε0).
Importantly, F is not current; it is voltage per unit distance. The
significant concept emerging from Coulomb’s law is that voltage
gradients are established without need for charge movement in space
(i.e., current flow). The converse is not true; for current to flow, charges
must move along a voltage gradient. This is a critical concept for
understanding nerve stimulation.
A voltage gradient between two points in a bulk conductor of uniform
composition causes current flow at right angles to the semicircular
equipotential lines of force as shown in Figure 3.1.5 No static charge
will accumulate in such a situation. In Figure 3.2, a mixture of positive
and negative charges arises because there are regions of differing
conductivity and paths of preferential current flow because tissue is best
described as nonhomogeneous and anisotropic.
Impedance
Tissue voltage/current electrical response is not just resistive as in
Ohm’s law (E = IR); it displays an impedance, comprising resistive and
capacitive components arranged in series and in parallel (RC circuit), as
shown in Figure 3.3. Electrical impedance is a measure of opposition to
time-variant electric current. The time function of impedance is
important to alternating current waveforms. Anesthesia nerve
stimulators, with duty cycles (the ratio of the pulse being “on” to that of
it being “off”) of 1:9,999 at 1 Hz with a 100 µs pulse width, do not
produce time-variant waveforms. When such square current pulses are
applied to a resistance, square voltage pulses result; however, when the
same current pulse is applied to an impedance, a voltage charging curve
is seen that reflects resistance and reactance. Resistance is fixed
opposition to current regardless of time or waveform considerations, but
reactance varies with the waveform (i.e., it reacts to the electrical field).
Impedance, resistance, and capacitive reactance are related in eq. 2,

where Z is impedance, f is frequency, and C is capacitance in farads.


Current, when introduced into a parallel RC circuit, initially distributes
charge onto capacitive surfaces. Very shortly after current flow
initiation, the repulsive action of accumulating charge on capacitive
surfaces shunts current through the resistance of the RC circuit. As a
progressively larger proportion of the charge flows through the
resistance, the voltage differential across the circuit increases in an
exponential fashion. When the current pulse is stopped, the voltage falls
in an exponential fashion; the complete curve is shown in Figure 3.4.

Direct current cannot flow through a capacitor; when the capacitive


surfaces are fully charged, the current flows entirely through the
resistive limb of the circuit, determining the final voltage value. The
charging portion of the curve is described by eq. 3,

where V0 is the final voltage, Vi is the initial voltage, t is time in seconds,


and τ is the circuit time constant equal to R × C. The discharge curve is
described by eq. 4,

The shape of this curve and the resultant impedance characteristics of


the total tissue electrical path are defined by the individual impedance
functions in the path, not just that of the nerve. These multiple
resistances and capacitances are related in eq. 5,

where C0 -Cn are constants and τ0 -τn are time constants (R × C) for the
component impedance sources.6 The initial portion of the charging curve
is determined by short time constants and has a steeper slope than the
final portion that is determined by longer time constants.
There are several sources of impedance in nerve stimulation, and
important among these are needle factors (e.g., the variable capacitance
of the insulated needle surrounded by conductive tissue). With
increasing depth of insertion, according to eq. 2, increasing needle
capacitance results in lower system impedance, requiring greater current
outputs to generate adequate voltage gradients for stimulation, alluded to
by Hadzic7 and modeled by Bashein et al.8 Impedance versus insertion
depth is demonstrated in saline (Fig. 3.5) and tissue (Fig. 3.6).9
Paresthesia current requirements at fixed depth with variable impedance
are shown in Figure 3.7 and demonstrate that developed voltage is the
critical factor for stimulation as current requirements varied inversely
with impedance.
Total system impedance is also affected by current density at the needle
tip/tissue interface. A critical current density for stainless steel, when
exceeded, is associated with nonlinear capacitive and resistive behaviors
shown in Figures 3.8 and 3.9.10 At a current output of 0.5 mA, the
current density at a 22G needle tip will be 77 mA/cm2, exceeding the
level of nonlinear behavior for capacitance and resistance of stainless
steel. Thus, current output increases from a stimulator applied to a 22G
needle are associated with smaller voltage changes than were the
relationship linear.
Cooper’s3 analysis also explains the importance of the neuronal
membrane time constant for depolarization by externally applied fields
shown in Figure 3.10. Spinomotor neuron time constants range from 1.0
to 12 msec.11,12 From these data, maximally effective nerve stimulators
should have pulse durations up to 5 to 6 msec. Aδ or C fiber–related
discomfort with longer pulses does not occur because pain is coded both
spatially (fiber class) and temporally (10 Hz or greater frequency of
impulses reaching the central nervous system). Pulses of 1 to 2 Hz
generated in these axons do not result in nociception.13,14
Nerve impedance
Neurons are low-impedance structures (i.e., long, uninterrupted
tubes).15–17 Neurons differ in this regard from vessels that contain large
numbers of cell membranes (the blood cellular constituents).1 Once a
stimulating needle penetrates the perineurium, current flow will partition
equally between intracellular and extracellular fluid over the length
constant (~1 mm) of the axons. Impedance to flow becomes the ohmic
resistance of these fluids.3,11 This is an electrical anisotropicity, or
facilitated conduction pathway.18 With intraneural sampling, short
axonal cell membrane time constants lead to higher initial developed
voltages for controlled current stimulation (Fig. 3.11).19,20 The greater
voltage/current ratio may be interpreted as an increased impedance.
However, such a determination performed on the ascending portion of
the charging curve does not reflect either resistance or impedance but
simply an instantaneous voltage/current sample for situations with
differing effective time constants. Time variant waveforms (i.e.,
alternating current) are required for accurate impedance determinations.

Finally, it is clear that controlled current output stimulators cannot


accurately create voltage gradients with increasing needle insertion
depth. This is due to variable needle capacitance and current density
effects at the needle tip. Only controlled voltage devices (e.g.,
radiofrequency lesioning stimulators and Grass stimulators) reliably
generate the necessary voltage gradients at depth since falling
impedance affects their current output and not voltage. Recalling Hille
and Figure 3.7, the magnitude of current flow is immaterial to
depolarization; the voltage gradient is the important factor. Further, local
voltage gradients created by controlled-current devices are determined
by the impedance of the total electrical path, not just that of the neuronal
cell membrane.
References
1. Cole KS. Membranes, Ions, and Impulses. Berkeley and Los
Angeles, CA: University of California Press; 1972:1–569.
2. Hille B. Ionic basis of resting and action potentials. In: Brookhart
JM, Mountcastle VB, Kandel ER, eds. The Nervous System.
Baltimore, MD: Waverly Press, Inc.; 1977:99–136.
3. Cooper MS. Membrane potential perturbations induced in tissue
cells by pulsed electric fields. Bioelectromagnetics. 1995;16:255–
262.
4. Armstrong CM. Voltage-dependent ion channels and their gating.
Physiol Rev. 1992;72:S5–S13.
5. van Nostrand RG, Cook KL. Interpretation of Resistivity Data.
Washington, DC: U.S. Government Printing Office; 1966.
Geological Survey Professional Paper 499.
6. Rall W. Membrane potential transients and membrane time constant
of motoneurons. Exp Neurol. 1960;2:503–532.
7. Hadzic A. Lower Extremity Peripheral Nerve Blocks. ASA 57th
Annual Refresher Course Lectures and Basic Science Reviews
2006;236:1–6.
8. Bashein G, Haschke RH, Ready LB. Electrical nerve location:
numerical and electrophoretic comparison of insulated vs
uninsulated needles. Anesth Analg. 1984;63:919–924.
9. Ragheb T, Riegle S, Geddes LA, et al. The impedance of a spherical
monopolar electrode. Ann Biomed Eng. 1992;20:617–627.
10. Geddes LA, da Costa CP, Wise G. The impedance of stainless steel
electrodes. Med Biol Eng. 1971;9:511–521.
11. Rall W. Core conductor theory and cable properties of neurons. In:
Brookhart JM, Mountcastle VB, Kandel ER, eds. The Nervous
System. Baltimore, MD: Waverly Press, Inc.; 1977:39–97.
12. Ranck JB. Which elements are excited in electrical stimulation of
mammalian central nervous system: a review. Brain Res.
1975;98:417–440.
13. Mayer DJ, Price DD, Becker DP. Neurophysiological
characterization of the anterolateral spinal cord neurons
contributing to pain perception in man. Pain. 1975;1:51–58.
14. Willis JD. Evidence for nociceptive transmission systems. In Willis
WD, ed.: The Pain System: The Neural Basis of Nociceptive
Transmission in the Mammalian Nervous System. Basel,
Switzerland: S. Karger; 1985:7–21.
15. Cooper MS, Miller JP, Fraser SE. Electrophoretic repatterning of
charged cytoplasmic molecules within tissues coupled by gap
junctions by externally applied electric fields. Dev Biol.
1989;132:179–188.
16. Prokhorov E, Llamas F, Morales-Sánchez E, et al. In vivo
impedance measurements on nerves and surrounding skeletal
muscles in rats and human body. Med Biol Eng Comput.
2002;40:323–326.
17. Cory PC, inventor; Nervonix, Inc., assignee. Non-invasive,
peripheral nerve mapping device and method of use. U.S. Patent
5,560,372. October 1, 1996.
18. Rudy Y, Plonsey R. The eccentric spheres model as the basis for a
study of the role of geometry and inhomogeneities in
electrocardiography. IEEE Trans Biomed Eng. 1979;26:392–399.
19. Cory PC. Increased impedance on nerve stimulator display may
actually reflect a decrease in total system impedance.
Anesthesiology. 2009;110:1192–1194.
20. Gabriel S, Lau RW, Gabriel C. The dielectric properties of
biological tissues: II. Measurements in the frequency range 10Hz to
20GHz. Phys Med Biol. 1996;41:2251–2269.
4
Microanatomy of the
Peripheral Nervous
System and
Stimulation
Thresholds Inside
and Outside the
Epineurium
Part 1: Microanatomy

ANDRE P. BOEZAART, PAUL E. BIGELEISEN,



AND NIZAR MOAYERI

The use of ultrasound has provided anesthesiologists a deeper


understanding of the location of peripheral nerves as well as a powerful
tool to anesthetize these nerves. The wavelength of medical ultrasound
used for nerve blocks ranges from 0.1 to 1 mm. In most cases, the nerves
that we wish to image range from 3 to 15 mm in diameter. Under
optimal conditions, we may, therefore, be able to “look” inside large
nerves or a nerve plexus and “see” individual fascicles. Unfortunately,
the resolution provided by ultrasound does not allow us to view the
perineurium, a protective sheath that surrounds the fascicles as the nerve
root emerges from the lateral recess of the spinal canal and travels
toward the periphery as a peripheral nerve. However, in small patients
and children, we may be able to image the dura mater where it surrounds
the cauda equina and as the nerve root emerges from the lateral recess in
the cervical spine. The latter may be of use when performing
transforaminal cervical blocks. Some anesthesiologists may not be
familiar with the ultrastructure of nerve roots and nerves and, for this
reason, may not appreciate the risks involved with paravertebral nerve
blocks—even with ultrasound guidance.
Catastrophic outcomes following paravertebral blocks at the cervical,1–4
thoracic,5 and lumbar levels6–8 have been reported, some of which were
reversible cases of extensive epidural/subdural block or total spinal
anesthesia. Unfortunately, other cases resulted in paraplegia,
quadriplegia, and death. The explanation for these catastrophes focused
on intracord injection, which complicated the blocks performed on
patients under general anesthesia. Interscalene block has also been
blamed for disastrous outcomes.9–14 The report by Benumof9 of four
cases of spinal injury after interscalene block attracted much attention
and generated a fierce debate on the safety of performing blocks on
patients under general anesthesia. The conclusions reached by Benumof
—that blocks should not be performed under general anesthesia—were,
regrettably, largely incorrect, particularly in the case of pediatric
patients. These tragic outcomes, like many others, were most likely
caused by intraroot injection and had nothing to do with the fact that the
patients were under general anesthesia when the blocks were placed.
Benumof’s conclusions were largely based on a misunderstanding of the
microanatomy of the connective tissue framework of the nervous
system. All of these cases have two things in common: root level nerve
blocks and thin, relatively sharp needles.15–17
Our understanding of the microanatomy of the peripheral nervous
system is already over 130 years old. Key and Retzius18 in 1876
(Richardson’s stain) and Horster and Whitman19 in 1931 (trypan blue)
studied the spread of intraneurally injected solutions. In 1948, French et
al.20 repeated this work with radiopaque contrast medium in dogs. In
1952, Moore et al.21 used methylene blue–stained exocaine, and in 1978,
Selander and Sjostrand22 used radioactive local anesthetic with
fluorescent dye to study the microanatomy of the sciatic nerve in rabbits.
In the interval, electron microscopy23 was used to confirm what was
already known.24
According to conventional teaching, (including many modern anatomy
and anesthesia textbooks), the cerebrospinal fluid (CSF) originates in the
choroid plexus, is discharged into the cerebral ventricles, and exits
through the foramina of Lushka and Magendie. The CSF then gathers in
the cisterns at the base of the brain, where it flows to the villi or
pacchionian bodies, and then into the peripheral venous circulation.
In 1948, Hassin25 proposed a new depiction of how the CSF courses
through the circulation.
In summary:

1. The CSF is the extracellular fluid of the brain and spinal cord.
2. The circulation of the CSF also involves the Virchow-Robin spaces
that surround the arterioles in the brain; these spaces form the blood–
brain barrier.
3. Absorption of the CSF is not through the villi or pacchionian bodies
only but also through the perineurial spaces of the cranial nerves and
spinal roots.
4. The CSF acts as the “lymph fluid” of the central nervous system and
carries away waste.
5. There is no central force, per se, that drives the CSF into the
circulation. The cardiac cycle causes expansion and contraction of the
brain and spinal cord, which are encased in a rigid compartment.
During systole, the entire brain and spinal cord expand, and pressure
in the CSF increases. Following a pressure gradient, the CSF flows
from the central space out into the perineural spaces of the cranial and
spinal nerve roots.
Peripheral nerves and plexuses
The tissue fluid deep to the epineurium, but outside of the perineurium,
in a peripheral nerve is lymph and drains to the regional lymph nodes.25
The axons of peripheral nerves are extensions of nerve cells in the
central nervous system. These axons, which are surrounded by
perineurium, form fascicles and are bathed by CSF. Under normal
conditions, the longitudinal flow of the CSF within the fascicle is
minimal.22 Lateral extension (centrifugal) of the perineurium is minimal
even under high pressure. As the nerve approaches dural penetration,
resistance to lateral extension increases and a peripherally injected
medium comes to lie in the clefts of the perineurium. Final emergence
into the subarachnoid space appears to occur, first, by way of the
subdural space and, subsequently, by breaking through the arachnoid
barrier into the subarachnoid space. Injection into peripheral nerve
fascicles, which is difficult to achieve under clinical conditions, provides
direct access to the CSF and interstitium of the spinal cord. Conversely,
penetration and injection into a spinal root is relatively easy under
clinical conditions, and this injectate similarly has direct access to the
CSF and spinal cord interstitium (Fig. 4.1). The clinical consequence of
an injection into a spinal root will depend on the volume, rate, and
pressure of the injectate.
Peripheral nerves are composed of numerous fascicles that contain
axons. Each fascicle is bounded by a dense perineurium, and a fine
epineurial membrane holds the fascicles together (Fig. 4.2). The
epineurium consists of a condensation of areolar connective tissue that
surrounds the perineural ensheathment of the fascicles on uni-and
multifascicular nerves. The attachment of the epineurium to surrounding
connective tissue is loose, so that the nerve is relatively mobile except
where tethered by entering blood vessels or nerve branches. Greater
amounts of connective tissue are normally present where nerves cross
over joints, and nerves usually carry sensory signals from branches from
those joints. In general, the more fascicles within a nerve, the thicker the
epineurium would be. Variable quantities of fat are also present in the
epineurium, particularly in larger nerves. This fat cushions the fascicles
against injury by compression; thus, large multifascicular nerves are less
susceptible to injury by compression than are smaller or unifascicular
nerves. The vasa nervorum enter the epineurium, where they
communicate with a longitudinal anastomotic network of arterioles and
venules. The epineurium also contains lymphatic vessels, which are not
present within the fascicles. These lymphatic channels accompany the
arteries of the peripheral nerves and pass into the regional lymph
nodes.25

The essential structure of the perineurium is a lamellated arrangement of


flattened cells separated by layers of collagenous connective tissue.18 It
provides ensheathment for both the somatic and peripheral autonomic
nerves and their ganglia. The cellular lamellae are composed of
concentric sleeves of flattened polygonal cells. These cells are equipped
to function as a metabolically active diffusion barrier, although they do
not have the morphologic features of a true epithelium. The
endoneurium refers to intrafascicular connective tissue, excluding the
perineurial portions that may subdivide fascicles. About 40% to 50% of
the intrafascicular space is occupied by nonneural elements, and about
20% to 30% of this is the endoneural fluid (CSF) and connective matrix,
which we call the endoneurium.24
The injection of local anesthetic deep to the epineurium, but outside the
perineurium, in humans and pigs causes a centrifugal expansion of the
epineurium; this is referred to as a subepineurial injection. The injection
of several milliliters of local anesthetic deep to the epineurium will
cause the diameter of the nerve to double. When 10 to 20 mL of local
anesthetic is injected into the subepineurial space, the epineurium
ruptures (neurolysis), causing the local anesthetic to surround the nerve.
The injection of several milliliters of local anesthetic into the
subepineurial space over a few seconds may cause the pressure within
the subepineurial space to rise to 15 pounds per square inch. Once the
injection is stopped, the pressure inside the epineurium returns to
baseline pressure within seconds.
When Selander and Sjostrand 22 injected blue tracer into a fascicle
(subepineurial injection) of rabbit sciatic nerves, the tracer spread
rapidly, both proximally and distally, within the fascicle. The injection
within the fascicle (0.05 to 0.1 mL) created pressures between 435 and
675 mm Hg. These pressures remained above capillary perfusion
pressure (50 mm Hg) for at least 10 minutes. In all cases, the injection
reached the sacral plexus. Distally, the tracer colored the tibial nerve,
sometimes reaching the foreleg. If the injection was made into a small
fascicle, the injectate did not extend beyond the sacral plexus. If the
tracer was injected into a large fascicle, the injectate passed the sacral
plexus and reached the spinal medulla; during slow injection, it spread in
the medulla superficially, under the pia mater. In some animals, the
spread was into the CSF, and the dura and arachnoid were also colored
with the tracer. In one animal, the blue stain extended to the cerebellum,
and the animal died shortly after injection22. In cross sections of the
spinal medulla, the tracer was noted in the dorsal root–medulla junction
area, extending into the substantia gelatinosa and into the anterior
median fissure. In the canine studies of French et al.,20 the radiopaque
tracer reached the lumbar plexus via the injected fascicle and then
tracked distally via an entirely different nerve (Fig. 4.3). The contrast
injected intrathecally under high pressure also spreads peripherally into
the fascicles of peripheral nerves.
Plexus trunks and spinal roots
The distal roots and trunks of the plexuses should be seen as transitional
areas where the fascicles are no longer ensheathed by the perineurium.
At the trunk level, the perineurium has split into septae (Fig. 4.4).
Functionally, the trunks should be regarded as a zone between peripheral
nerves with clearly defined fasciculi and rigid perineuria, and the root
area is where perineurial septae have joined to form the dura mater. The
perineuria of peripheral nerves, therefore, are continuations of the dura
mater. The axons inside roots are no longer protected by the
perineurium, and the tissue fluid inside the root is the CSF (Fig. 4.4).
When Bigeleisen26 injected 2 mL of India ink over 2 seconds into the C5
nerve root of a fresh cadaver, the pressure exceeded 30 pounds per
square inch. The tracer stained the subarachnoid space at the C5 level
and traveled distally into a fascicle of the posterior cord (Fig. 4.1).

The mesothelial cells of the arachnoid membrane become hyperplastic at


the point where the nerve leaves the spinal cord and form a cuff around
the roots just after they penetrate the dura mater. Beyond this cuff, no
tissue that is recognized as arachnoid can be seen. The connective tissue
framework of the peripheral nervous system, therefore, appears to arise
entirely from a continuation of the perineurium, starting at the dura
mater. As the nerve progresses peripherally, it is further subdivided by
perineural septations until each fascicle eventually has its own perineural
sheath.
In a series of seminal papers, Moayeri27,28 quantified the extent of neural
and nonneural tissue in the nerve root and peripheral nerves of the
brachial plexus and sciatic nerve (Figs. 4.5 and 4.6). In particular, he
found that the percentage of neural tissue within the nerve was greatest
at the nerve root and diminished in the more distal parts of the nerve
(Figs. 4.7 and 4.8). This most likely gives rise to the well-known dictum
that the more distal a nerve block is placed, the safer it is to block it. At
the level where the supraclavicular block is performed, the deep cervical
fascia is adherent to the epineurium surrounding the plexus on the
superior surface of the plexus (Fig. 4.4B). Some practitioners refer to
this tissue as the plexus sheath. Puncture and injection of this sheath
invariably results in an intraneural injection. Injection outside this fascial
layer rarely results in a clinical block. In contrast, the inferior surface of
the plexus is covered only by epineurium. Thus, injection deep to the
plexus often results in a satisfactory block as the local anesthetic diffuses
through the epineurium.
Practical matters and clinical consequences
With the recent increase in ultrasound utilization, it has become apparent
to practitioners that nerves can range from hypoechoic to hyperechoic in
appearance.29 In most cases, the closer the nerve lies to the spine, the
more likely it is to be hypoechoic. With the insight of microanatomy,
this is easy to understand. At the root level, there are mostly axons
interspersed with CSF, which, similar to the fluid within the axons, is
hypoechoic. In the periphery, the nerve has stroma and fat outside the
perineurium, which is hyperechoic. The perineurium itself, which
divides the nerve into fascicles, is also hyperechoic. This architecture
gives rise to the honeycomb appearance of peripheral nerves.29
Recent ultrasound work on axillary nerve blocks has also confirmed our
clinical impressions. Intraneural injections have no bad consequences as
long as the injection is deep to the epineurium but outside the
perineurium.30,31 Intrafascicular injections are difficult when blunt
needles are used. Nonetheless, the work of French et al.20 showed that
contrast medium injected into fascicles lingered for up to 5 weeks. This
finding makes the occasional numb finger that we encounter in our
patients from time to time, several weeks after an interscalene or
supraclavicular block, more comprehensible.
Injections at the root level and trunk level in some individuals should be
regarded as epidural injections because the injection is made directly
outside the dura or peridural space. All of the time-tested rules for
epidural injections should, thus, apply for root-level or paravertebral
injections. This should include the avoidance of sharp, thin needles.
Instead, the use of large, blunt needles for continuous and single
injections should be the norm.32 Fractionation of the dose, with frequent
aspiration for vascular and intrathecal injection, should also be
performed; perhaps, it also would be wise to follow the anticoagulation
guidelines for epidural placement.
All of the tragic cases referred to previously could reasonably be
attributed to intraroot injections with relatively thin, sharp needles that
can easily penetrate the dura. These blocks were performed with needles
that should never be used for an epidural block, and yet they were used
for a form of epidural block: the paravertebral or paraspinal block. The
differences in the ultimate clinical presentations were merely a function
of the volume, rate, and pressure of the injection. All or most of these
tragic outcomes may have been avoided by using large-bore Tuohy
needles, test dosing, and ultrasound. An understanding of the functional
microanatomy would have gone a long way toward understanding what
was happening to these unfortunate patients.
Sadly, many cases of total spinal anesthesia and permanent nerve
damage continue to occur in the anesthesia community because of a lack
of understanding of the ultrastructure of nerves. Because these injuries
have previously been reported in the literature, these additional cases no
longer reach the standard anesthesia community as case reports. In
addition, the risk of monetary and legal punishment to practitioners
discourages practitioners from discussing these complications publicly.
It is the authors’ hope that the information in this chapter will diminish
this problem.
References
1. Voermans NC, Crul BJ, de Bondt B, et al. Permanent loss of
cervical spinal cord function associated with the posterior approach.
Anaesth Analg. 2006;102:326–335.
2. Aramideh M, van den Oever HLA, Walstra GJ, et al. Spinal
anesthesia as a complication of brachial plexus block using the
posterior approach. Anaesth Analg. 2002;94:1338–1339.
3. Grefkens JM, Burger K. Total spinal anaesthesia after an attempted
brachial plexus block using the posterior approach. Anaesthesia.
2006;61(11):1105–1108.
4. Gentili M, Aveline C, Bonnet F. Total spinal anesthesia after
posterior plexus block. Ann Fr Anesth Reanim. 1998;17(7):740–
742.
5. Lekhak B, Bartley C, Conacher ID, et al. Total spinal anaesthesia in
association with insertion of a paravertebral catheter. Br J Anaesth.
2001;86(2):280–282.
6. Houten J. Paraplegia after lumbosacral nerve root block, report of
three cases. The Spine J. 2002;2(1):70–75.
7. Pousman RM, Mansoor Z, Sciard D. Total spinal anesthetic after
continuous posterior lumbar plexus block. Anesthesiology.
2003;98(5):1281–1282.
8. Litz R. Misplacement of a psoas compartment catheter in the
subarachnoid space. Reg Anesth Pain Med. 2004;29(1):60–64.
9. Benumof JL. Permanent loss of cervical spinal cord function
associated with interscalene block performed under general
anesthesia. Anesthesiology. 2000;93(6):1541–1544.
10. Ross S, Scarborough CD. Total spinal anesthesia following
brachial-plexus block. Anesthesiology. 1973;39(4):458.
11. Dutton RP, Echhardt III WF, Sunder N. Total spinal anesthesia after
interscalene blockade of the brachial plexus. Anesthesiology.
1994;80:939–941.
12. Tetzlaff JE, Yoon HJ, Dilger J, et al. Subdural anesthesia as a
complication of an interscalene brachial plexus block. Reg Anesth.
1994;19:5357–5359.
13. Dutton RP, Eckhardt III WF, Sunder N. Total spinal anesthesia after
interscalene blockade of the brachial plexus. Anesthesiology.
1994;80(4):939–941.
14. Lee WJ. Catastrophic complications of the interscalene nerve block.
Anesthesiology. 2001;95(5):1301.
15. Capdevila X, Macaire P, Dadure C, et al. Continuous psoas
compartment block for postoperative analgesia after total hip
arthroplasty: new landmarks, technical guidelines, and clinical
evaluation. Anesth Analg. 2002;94:1606–1613.
16. Boezaart AP, Koorn R, Rosenquist RW. Paravertebral approach to
the brachial plexus: an anatomic improvement in technique. Reg
Anesth Pain Med. 2003;28:241–244.
17. Boezaart AP, de Beer JF, Nell ML. Early experience with
continuous cervical paravertebral block using a stimulating
catheter. Reg Anesth Pain Med. 2003;28:406–413.
18. Key A, Retzius G. Stdien in der Anatomie des Nervensystems und
des Bindegewebes (2.Hafte). Stockholm: Samson & Wallin; 1876.
19. Horster H, Whitman L. Die Methode der intraneuralen Injektion. Z
Hygiene. 1931;113:113.
20. French JD, Strain WH, Jones GE. Mode of extension of contrast
substances injected into peripheral nerves. J Neuropathol Exp
Neurology. 1948;7:47–58.
21. Moore DC, Hain RF, Ward A, et al. Importance of the perineural
spaces in nerve blocking. J Amer Med Assoc. 1954;156:1050–1055.
22. Selander D, Sjostrand J. Longitudinal spread of intraneurally
injected local anesthetics. Acta Anaesth Scand. 1978;22:622–634.
23. Rohlich P, Knoop A. Elektronenmikroskopische Untersuchenen an
den Hullen des N. Ischiadicus der Ratte. Z Zellforsch Mikrosk Anat.
1961;53:288–305.
24. Thomas PK, Berthold CH, Ochoa J. Microscopic anatomy of the
peripheral nervous system. In: Dyck PJ, Thomas PK, eds.
Peripheral Neuropathy. 3rd ed. Philadelphia, PA; 1993:28–91.
25. Hassin GB. Cerebrospinal fluid: its origin, nature, and function. J
Exp Neuropath Axp Neurol. 1948;7(1):172–181.
26. Bigeleisen PE. Intraroot injection of India ink and contrast dye in
fresh cadaver. Where does the dye go and what does it mean?
Submitted to Anesthesiology
27. Moayeri N, Bigeleisen PE, Groen GJ. Quantitative architecture of
the brachial plexus and surrounding compartments, and their
possible significance for plexus blocks. Anesthesiology.
2008;108(2):299–304.
28. Moayeri N, Groen GJ. Differences in quantitative architecture of
sciatic nerve may explain differences in potential vulnerability to
nerve injury, onset time, and minimum effective anesthetic volume.
Anesthesiology. 2009;111(5):1128–1134.
29. Chan VWS. The use of ultrasound for peripheral nerve blocks. In:
Boezaart AP, ed. Anesthesia and Orthopaedic Surgery. New York,
NY; 2006:283–290.
30. Bigeleisen PE. Nerve puncture and apparent intraneural injections
during ultrasound-guided axillary block does not invariably result
in neurologic injury. Anesthesiology. 2006;105:779–783.
31. Bigeleisen PE, Moayeri N, Groen JG. Extraneural versus intraneural
stimulation thresholds during ultrasound-guided supraclavicular
block. Anesthesiology. 2009;110(6):1235–1243.
32. Boezaart AP, Franco CD. Thin sharp needles around the dura
(letter). Reg Anesth Pain Med. 2006;31:388–389.
Part 2: Nerve Stimulation
Thresholds
RON SAMET AND PAUL E. BIGELEISEN

Historically, peripheral nerve blocks were performed via the landmark


technique. Based on the assumed location of nerves relative to palpable
bones, muscles, and vessels, a needle was inserted “blindly” through the
skin. When a paresthesia (i.e., an abnormal sensation such as tingling or
burning) was elicited in the distribution of the nerve, the needle tip was
presumed to have made contact with the nerve and local anesthesia was
injected. In contrast, if a dysesthesia (i.e., a painful sensation often
described as an electric shock) was elicited in the distribution of the
nerve, needle penetration into the nerve was suspected. Practitioners
were advised to withdraw the needle until the dysesthesia resolved to
avoid intraneural injection and theoretical nerve injury.
In the 1960s, the first nerve stimulation–assisted nerve blocks were
performed by Greenblatt and Denson and were later popularized in the
1980s by Raj. The technique involved inserting an insulated electrical-
stimulating needle blindly through the skin via surface landmarks until a
contraction was observed in the muscle innervated by the nerve. Motor
contraction due to stimulation of the nerve at stimulation currents of 0.3
to 0.5 mA was proposed to suggest that the needle tip was in close
proximity to the nerve and that local anesthetic administration at these
settings would induce successful regional anesthesia. Sustained motor
contraction, however, at less than 0.3 mA was assumed to indicate
intraneural needle placement, and clinicians were warned to withdraw
the needle to avoid potential nerve damage when this occurred.
With the advent of ultrasound and the ability to sonographically
visualize the needle-to-nerve relationship, prior theories came into
question. Investigators began to study the symptomatology and motor
response of various placements of the needle relative to the nerve. In
2006, Bigeleisen1 demonstrated under direct ultrasound imaging that
needle-to-nerve contact does not generate paresthesias. Rather, only
when the needle breaches the epineurium are paresthesias produced.
Moreover, Bigeleisen discovered that this phenomenon does not occur in
all patients. Clear sonographic evidence of epineurium penetration does
not reliably cause paresthesias in many patients and, surprisingly,
intraneural placement of a needle rarely caused a dysesthesia in sedated
patients.1 Investigators began studies of combined ultrasound and
stimulation-guided nerve blocks of the brachial plexus and sciatic nerve
and found that stimulation thresholds of 0.5 mA frequently resulted in
penetration of the epineurium. Stimulation thresholds of 0.2 mA or less
almost always resulted in intraneural placement of the needle.2,3
For practitioners who wish to avoid intraneural injection, reevaluation of
the appropriate stimulation threshold for successful, yet extraneural,
needle placement is indicated. Ultrasound guidance for most superficial
blocks provides adequate resolution to visualize the needle adjacent to
the nerve but cannot determine whether the needle is within a fascicle
once the epineurium has been breached. With an appropriate local
anesthetic volume administered, and enough time for the block to set,
conditions to provide surgical anesthesia can usually be met when the
needle is seen outside the nerve. When deep blocks are performed,
however, and ultrasound imaging is less accurate, stimulation remains
the technique of choice. If a surgical block is desired, sustained muscle
contractions at stimulation thresholds of 0.4 mA are advised, with the
understanding that many of these blocks will be subepineurial. Whereas,
if the objective is to provide analgesia, stimulation thresholds of 1.0 mA
or greater are usually sufficient to provide analgesia with the likely
security of an extraneural needle placement. Regardless of the intent or
technique used, intrafascicular injections, heralded by injection pressures
greater than 1 atm or significant dysesthesias on injection, should be
terminated immediately and the needle repositioned.
References
1. Bigeleisen PE. Nerve puncture and apparent intraneural injection
during ultrasound-guided axillary block does not invariably result
in neurologic injury. Anesthesiology. 2006;105(4):779–783.
2. Sala-Blanch X, Lopez AM, Pomes J, et al. No clinical or
electrophysiologic evidence of nerve injury after intraneural
injection during sciatic popliteal block. Anesthesiology.
2011;115(3):589–595.
3. Bigeleisen PE, Moayeri N, Groen JG. Extraneural versus intraneural
stimulation thresholds during ultrasound-guided supraclavicular
block. Anesthesiology. 2009;110(6):1235–1243.
5
Ultrasound
Equipment
STEVEN L. OREBAUGH AND PAUL E.

BIGELEISEN

A variety of equipment is necessary for conducting ultrasound-guided


peripheral nerve block. Some practitioners prefer to use nerve
stimulation along with sonography to verify or confirm the nerve target
before injection. Although both methods have proponents, it is not clear
that either technique is superior, and efficacy appears similar.1
Equipment
Whether blocks are conducted in the operating room, the preoperative
holding area, or a formal block area, patients should have an intravenous
catheter initiated, monitors placed, and supplemental oxygen delivered
prior to the block (Fig. 5.1). The induction of regional anesthesia has
risks that are similar to those of general anesthesia. For this reason, a
block cart stocked with equipment for regional block should also contain
appropriate equipment and drugs for resuscitation in the event of an
anesthetic catastrophe (Fig. 5.2).2 Judicious sedation and analgesia,
along with a kind bedside manner, will prepare most patients for
regional anesthesia but will also avoid a depth of anesthesia that
precludes feedback to the anesthesiologist (Fig. 5.3).
Intravenous catheter and balanced salt solution Blood pressure
monitoring
Pulse oximetry monitoring
Electrocardiographic monitoring (necessary for patients with cardiac
disease or dysrhythmias) Supplemental oxygen by face mask or
cannula Medications for anxiolysis (e.g., midazolam, propofol,
dexmedetomidine) and analgesia (fentanyl) Equipment for
airway management (e.g., laryngoscopes, endotracheal tubes, bag
and face mask, oral airways, laryngeal airways) Pharmaceuticals
for resuscitation (e.g., propofol, epinephrine, 20% intralipid,
vasopressin, amiodarone) The skin is prepared in sterile fashion,
and subcutaneous local anesthetic solution is injected at the site
at which the block needle is to be inserted. The ultrasound probe
is covered with a sterile, transparent membrane for single-shot
blocks, and the anesthesiologist dons sterile gloves. Most authors
recommend use of short-bevel needles, as it appears to be more
difficult to penetrate the perineurium with this type of needle,3
although clinical outcome data are lacking.
Antiseptic for skin preparation
Transparent, sterile membrane to cover ultrasound probe Lidocaine
in 3-or 5-mL syringe with 27G or 30G needle for skin anesthesia
Short-bevel block needle, length appropriate to depth of nerve
The local anesthetic solution utilized will depend on the
anesthesiologist’s intent. Blocks intended for rapid onset and
short duration may be conducted with mepivacaine, lidocaine, or
chloroprocaine, whereas longer acting blocks will require
ropivacaine or bupivacaine. Mixtures probably add little to speed
onset while significantly reducing the duration of long-acting
agents. Although it is axiomatic that injection must not proceed
when injection pressures are high, it is not yet clear whether
pressure monitoring will influence the occurrence of nerve injury
in the clinical setting.4
Appropriate volume of desired local anesthetic solution Desired
additives for local anesthetic (e.g., clonidine, dexamethasone,
buprenorphine, epinephrine) Peripheral nerve stimulator,
attached to block needle and patient Pressure-monitoring device
in line with injectate system, if desired The choice of an
ultrasound system depends on the resources and needs of the
user. The system should be portable with high definition and a
choice of probes to allow imaging of both superficial (high-
frequency probe) and deep (low-frequency probe) nerves (Fig.
5.4).5
Portable, high-fidelity ultrasound machine Ultrasound gel, preferably
sterile
Ultrasound probe appropriate for nerve to be blocked For catheter
insertion, a higher degree of sterility is necessary since it is an
indwelling device, with relatively high rates of colonization.6
This should mandate the use of sterile gowns, masks, drapes, as
well as the catheter system itself (Fig. 5.5). The choice remains
for the individual anesthesiologist as to whether peripheral nerve
stimulation is utilized for confirmation of the target nerve and, if
so, whether a stimulating or nonstimulating type catheter is
preferred.
Perineural needle/catheter set (stimulating or nonstimulating) Sterile
drapes
Sterile gown, mask, gloves
Sterile adhesives to hold catheter in place Sterile, transparent
membrane to cover catheter site Peripheral nerve stimulator, if
desired
References
1. Chan VWS, Perlas A, McCartney CJL, et al. Ultrasound guidance
improves success rate of axillary brachial plexus block. Can J
Anaesth. 2007;54:176–182.
2. Tucker MS, Nielsen KC, Steele SM. Nerve block induction rooms
—physical plant setup, monitoring equipment, block cart and
resuscitation cart. Int Anesth Clin. 2005;43:55–68.
3. Selander D, Dhuner KG, Lundborg G. Peripheral nerve injury due
to injection needles used for regional anesthesia. An experimental
study of the acute effects of needle point trauma. Acta Anaesth
Scand. 1977;21:182–188.
4. Hadzic A, Dilberovic F, Shah S, et al. Combination of intraneural
injection and high injection pressure leads to fascicular injury and
neurologic deficits in dogs. Reg Anesth Pain Med. 2004;29:417–
423.
5. Gray AT. Ultrasound-guided regional anesthesia: current state of the
art. Anesthesiology. 2006;104(2):368–373.
6. Capdevila X, Pirat P, Bringuier S, et al. Continuous peripheral nerve
blocks in hospital wards after orthopedic surgery: a multicenter
prospective analysis of the quality of postoperative analgesia and
complications in 1416 patients. Anesthesiology. 2006;103:1035–
1045.
6
Principles of
Sonography
PAUL E. BIGELEISEN, DANIELLA SMITH, AND

STEVEN L. OREBAUGH

The frequency of medical ultrasound ranges between 2 and 20 MHz.


The speed of sound in tissues is about 1,500 m/sec. The average
wavelength of the ultrasound beam in this band is less than 1 mm. This
limits the use of ultrasound to structures that are 1 mm in diameter or
larger (Fig. 6.1). Most nerves of interest range from 2 to 10 mm. Veins
and arteries of interest are typically 3 to 15 mm.

Many factors contribute to the quality and resolution of the ultrasound


image. In general, higher frequency probes generate higher resolution
images. Unfortunately, the signal strength or intensity (I) of high-
frequency ultrasound waves (10 to 20 MHz) is rapidly attenuated in
tissue. Thus, high-frequency probes are best suited for structures less
than 4 cm deep from the skin. For deeper structures, probes in the 2 to 5
MHz range are more useful (Fig. 6.2). The speed of sound in air and
tissues varies from 300 m/sec in air to 4,000 m/sec in bone. In soft
tissues, blood, and cystic fluid, the speed of sound is about 1,500 m/sec.
Most ultrasound platforms assume an average speed for beam
propagation of about 1,500 m/sec. Because the speed of sound differs
slightly in different types of soft tissues, small imaging artifacts may
occur as the beam passes from one tissue type to another. This occurs
most frequently when the beam passes from soft tissue through blood or
cystic fluid.

The ultrasound beam may be refracted as it passes through tissue. When


this occurs, a nerve or other organ may appear at a different anatomical
location than its actual site. This is the same phenomenon responsible
for the apparent bending of your forearm when you place it in a bucket
of water (Fig. 6.3). Notice that the needle appears to be bent as it enters
a fluid-filled cyst in Figure 6.3B. Fat globules below the skin, in the
muscle, and around nerves are about 1 mm in diameter. These globules
(Fig. 6.4A) serve as scattering and diffraction sites for the incident and
reflected ultrasound beam and cause a speckled appearance in the image
(Fig. 6.4B, C). This is called speclation. For these reasons, obese
patients can be very difficult to image (Fig. 6.5A, B).
The image formed of a nerve on ultrasound is very sensitive to the angle
of incidence of the beam relative to the nerve. This is referred to as the
angle of insonation. Sometimes, changing the angle of insonation by
only a few degrees can bring the nerve into focus. Diffraction and
scattering, as well as the heterogeneous three-dimensional structure of
the nerves and their surrounding tissues, are thought to cause this
phenomenon. An analogy is shown in Figure 6.6A. Here, light is shown
on a three-dimensional icon. Varying the direction of the incident light
beam by only a few degrees will result in a radically different shadow in
any direction. To obtain the optimal ultrasound image, the image should
be centered on the screen by sliding or rotating the probe on the patient’s
skin. For deep structures, compressing the tissue may improve the
image. Once these maneuvers have been completed, toggling will
produce the best image (Fig. 6.6Gi). Notice how the nerve disappears in
the ultrasound image of Figure 6.6G(ii) with additional toggling.
Modern platforms allow the user to adjust the brightness (gain) of the
entire image or more superficial (near field) and deep (far field)
structures. Increasing the gain makes the entire image whiter. Increasing
the gain too much creates a snowy background in which all of the
structures become indistinguishable. In general, the gain should be set so
that most of the background is black and only the structures of interest,
such as nerves and vessels, are easily seen. Many machines have an
autogain button that adjusts the gain for the user.
Modern machines also allow the user to adjust the contrast. The formal
term for contrast is dynamic range compression. Increasing the dynamic
range compression makes the white images whiter and the black images
blacker (Fig. 6.7A). This may bring the edges of an anatomic structure
into better view. Decreasing the dynamic range compression makes
everything in the image a more homogeneous grey.
All machines allow the user to adjust the depth to which the probe
penetrates. Whenever possible, the depth should be set to the shallowest
setting in which the structures of interest are all imaged. Some machines
come with a preset focal zone, and some machines allow the user to set
multiple focal zones. In general, the focal zone should encompass the
area of greatest interest in the image. The near focal zone has the
greatest resolution and is referred to as the Fresnel zone. The deeper
focal zone (far field) is referred to as the Fraunhofer zone. This zone has
poorer resolution (Fig. 6.8A). Most machines allow the user to adjust the
depth of the focal zone or to set multiple focal zones (Fig. 6.8B) by
adjusting the time delays between transducer elements in the probe.
Arteries can usually be distinguished by their pulsatile nature. Veins can
be distinguished by their compressibility. Pressing on the skin with the
probe will usually cause the vein to collapse. Color-flow Doppler
imaging can also be used to identify and distinguish arteries and veins.
Blood flowing perpendicular to the probe is black (Fig. 6.9B). By
convention, blood flowing toward the probe is colored red (Fig. 6.9C).
Blood flowing away from the probe is colored blue (Fig. 6.9C). Velocity
gates can be set to measure the flow velocity. High velocities are usually
arteries; low velocities are usually veins.
Transducer elements can be arranged in linear or curved arrays (Fig.
6.10). Linear arrays create rectangular images and are most useful for
superficial structures. Curved arrays create wedge-shaped images and
are most useful for deeper structures. When a curved array is used, the
beam disperses laterally, resulting in a lower resolution than a linear
array. A phased array retains the transducer elements in a linear
arrangement, but the elements fire in sequence, creating a phase delay
between each element. The net result is a wedge-shaped image from a
set of linear transducers. Because this signal is averaged, its resolution is
also lower than a standard linear array. The array in a transducer can also
be programmed to fire with multiple phase delays. This creates a simple
two-dimensional tomogram called a compound beam image. This signal
can be processed to highlight nerves (Fig. 6.11A). The image formed is
less sensitive to the angle of incidence because the beam comes from
many directions. Thus, even though transducer alignment may not be
optimal, a recognizable image may still be produced (Fig. 6.11B). The
resolution in the compound beam mode is blurred compared to a
standard linear or curved array without this type of signal averaging.
Most probes have transducers that emit the highest amplitude of their
ultrasound wave at a specific fundamental frequency. Higher harmonics
of this frequency are also emitted at lower amplitudes (Fig. 6.12). By
listening for the echo at these higher harmonic frequencies, image
resolution can be enhanced. Because the harmonics are very low
amplitude, only transducers that have sufficient power output can be
used for harmonic imaging. This type of image enhancement is referred
to as tissue harmonic imaging. It is very useful in cardiac imaging, but it
has not been shown to be useful in nerve imaging at this time.
Compound beam imaging negates most of the advantages of tissue
harmonic imaging.

Modern probes operate in pulsed mode. The transducer elements in the


probe are used as both transmitters and receivers. The transducer
elements emit a short ultrasound burst and then wait for the echo before
emitting another ultrasound burst. This allows the probe to be smaller
compared to continuous wave probes where there are separate emitters
and transducers (Fig. 6.13).

When a needle is inserted into tissue perpendicular to the ultrasound


beam, it is a good reflector and easy to image. Ghosts on the deep side of
the needle are caused by continued needle vibration after the ultrasound
wave strikes the needle (Fig. 6.14). These reverberations return to the
receiver later than the first volley of echoes. Consequently, they are seen
as deeper in the tissue. In some cases, it may be necessary to insert the
needle nearly parallel to the beam in order to reach the targeted organ. In
this case, most of the echo is lost and the needle image is much fainter
(Fig. 6.15).

Above the collar bone, nerves are usually dark (hypoechoic) (Fig. 6.16).
Nerves located below the collarbone are usually white (hyperechoic)
(Fig. 6.5B). The reasons for this dichotomy are not known but are
thought to be related to the depth of the nerves, the amount of fat around
the nerves, and the relative amounts of fat and stroma within the nerves
themselves. On ultrasound cross section, nerves are round, hypo-or
hyperechoic, reticulated structures. When imaged along their long axis,
nerves appear as linear, hypo-or hyperechoic streaks. Bones are
hyperechoic (Fig. 6.5B) and usually very bright white. Arteries and
veins are black unless color-flow Doppler imaging is used (Fig. 6.9). If
the transducer is perpendicular to the blood flow, arteries and veins will
be black even with color-flow Doppler imaging.

Most nerves have some fascia around them. There is usually a potential
space between the fascia and the epineurium. When a needle punctures
the fascia, local anesthetic can usually be deposited between the fascia
and nerve (Fig. 6.17A). This creates a black (hypoechoic) ring around
the nerve. In some cases, the fascia adheres to the epineurium or is
missing. In that case, the needle may puncture the nerve, and the nerve
will swell as the local anesthetic is injected (Fig. 6.17B).
Artifacts are inherent to ultrasound imaging. Typical artifacts mentioned
earlier include refraction, diffraction (speclation), and reverberation.
Another common artifact is called a side lobe artifact. When the beam is
formed, most of the output is focused in front of the transducer, but a
small amount of the beam is transmitted as side lobes (Fig. 6.18A, B).
Figure 6.18A shows a transducer in a plastic cup of water. Figure 6.18B
shows side lobes that arise from multiple reflections (yellow and purple
arrows in Fig. 6.18A) at the corner of the cup and mirroring of the
bottom (mb in Fig. 6.18B). Side lobes can also be reflected back to the
transducer by tissue that is lateral to the main beam. These artifacts will
appear in the image as if they arose from tissue in front of the transducer.

The imaging algorithm of the machine assumes that echoes that return to
the transducer later arise from deeper tissues (far field). Because the
beam is attenuated as it travels into deeper tissues, these echoes are
weaker, and the resulting far-field image is faded relative to the near-
field image. The algorithm compensates for this by brightening the far-
field image. This is called time gain compensation (Fig. 6.19). In some
cases, the tissue in the near field may be a blood vessel or cyst. When
the tissue in the near fields is fluid, the beam will pass through the blood
or water with little attenuation. When this beam strikes a strong reflector
in the far field, such as a vessel wall, its echo will be minimally
attenuated. This echo, which arises from the far field, is brightened by
the algorithm even though it has been little attenuated by the near-field
blood vessel or cyst. When this occurs, the tissue deep to the cyst or
vessel may appear artificially bright. In some cases, time gain
compensation may cause an artifact deep to a vessel to appear like a
nerve, when in fact no nerve exists deep to the vessel. In Figure 6.19, the
brachial plexus, axillary artery, and vein are real. The area inside the
blue outline looks like nerve tissue, when in fact it is time gain
compensation artifact. The deep vessel wall appears hyperechoic due to
time gain compensation.

Shadowing occurs when a strong reflector prevents the beam from


penetrating the reflector. Strong reflectors include bone, fascia, and
calcified tissues. Notice the shadowing (hypoechoic areas) deep to the
transverse processes in Figure 6.20. Mirror artifacts occur when the
beam encounters strong reflectors deep to the structure of interest (Fig.
6.21A, B). Echoes that reverberate between the tissue of interest and the
strong reflector will return at a later time to the transducer. These later
echoes are imaged as an artifactual mirror image deep to the true
structure (Fig. 6.21B).

Technical Tips 1. Stimulation with a peripheral nerve stimulator is not necessary


if the operator is certain of the nerve’s identity on ultrasound. When a stimulator
is used, it should only be used to confirm the target nerve, then it can be turned
off. Because the local anesthetic is injected around or into the target nerve under
ultrasound visualization, there is no need to “titrate” the current to the twitch.
2. It is challenging to keep the needle perfectly parallel to the long axis
of the transducer. Frequent fine adjustment of the transducer may be
necessary along with switching the line of site of the operator from
the ultrasound screen to the site of needle insertion. Some people
prefer to use a needle guide to help keep the needle aligned with the
ultrasound beam.
3. As local anesthetic is injected, each increment should cause visible
expansion of the tissues at the tip of the needle. This provides
evidence that the tip of the needle is not intravascular.
4. If all of the local anesthetic solution seems to accumulate on only one
side of the target nerve, the needle should be gently advanced or
moved to another site around the nerve, to allow accumulation of the
solution around the entire nerve (the “halo” effect). This minimizes
block setup time. It should be clear that each aliquot of local
anesthetic injected should cause distension of the tissues at the tip of
the needle. It is not necessary to restimulate as the needle is moved
around the nerve, but the patient should be assessed for pain in the
territory of the nerve.
5. As with any block guided by the nerve stimulator, mild injection
paresthesias may occur during injection of local anesthetic in blocks
carried out by ultrasound guidance. This should be differentiated
from the severe pain that is more likely indicative of intrafascicular
injection.
6. Time gain compensation may make it difficult to see the posterior
cord or radial nerve deep to the axillary artery.
7. Shadowing will make it difficult to image nerves deep to bones or
thick fascia. When this occurs, it is useful to image the nerve where
the nerve lies superficial or lateral to the bone. For example, the
sciatic nerve can be easily visualized from the lateral, posterior, or
medial approach, but not with the transducer anterior to the femur.

Additional Reading Jan J. Medical Image Processing, Reconstruction


and Restoration. New York: Taylor and Francis; 2006.
7
Three-and Four-
Dimensional
Ultrasound in
Neuraxial Anesthesia
DAVID BELAVY

There is scant literature on the use of three-dimensional (3D)/four-


dimensional (4D) ultrasound in regional anesthesia and only one
publication that explores the potential use of 4D ultrasound for neuraxial
procedures.1–13
Neuraxial 4D ultrasound builds on the knowledge of 2D ultrasound
scanning procedures, sonoanatomy, and needle technique. There are
potential uses in preprocedural scanning and for needle guidance. The
use of 4D ultrasound greatly depends on how the images are presented
to the operator.
This chapter discusses how 3D/4D ultrasound data is acquired and
presented, how the spine appears using this technology, and how it could
be applied to epidural and spinal techniques.
3D/4D image acquisition
In two-dimensional (2D) ultrasound imaging, a 3D appreciation of the
anatomy is built in the sonographer’s mind by examining structures in
multiple 2D views, or sections, acquired through movement of the
probe. Abdominal and obstetric curvilinear low-frequency transducers (2
to 5 MHz) are recommended for neuraxial imaging in adults.14 Their
low frequency improves imaging of deeper structures, and curved probes
provide a wider field of view. These 2D ultrasound probes typically
consist of a single line of up to 256 piezoelectric crystals, which
examine a 2D area and produce a B-mode image. The two dimensions
are the depth and width of the imaging field. In contrast, static 3D
ultrasound captures a volume of data that may then be digitally
manipulated and displayed without movement of the probe. 3D imaging
adds an elevation dimension to the depth and width of 2D imaging to
describe a 3D volume of data. Each data element, or voxel, represents
the echogenicity of that point in the 3D volume of tissue.
For the purposes of neuraxial ultrasound, there are two methods that
3D/4D ultrasound systems use to examine a 3D volume.15

1. Mechanically steered arrays. In these systems, a standard curvilinear


transducer with a single line of piezoelectric crystals is contained
within a fluid-filled housing. The transducer is mechanically steered
within the housing and sweeps over the area examined, building a
volume of data (Fig. 7.1A). The time taken to acquire one volume of
data is the time taken for the transducer to cover the sweep angle.

2. 2D phased array transducer. In contrast to conventional transducers


that contain a line of piezoelectric crystals, these transducers contain
a 2D array of crystals. From this array, the ultrasound beam is
electronically swept through the volume being examined, resulting in
a more rapid examination. These transducers require more
piezoelectric crystals; an example is the Philips X6-1 xMATRIX
Array transducer which contains 9,212 elements (Fig. 7.1B).
Both techniques produce data covering a pyramidal volume (Fig. 7.2).
Real-time 3D (RT3D) or 4D ultrasound simply refers to repeated
acquisition of 3D ultrasound volumes over time, producing a real-time
examination. The image resolution and the dimensions of the volume of
interest (imaging depth, sector width, and sweep angle) are key
determinants of the frame rate in 4D ultrasound imaging.

3D/4D image display


The data acquired during a 3D scan can be examined on the ultrasound
machine or off-line on a workstation. The two main postprocessing
techniques relevant for displaying neuraxial images from a 3D dataset
are multiplanar reconstruction (MPR) and 3D rendering.
Multiplanar reconstruction
MPR allows any slice taken from a 3D dataset to be displayed, but
initially, three orthogonal planes are chosen. MPR images are routinely
displayed from spiral computed tomography (CT) and magnetic
resonance imaging (MRI) scans in coronal and sagittal planes in addition
to the traditional transverse or axial plane. Similarly, with 3D ultrasound,
two or three orthogonal images are displayed from the acquired data
(Fig. 7.3). The images of the three planes are usually designated A, B,
and C. The first plane is the familiar 2D ultrasound image. Plane B or y
is perpendicular to the first. If plane A is positioned over the long axis of
a structure, plane B would display its short axis. Plane C or z is
perpendicular to the other images and represents echoes from a plane
positioned a distance from the probe. Remembering the terms can be
aided by using the mnemonic “Actual, Bisecting, and Coronal.” Figures
7.3 and 7.4 demonstrate these views at the lumbar vertebral level.
MPR using 4D ultrasound displays two or more planes in real time.
Most commonly, two orthogonal planes are displayed. The terminology
for this type of imaging varies; for example, Philips machines use
xPlane mode, whereas GE calls it MPR. These imaging modes have
potential application in neuraxial ultrasound and will be discussed in the
following text.
3D rendering techniques
There are numerous computational techniques to display 2D images of
3D datasets. The majority of 3D techniques involve ray casting. An
imaginary ray is passed from the viewing screen through the dataset.
Surface rendering or shaded-surface display uses algorithms to
determine the surfaces in the dataset and represents them as shaded 2D
polygons arranged in 3D space. The software can then apply lighting
effects to the model.16
Maximal intensity projection projects rays through the dataset. The point
of maximal intensity along that ray is displayed and creates x-ray–like
images. This rendering technique is well suited to CT angiography,
where the point of maximal intensity will be the contrast-filled lumen of
a vessel.16 Minimal intensity projection is a similar technique except that
black or empty voxels are rendered as solid and white voxels are shown
empty. This is useful to demonstrate the shape or volume of an echo-free
structure.
Volume rendering is the most frequently used technique in ultrasound. A
virtual ray is passed from the viewer through the 3D volume of data by
the software. When that ray passes through an echogenic voxel, the point
is displayed on the screen. In vivo, surfaces are most easily recognized
when surrounded by echolucent fluid. The structures best displayed by
3D ultrasound using surface rendering are heart valves surrounded by
blood or the fetus surrounded by amniotic fluid.15 How the structures are
displayed depends heavily on a variety of user-defined controls.
Acquisition and optimization of 3D/4D imaging of the spine
A 3D/4D study of the spine begins with an optimal 2D image, and all of
the usual image optimization procedures need to be followed prior to
beginning 3D/4D imaging. A poor-quality 2D image usually results in a
much worse 3D image. The structure of interest should be positioned in
the center of the 2D image and the probe frequency, depth, focal zone,
gain, compression, and gray scale optimized. The time gain
compensation can be adjusted to decrease gain from nearer structures
that produce near-field clutter in 3D-rendered images, but this is less
important in MPR images. In addition, near-field clutter can be removed
by cropping during postprocessing. Techniques to reduce artifacts such
as frequency compounding, spatial compounding, and tissue harmonic
imaging may also improve the image at the cost of frame rate. Adjusting
compression may improve the contrast of the image.
Elevation compounding or volume contrast imaging is an additional
option to improve image quality in 2D imaging that is available with
some 3D systems. This is a 3D technique that uses a small sweep angle
(elevation) to image a limited number of adjacent 2D slices, essentially
producing a thick 2D image. The software compares the adjacent slices
to remove speckles and enhance the signal-to-noise ratio.
The 3D examination can proceed in a number of ways: acquisition of a
single 3D volume and subsequent postprocessing, real-time MPR
imaging, and real-time 4D rendering. The terminology and methods of
acquiring these images varies between manufacturers.
Acquisition of a single 3D volume
A single 3D volume can usually be quickly acquired without much
adjustment. A region-of-interest box may be used to limit the
unnecessary data recorded and speed that acquisition time. Some
systems allow the user to switch to an MPR mode where at least two
perpendicular planes are displayed. Using the two perpendicular 2D
images, the three dimensions of the volume of interest can be defined
(height, width, and elevation). After the volume of interest is defined, a
single 3D volume can be acquired including all the relevant anatomy.
The image shows the MPR data after acquisition.
It is difficult to produce the reconstructed image shown (Fig. 7.3) in the
majority of circumstances. To render structures such as bone clearly, the
image must be manipulated to show only the echogenic bone and
remove other tissues. Numerous controls are available in 3D volume
rendering to achieve this goal and are illustrated in images. However, a
high-quality image is rarely achieved because the overlying tissues are
not echo free.
Threshold refers to the minimal echo intensity required for a voxel to
render as a solid point. Increasing the threshold removes less-intense
echoes and can reveal deeper structures.
Opacity (and the inverse, transparency) is the degree to which structures
close to the viewer obscure structures that are farther away. Decreased
opacity (or increased transparency) allows deeper structures to be seen
through more superficial structures. High levels of opacity produce
images that are similar to surface rendering, whereas low levels of
opacity are useful to see bright deeper structures that would otherwise be
obscured.
Window width (contrast) and level (brightness) should be optimized
during 2D imaging but can be further adjusted during volume rendering
and determine the intensity of each displayed pixel.
Smoothing performed by rendering algorithms can alter the average gray
values between nearby voxels to make surfaces appear smoother and to
potentially improve the anatomical representation. This results in loss of
detail but may make the image easier to interpret. Reducing smoothing
makes surfaces appear more textured.
Even with all these adjustments, it is difficult to render the vertebrae
using ultrasound.
Real-time MPR imaging
MPR or xPlane imaging modes are more useful in neuraxial imaging,
both during the scan-and-mark procedure and for needle insertion.3 The
A plane is positioned as in the 2D examination and the position of the B
plane can be adjusted to give the perpendicular view. An acoustic
window can be assessed in two planes at once, which is potentially
useful in the thoracic spine where the windows can be small. The two
planes can be seen without changing probe position, speeding the
examination.
Compared to 2D imaging, the frame rate is reduced with MPR imaging.
The frame rate reduction is greater with mechanically steered arrays than
phased array (electronically steered) probes. The size of the region of
interest is an important determinant, as the frame rate is inversely
proportional to the time taken to acquire one volume of data. The sweep
angle should be minimized, particularly for mechanically steered
transducers.
Decreasing the resolution of the image may improve the frame rate but
decreases image quality and probably needle visibility.
Real-time 4D rendering
Although very useful in echocardiography, 4D rendering has a limited
role in neuraxial imaging because of the poor-quality rendering of 3D
structures surrounded by echogenic tissues.
Real-time 4D needle insertion for neuraxial anesthesia
The landmark-guided midline or paramedian needle techniques offer the
most direct approach to the epidural or subarachnoid spaces. Real-time
ultrasound-guided combined spinal-epidural procedures were first
described as an out-of-plane technique with the needle inserted in the
midline and the probe in a paramedian position.14 For in-plane needle
visualization, the needle may be inserted using a paramedian
approach.14,17,18
We performed a feasibility study on epidural insertions in embalmed
cadavers using real-time 4D ultrasound with a mechanically steered
transducer.1
We attempted to identify the needle using 3D rendering but found that
the echogenicity of the paraspinal musculature completely obscured the
16G Tuohy needle. The speckle from the erector spinae muscle appear
as a solid mass of tissue in 3D rendering, and the needle could not be
recognized.
We found that needle visibility was best in the primary imaging plane (A
plane or x) during MPR imaging. Needle visibility was poor outside of
the primary imaging plane. Compared to 2D imaging, image resolution,
frame rate, and needle visibility with 4D MPR is inferior.
As a result of this experience, we observed that the potential future use
of 4D ultrasound is likely to involve real-time MPR. The technique is
very similar to the real-time 2D paramedian in-plane technique, but a
perpendicular plane is also displayed on the ultrasound screen. This
technique offers the significant potential advantage by improving the
orientation of the operator. The primary imaging plane (A or x) is
positioned over the laminae in a longitudinal paramedian position such
that the intrathecal space and posterior border of the vertebral body is
visible. The perpendicular (B or y) plane can then be used to medially or
laterally adjust the angle of the transducer so that the probe is aimed
more medially over the lamina, nearer the base of the spinous process.
The needle is then inserted, in-plane, under the inferior end of the probe,
similar to the 2D technique. The needle can then be passed in-plane
toward the target. In our study, medial or lateral adjustments were made
to the needle path in half of the approaches as a result of the information
derived from the perpendicular MPR plane. However, at the time of
writing, no publications have discussed the clinical use of this technique.
Challenges for 3D/4D ultrasound in neuraxial anesthesia and potential solutions
There are numerous challenges to overcome prior to the use of 3D
ultrasound in neuraxial anesthesia.

Image Quality and Frame Rate


Compared to 2D ultrasound, image resolution is inferior in both MPR
and 3D rendering, which adversely impacts on anatomical clarity and
needle visibility. Mechanical transducers also have a much lower frame
rate, as the transducer must sweep through the entire volume to generate
images. Lower frame rates affect real-time needle insertion when the
movement of tissues used to identify needle location becomes harder to
recognize. Minimizing the volume of interest and reducing the sweep
angle results in higher frame rates. As matrix transducer technology
improves and imaging in multiple planes at high frame rates becomes
possible, these disadvantages should be overcome.

Needle Visibility and Path Length


Neuraxial blocks are a relatively deep block, typically 4 to 6 cm but
sometimes significantly more. The 4D ultrasound probes also have a
relatively large footprint. With larger probe footprint and significant
target depth, the needle must be inserted further from the center of the
probe for good needle visualization using an in-plane technique, even
when obesity is not a factor. As a result, the needle follows a longer
paramedian path during real-time ultrasound-guided needle insertion.
Even 16G Tuohy needles can prove difficult to see, and tissue movement
must be relied upon for determining needle location. Needle visibility
may be improved in the future with echogenic Tuohy needles. For spinal
anesthesia, a needle-through-needle combined spinal-epidural technique
is likely to improve needle visibility and manage the flexibility of long
spinal needles. Stereotactic ultrasound guidance, where needle location
is displayed on-screen independent of needle visibility, is a technologic
direction that may overcome these difficulties.

Assistance and Identifying the Epidural Space


Performing ultrasound-guided neuraxial anesthesia involves several
physical tasks, including holding the probe, inserting the needle,
applying a loss-of-resistance technique, feeding a catheter, and
controlling the ultrasound machine. It is impractical for all of these tasks
to be performed by one clinician alone. Several options are available to
manage the tasks.
The simplest solution is to have a scrubbed assistant hold the ultrasound
probe while the anesthesiologist performs the procedure.
There are several alternative techniques of identifying the epidural space
that may be suitable for ultrasound-guided techniques. The hanging drop
technique may be used, or mechanical devices such as the Episure
AutoDetect LOR Syringe (Indigo Orb, Irvine, CA) or the Epidrum
(Exmoor Innovations, Somerset, UK) are available.19,20 A bag of saline
positioned above the patient and connected with sterile tubing to the
Tuohy needle will begin to flow when the epidural space is entered,
demonstrating a loss of resistance; however, there is little literature on
this technique. In all of these techniques, the primary operator will be
watching the ultrasound screen so an assistant needs to observe for loss
of resistance. Alternatively, an assistant may perform the loss-of-
resistance technique at some distance from the primary operator by
using low-compliance extension tubing, such as that used for invasive
pressure monitoring. Finally, new techniques such as optical
spectroscopy are also potential solutions.21

Ultrasound Gel
The composition of ultrasound gels is often unknown but may include
low concentrations of propylene glycol or glycerol that are both known
to have neurolytic properties at high concentrations. There is a risk of
introducing ultrasound gel into the epidural or intrathecal spaces if it
comes in contact with the needle.22 To avoid this, small amounts of gel
should be used with particular care to ensure that the needle does not
contact the gel, or alternative ultrasound conduction media can be
considered such as repeated application of saline to the skin.

Training and Certification


Guidelines are only recently being developed for training in ultrasound-
guided regional anesthesia.23 It has been suggested that experience of 40
cases is required to attain competence with 2D scanning of the lumbar
spine. Neuraxial ultrasound-guided procedures are technically
demanding and, as 4D ultrasound is a further increment in cognitive and
technical challenge, substantially more experience may be required. As
the role of 4D ultrasound remains undefined, the training requirements
for performing procedures are unknown.

Clinician Resistance
Landmark-guided neuraxial anesthesia techniques are well established
and 2D ultrasound is not required for successful anesthesia in the
majority of patients.

Cost
3D ultrasound equipment is substantially more expensive than 2D but, in
time, is likely to become more available.
Conclusions
The 2D ultrasound examination of the spine has become increasingly
useful in anesthesiology. 3D/4D technology may offer some advantages
with MPR to speed the examination and improve needle direction during
real-time procedures. The limitations in frame rate, resolution, and cost
may be overcome as technology develops. The relevance of 4D
ultrasound needs to be demonstrated in well-conducted clinical studies.

Acknowledgements
I wish to thank Iain Dunn and Philips Healthcare UK for loaning the
ultrasound machine that was used to produce many of the pictures for
this chapter.

References
1. Belavy D, Ruitenberg MJ, Brijball RB. Feasibility study of real-
time three-/four-dimensional ultrasound for epidural catheter
insertion. Br J Anaesth. 2011;107:438–445.
2. Clendenen SR, Riutort KT, Feinglass NG, et al. Real-time three-
dimensional ultrasound for continuous interscalene brachial plexus
blockade. J Anesth. 2009;23:466–468.
3. Clendenen SR, Riutort K, Ladlie BL, et al. Real-time three-
dimensional ultrasound-assisted axillary plexus block defines soft
tissue planes. Anesth Analg. 2009;108:1347–1350.
4. Clendenen SR, Robards CB, Clendenen NJ, et al. Real-time 3-
dimensional ultrasound-assisted infraclavicular brachial plexus
catheter placement: implications of a new technology. Anesthesiol
Res Pract. 2010;2010.
5. Clendenen SR, York JE, Wang RD, et al. Three-dimensional
ultrasound-assisted popliteal catheter placement revealing aberrant
anatomy: implications for block failure. Acta Anaesthesiol Scand.
2008;52:1429–1431.
6. Feinglass NG, Clendenen SR, Torp KD, et al. Real-time three-
dimensional ultrasound for continuous popliteal blockade: a case
report and image description. Anesth Analg. 2007;105:272–274.
7. Foxall GL, Hardman JG, Bedforth NM. Three-dimensional,
multiplanar, ultrasound-guided, radial nerve block. Reg Anesth Pain
Med. 2007;32:516–521.
8. Grau T, Leipold RW, Conradi R, et al. Ultrasound control for
presumed difficult epidural puncture. Acta Anaesthesiol Scand.
2001;45:766–771.
9. Grau T, Leipold RW, Conradi R, et al. Ultrasound imaging
facilitates localization of the epidural space during combined spinal
and epidural anesthesia. Reg Anesth Pain Med. 2001;26:64–67.
10. Grau T, Leipold RW, Conradi R, et al. Efficacy of ultrasound
imaging in obstetric epidural anesthesia. J Clin Anesth.
2002;14:169–175.
11. Grau T, Leipold RW, Fatehi S, et al. Real-time ultrasonic
observation of combined spinal-epidural anaesthesia. Eur J
Anaesthesiol. 2004;21:25–31.
12. Karmakar M, Li X, Li J, Sala-Blanch X, et al. Three-
dimensional/four-dimensional volumetric ultrasound imaging of the
sciatic nerve. Reg Anesth Pain Med. 2012;37:60–66
13. Kil HK, Cho JE, Kim WO, et al. Prepuncture ultrasound-measured
distance: an accurate reflection of epidural depth in infants and
small children. Reg Anesth Pain Med. 2007;32:102–106.
14. Chin KJ, Karmakar MK, Peng P. Ultrasonography of the adult
thoracic and lumbar spine for central neuraxial blockade.
Anesthesiology. 2011;114:1459–1485.
15. Hoskins PR, Martin K, Thrush A. Diagnostic Ultrasound Physics
and Equipment. 2 ed. Cambridge University Press; 2010.
16. Calhoun PS, Kuszyk BS, Heath DG, et al. Three-dimensional
volume rendering of spiral ct data: theory and method.
Radiographics. 1999;19:745–764
17. Karmakar MK, Li X, Ho AM, et al. Real-time ultrasound-guided
paramedian epidural access: evaluation of a novel in-plane
technique. Br J Anaesth. 2009;102:845–854.
18. Tran D, Kamani AA, Al-Attas E, et al. Single-operator real-time
ultrasound-guidance to aim and insert a lumbar epidural needle.
Can J Anaesth. 2010;57:313–321.
19. Riley ET, Carvalho B. The Episure syringe: a novel loss of
resistance syringe for locating the epidural space. Anesth Analg.
2007;105:1164–1166, table of contents.
20. Sawada A, Kii N, Yoshikawa Y, et al. Epidrum(®): a new device to
identify the epidural space with an epidural Tuohy needle. J Anesth.
2012;26:292–295.
21. Rathmell JP, Desjardins AE, van der Voort M, et al. Identification of
the epidural space with optical spectroscopy: an in vivo swine
study. Anesthesiology. 2010;113:1406–1418
22. Belavy D. Brief reports: regional anesthesia needles can introduce
ultrasound gel into tissues. Anesth Analg. 2010;111:811–812.
23. Sites BD, Chan VW, Neal JM, et al. The American Society of
Regional Anesthesia and Pain Medicine and the European Society
Of Regional Anaesthesia and Pain Therapy Joint Committee
recommendations for education and training in ultrasound-guided
regional anesthesia. Reg Anesth Pain Med. 2009;34:40–46.
8
Anatomical
Anomalies in
Ultrasound
Simulation
JEREMY KAPLOWITZ AND PAUL E.

BIGELEISEN

Ultrasound-guided regional anesthesia has provided the regional


anesthesiologist with the ability to visualize pertinent nerves and local
anesthetic spread. One consequence of this newfound “vision” in
regional anesthesia is the ability to visualize anatomic variations and
sonopathology. These findings may impact the performance of your
regional anesthetic and may lead to an altered anesthetic technique.
Currently, there is some debate about the role of regional
anesthesiologists in diagnosing pathologic abnormalities while
performing ultrasound-guided nerve blocks. Although this question has
important medicolegal and clinical implications, it is beyond the scope
of this chapter to delve into this issue. Nonetheless, practitioners agree
that knowledge of the common anatomical variations is essential in
allowing the practitioners to perform effective and safe regional
anesthesia.
In order to appreciate the anatomic variants, we first discuss some of the
common anatomic artifacts that frequently confuse novices. Then we
discuss some of the common anatomic variants. Last, we highlight some
of the more common sonopathologic findings that you may encounter
while performing ultrasound-guided regional anesthesia.
Anatomic artifacts
There are various anatomic artifacts that one may encounter while
performing a regional anesthetic block. Anatomic artifacts occur when
tissue structures are confused with each other. These are usually
structures that may appear similar to nerves. Some common examples
include tendons, muscles, lymph nodes, and blood vessels.
Tendons
Tendons may be confused with nerves that have a hyperechoic
appearance. This artifact confusion most commonly occurs while
performing sciatic nerve blocks at the popliteal fossa or during selective
nerve blocks in the arm and forearm. The normal sonographic image of a
tendon is groups of hyperechoic dots mixed in among a hypoechoic
background. Nerves often have fascicles with a hypoechoic center (Fig
8.1A), which may also be apparent in a long-axis view (Fig 8.1B). In
Figure 8.1B, notice the epineurium of the nerve and the hypoechoic
fascicles within the nerve. The primary distinction is that tendons have
hyperechoic continuous fibrils, whereas nerves have fascicles. These
fascicles have a hyperechoic border caused by the perineurium and
hypoechoic center caused by the axons. Tendons may also show greater
anisotropy than nerves, showing a greater change in echogenicity when
toggling the probe. Moreover, a tendon should be seen taking origin
from a muscle or inserting on a bone. This can be seen by scanning
along the length of the structure in question. Finally, a tendon should
move in the image when the appropriate muscle/joint is contracted (Fig.
8.1C–E).
Vasculature
Larger blood vessels are not usually mistaken for nerves. In fact, most
regional anesthesiologists use vascular landmarks to help identify many
nerves. Smaller blood vessels are easier to confuse with nerves,
especially in the short-axis view. Color-flow Doppler is invaluable in
helping to distinguish small blood vessels from nerves. This problem is
frequently encountered while performing brachial plexus nerve blocks
(Figs. 8.2 and 8.3). If in doubt, use Doppler to ensure that the nerve in
question is truly a nerve.
Lymph Nodes
Lymph nodes may be flat or oval and either hypoechoic or hyperechoic.
Lymph nodes can be differentiated from nerves by their lack of fascicles.
They are most commonly seen in axillary, femoral, sciatic, and
interscalene blocks (Fig. 8.4).
Muscle
Muscles have either fusiform or pennant structures. Fusiform muscles
have fascicles with a parallel orientation that converge toward the tendon
at the end of the muscle. Pennate muscles have fascicles that appear
more featherlike in relation to their aponeurosis or tendon. Muscles have
a relatively hypoechoic appearance with some hyperechoic structures
(connective tissue or fat) embedded in the muscle. Some practitioners
liken this appearance to a “starry night.” The ratio of hypoechoic to
hyperechoic muscle is based on the ratio of connective tissue to fascicle
within each muscle. In the long axis, muscle may have a featherlike
appearance (pennate muscle), and it is usually well organized.
Intramuscular tendons and aponueroses are best imaged in the short axis.
It may be difficult to distinguish muscle from nerve when performing
gluteal or popliteal sciatic nerve blockade (Fig 8.5). Small muscles may
have the appearance of a small vessel or nerve. Use of nerve stimulator
can usually allow the practitioner to differentiate nerves from muscles.
Anatomic variation
There have been numerous examples of anatomic variations that have
been described while performing ultrasound-guided regional anesthesia.
In one ultrasound study, nearly half of patients scanned had a variation
of the expected interscalene anatomy. Awareness of the prevalence is
important, especially for the newcomer to regional anesthesia. It is
important to systematically scan patients and remember the anatomic
relationship of muscles, bones, vessels, and nerves.
Anatomic variation by region
Interscalene brachial plexus

1. Intramuscular plexus: Where a part or the entire brachial plexus is


located within one of the scalene muscles, as opposed to between
them. This has been described for both the SC (supraclavicular) and
IS (interscalene) blocks (Fig. 8.6).
Supraclavicular brachial plexus

1. The rainbow plexus: The brachial plexus is arranged in a rainbow


shape around the subclavian artery. This is seen while performing SC
blocks (Fig. 8.7).

2. A divided plexus: This is commonly seen while performing the SC


block. In our experience, the most common cause is the transverse
cervical artery (Fig. 8.3). Another possible cause is the presence of a
cervical first rib.
Axillary brachial plexus

1. Nerves or blood vessels in an unexpected distribution: This is


commonly seen while performing the IS and axillary nerve block.
There have been countless anatomical variations described for the
location of the nerves while performing axillary nerve blocks (Fig.
8.8).
Femoral nerve

1. Femoral nerve lying outside the correct fascial plane


Popliteal sciatic nerve block

1. Inability to visualize the common peroneal nerve


2. Common peroneal and posterior tibial nerves joining together to form
the sciatic on top of each other as opposed to side by side (Fig. 8.9)
Sonopathology of anatomic structures
There are numerous pathologic findings that may be seen while
performing ultrasound-guided regional nerve blocks. These include
abnormalities of the viscera, the vasculature, and nerves.
Vascular Sonopathology
The most common abnormalities seen are atherosclerotic plaques due to
the prevalence of this disease in the U.S. population. Atherosclerotic
plaques are usually calcified and are hyperechoic. A dropout shadow
may be observed deep to the plaque. Plaques with a high lipid content
may be echolucent (Figs. 8.10 and 8.11). The thickness of the intima and
media can be used to quantify the degree of disease as well as the
associated risk of stroke or myocardial infarction. Plaques are most
commonly seen in the carotid artery during interscalene blocks or the
femoral artery during a femoral nerve block.
Another pathologic finding is a deep venous thrombosis (DVT). Fresh
DVTs have variable echogenicity, are usually noncompressible, and
appear as a dilated homogeneous circle (Fig. 8.12). Chronic DVTs may
shrink and have a hyperechoic and heterogeneous appearance (Fig.
8.13A, B). Color-flow Doppler will confirm lack of blood flow.
Arterial Dissections, Aneurysms, or Pseudoaneurysms
A dissection is caused by an intimal tear leading to a false chamber or
passage in the wall of the artery. In this setting, color-flow Doppler
usually shows a to-and-fro flow between the arterial lumen and the
passage in the wall of the artery. This is most commonly seen in the
aorta. In smaller arteries, the etiology is usually iatrogenic due to
previous arterial cannulation.
An aneurysm is a ballooning of an artery due to structural weakening of
the artery. The ultrasound image will show an irregularly shaped,
fusiform dilation of the artery (Fig. 8.14).
A pseudoaneurysm is due to an injury to an artery leading to a hematoma
in congruity with an artery. Doppler exam shows to-and-fro blood flow
between the hematoma and artery (Fig. 8.15A).
Nerves
Ultrasound is currently being used as a tool to diagnose neuropathology,
including neural tumors (Fig. 8.16A–C), traumatic disruptions, neuritis,
and neuropathies (Fig. 8.17B). Identification of neural pathology is
difficult and usually requires scanning the nerve in long-and short-axis
views. Enlarged fascicles and loss of the hyperechoic epineurium may
signal acute neuropathy. Patients with chronic neuropathy may have
enlarged nerves with thickening of the epineurium. A contralateral scan
for comparison is necessary. When new neuropathology is seen, consider
abandoning the regional technique and sending the patient for a more
formal examination.
Thyroid
The thyroid is imaged during stellate ganglion blocks and frequently
during interscalene block. The normal thyroid is usually hypoechoic and
has a homogenous ground glass appearance. Anechoic circles are
usually blood vessels but may represent thyroid pathology. Doppler
imaging will help define these. “If you see something, say something,”
and refer the patient for more definitive testing (Figs. 8.18 and 8.19).
Suggested Readings
Bianchi S. Ultrasound of the peripheral nerves. Joint Bone Spine. 2008;75:643–649.
Bigeleisen PE. The bifid axillary artery. J Clin Anesth. 2004;16:224–225.
Martinoli C, Bianchi S, Derchi LE. Tendon and nerve sonography. Radiol Clin North Am.
1999;37:691.
Silvestri E, Martinoli C, Derchi LE, et al. Echotexture of peripheral nerves: correlation between
US and histologic findings and criteria to differentiate tendons. Radiology. 1995;197:291–296.
Sites BD, Brull R, Chan V, et al. Artifacts and pitfall errors associated with ultrasound guided
regional anesthesia. Part II a pictorial approach to understanding and avoidance. Reg Anesth
Pain Med. 2007; 32:419–433.
Sites BD, Macfarlane AJR, Sites VS, et al. Clinical sonopathology for the regional
anesthesiologist. Part I: vascular and neural. Reg Anesth Pain Med. 2010;35:272–280.
Sites BD, Macfarlane AJR, Sites VS, et al. Clinical sonopathology for the regional
anesthesiologist. Part II: bone, viscera, subcutaneous tissue, and foreign bodies. Reg Anesth
Pain Med. 2010;35:281–289.
Stuart RM, Koh ES, Breidahl WH. Sonography of peripheral nerve pathology. Am J Roentgenol.
2004;182:123–129.
Sutin KM, Schneider C, Sandhu NS, et al. Deep venous thrombosis revealed during ultrasound
guided femoral nerve block. Br J Anaesth. 2005;94:247–248.
Zamorani MP, Valle M. Muscle and tendon. In: Bianchi S, Martinoli C, eds. Ultrasound of the
Musculoskeletal System. New York, NY: Springer; 2007:45–96.
9
Ultrasound Simulator
Training
PAUL E. BIGELEISEN, ERIC STAVNITSKY, AND

KARL REILING

Developing competence in ultrasonography and applying the results to


clinical care is a complex process. It requires psychomotor and optimal
image window acquisition skills. Once an optimal image window is
acquired and correctly interpreted, the information needs to be correctly
applied to patient care. The opportunity cost of training health care
providers on ultrasonography is extremely high. Optimal training
requires (1) a qualified instructor; (2) trainees; (3) an ultrasound
machine; and (4) patients with a variety of anatomic norms, variations,
and abnormalities. All of these elements must come together in one
physical space and time, and the process must repeat with a new patient
presenting these normal and variable conditions over an extended period
of time. It may take months to years before a care provider is able to
scan a sufficient number of patients with certain conditions to develop
competence because some clinical presentations are rarely encountered.
The inability to train on sufficient numbers of variant cases is a
recognized impediment to ultrasound competency.
Many of the currently used training methods have significant limitations.
Traditionally, these consisted of clinical bedside teaching and attending
hands-on training courses. Simple phantoms models were developed
quickly along with didactic videos available on the Internet and from
commercial vendors. These were followed by interactive virtual reality
simulators and then high-fidelity ultrasound simulator workstations.
Each of these training devices has a different cost-benefit ratio.
Examples of simple phantoms, virtual reality simulators, and high-
fidelity simulators are described in the following sections.
High-fidelity ultrasound simulators
High-fidelity simulators require relatively expensive large modules that
may require the user to visit a simulation center. Most devices with good
haptic feedback cost between $30,000 and $100,000. Specific training
simulators involve bulky training platforms that require trainees to visit
a simulation. These ultrasound-training solutions employ dedicated
computer hardware, software, and mannequins that do not deploy over
the Internet. An example of such a simulator is described in the
following text.
Method
Volumetric Description of Virtual Patient
The method starts with a volumetric description of the virtual patient.
Popular sources include computed tomography (CT), magnetic
resonance imaging (MRI), or even computer-aided design. Our data
come from the Visible Human Data Set.
In November 1994, the Center for Human Simulation (CHS) at the
University of Colorado Medical School delivered the Visible Human
Male (VHM) to the National Library of Medicine (NLM) as part of a
contract under the Visible Human Project.1 In November 1995, it
delivered the Visible Human Female. The Visible Humans represent
complete submillimeter visual anatomic descriptions of a male and
female. The voxel (volume element) resolution of the male is ⅓ mm in x
and y (orthogonal coordinates in the axial plane) and 1 mm in the axial
direction. The female has the same x and y resolution with three times
the axial resolution, giving her ⅓-mm voxels. The full-color voxel
anatomy data are supplemented with full-body CT and MRI images. The
CHS has since cut multiple specimens having 1/10-mm resolution in all
three directions. Figure 9.1 is a zoomed-in image taken from a foot and
ankle that were cryosectioned at 1/10 mm.
Segmentation and Classification
Touch of Life Technologies (ToLTech), working in conjunction with the
CHS, has been developing fundamental tools that are necessary for both
visual and haptic interaction with the Visible Human and similar
datasets. These fundamental tools provide the foundation for robust
anatomically accurate simulators. A major part of this effort deals with
segmenting and classifying the data.
Together, ToLTech and the CHS have developed multiple strategies for
assigning to each voxel a number that can be related to a specific
structure. We refer to this three-dimensional augmentation as the alpha
channel. These methods fundamentally produce borders of structures,
either through hand drawing, automatic edge detection, or surface
splines. The tissue within the border is then assigned an identifier.
Figure 9.2 shows a surface spline being used to outline a kidney in a
single image of the VHM.

ToLTech and the CHS have worked together to segment and classify the
entire VHM. Although the effort to refine the data is ongoing, most, if
not all, structures that have a three-dimensional extent of over a
millimeter are now in the database. The alpha channel has become the
foundation for ToLTech’s three-dimensional display, both haptic and
graphic, of individual anatomic structures.
Simplified Physics
In the early part of 2000, ToLTech developed algorithms to create
simulated ultrasound from the VHM alpha channel. The basic idea was
to send rays out from the simulated probe through virtual tissue and
determine the expected energy return. The method assumes that the
information contained in a clinical ultrasound can be closely
approximated by superposition of the following:

1. Impedance mismatch between structures


2. Attenuation of signal based on the normal of the interface compared
with the position of the transmitter/receiver
3. Energy loss proportional to material attenuation
4. Statistically described angle-independent texture associated with
anatomic structures.
Christian Lee was instrumental in all aspects of this development, from
distilling the primary acoustic properties through writing the associated
software.
ToLTech currently associates impedance, sound velocity, and attenuation
with the following tissue types: blood, muscle, air, connective tissue,
lung, water, fat, bone, brain, and nerve.
This abbreviated list has served well for ToLTech’s initial prototypes and
can be continually refined. Creating virtual ultrasound from the VHM
data brings the following attributes:
• Currently, the entire VHM is available with high enough resolution to
simulate today’s ultrasound. This allows the user to seamlessly
interrogate all areas of interest.
• Having a segmented and classified foundation for the simulation
allows us to provide interactive interrogation of the tissues displayed
in the simulated ultrasound (the user can place a cursor on the
simulated ultrasound and have the structure identify itself). This is
important for feedback-based mentoring and testing.
• Deformations of the data due to the probe or anatomic motion can be
immediately rendered in ultrasound.
• The simulated ultrasound can be combined with the virtual anatomy.
Allowing for the practice of ultrasound-guided needle insertion.
Modifying Posture
ToLTech and the CHS have collaborated to develop off-line techniques
for altering the posture of the virtual patient. These techniques utilize
finite element modeling (FEM) to deform tissues. With this technique,
tetrahedra(*) are used to volumetrically describe the virtual patient. The
tetrahedra individually obey laws of deformation and interpenetration.
The aggregate then describes the overall deformation of the virtual
patient. This creates a set of transforms, one for each tetrahedra from the
original space to the deformed space. One common use for the
transforms takes the vertices associated with the polygonal surfaces of
the virtual patient and creates the vertices for the altered posture. This
can be done either statically, creating one posture for the simulation, or
loaded into splines used to dynamically alter the posture in real time.
Figures 9.3 and 9.4 show the VHM in postures altered to correspond to
laterocollis and anterocollis. The images were taken from a simulator
used to train neuromuscular injection techniques. Its development was
funded by Allergan, Inc.
To simulate ultrasound relative to the altered posture, we run the
simulated acoustic rays through the altered volumetric description and
use the transforms to define their paths in the original dataset, where the
acoustic properties are defined for the individual voxels.
Combining with a virtual patient
We originated the simulated ultrasound to give gastroenterologists the
ability to compare “ultrasound anatomy” with three-dimensional models
and oblique views through the VHM
(http://www.vhjoe.org/index.php/vhjoea). Figure 9.5 shows the
combination of an interactive method for determining the relation of the
ultrasound probe and the anatomy, which we refer to as the interactive
atlas, with the resulting simulated ultrasound. The green pie shape in the
far left of Figure 9.5A and green outline in the right image of Figure
9.5A indicate the placement and orientation of the probe in three-
dimensional space and the oblique slice through the VHM data,
respectively.

As our virtual patient evolved, the ultrasound probe became a standard


tool that the trainee can choose.
Simulated Ultrasound for Needle Guidance
To create a virtual patient on which to practice placement of needles,
ToLTech combines stereographic display with collocated force feedback
from haptic devices, giving the ability for a trainee to both see and feel a
virtual patient.
Real-time display of complex three-dimensional scenes became practical
with the development of texture-mapped polygonal rendering. Spurred
by computer gaming, the computer graphics industry has responded with
inexpensive graphics cards capable of rendering millions of texture-
mapped polygonal models per second. By rendering the scene separately
from the perspective of each eye, stereo pairs can be produced, adding
depth to the scene.
Haptic devices went from laboratory to commercial application when
SensAble Technologies introduced the PHANToM Premium in the early
1990s (http://sensable.com/). Haptic devices may be thought of as robots
whose paradigm is not one of moving joints to achieve a position of the
end effector but instead to produce torques that are translated to a
controllable force at the end effector. If the user grasps a tool that is
connected to the end effector, the haptic device can produce the same
force, through the handle of the tool, that the user would feel should he
or she be interacting with a real object.
In general, haptic devices cannot be placed directly in the source of the
stereographic display. To give the appearance of collocating the haptics
and graphics, ToLTech uses a mirror system to project the apparent
image of the stereo display into workspace of the haptic devices. The
resulting system, which ToLTech refers to as the Common Platform
Medical Skills Trainer (CPMST), produces a highly adaptable
environment in which to practice bimanual skills. Figure 9.6 shows a
trainee using the CPMST to practice fasciotomy. In this image, the top
screen shows a two-dimensional version of what the trainee sees in
stereo; the middle screen displays the Mentor, which guides, critiques,
and scores the progress of the trainee; and the screens to either side of
the trainee’s head combine through a set of mirrors to collocate the
stereo view of the patient with the haptic devices that the trainee holds.

Results for training ultrasound-guided needle placement


The following images show some of the steps that a trainee goes through
using the CPMST to practice a femoral block.
In Figure 9.7, the Mentor has selected a palpating tool and the trainee
uses it to locate the femoral pulse along the inguinal crease.
When the trainee has successfully found the pulse, the Mentor
substitutes a marking pen for the palpator tool (Fig. 9.8) and requests
that the trainee mark the site of the pulse as well as a spot 1.5 cm lateral
to the pulse. This becomes the skin puncture site for either a stimulating
needle or a nonstimulating needle when accompanied by ultrasound.
In Figure 9.9, the Mentor has placed a transparent ultrasound probe at
an optimal site in the virtual scene. The resulting ultrasound for the
chosen site is shown as the ultrasound image to the right. The trainee is
controlling the opaque ultrasound probe and sees the resulting
ultrasound image (the one to the left) update in real time.
As the trainee maneuvers the probe to match the static probe, the
transparent probe disappears and the two simulated ultrasound images
become similar. The trainee is free to translate and rotate the virtual
probe to create comparisons with the static image. In Figure 9.10, the
trainee has chosen a suitable position and orientation for the ultrasound
probe and used it to guide the needle to the femoral nerve.
Figure 9.11 shows a transparent view of the scene, confirming the
success of the trainee. When scoring the ability of the trainee, the
Mentor is fairly lax about the placement of the ultrasound probe.
However, its measures for the needle are transformed into the space of
the ultrasound probe. This allows the authors to easily describe tight
parameters for success based on how well the needle is showing in the
ultrasound image.
The development of the femoral block simulator was greatly aided
through guidance from Drs. Adrian Hendrickse and John Armstrong,
both from the Department of Anesthesia at the Anschutz Medical
Campus of the University of Colorado.
Near-term research
The current simulated ultrasound responds, in real time, to the changes
in the position and orientation of the virtual probe. In addition, it can
respond to dynamic posture changes as described earlier. The ability to
transform rays from the original posture to the deformed posture lays the
foundation for the ultrasound to display the results of any movement
occurring in the deformed space. Thus, all modifications that can be
made to the virtual patient using finite modeling techniques could be
displayed in the simulated ultrasound. It does not take a great deal of
computational power to modify the vertices of the tetrahedra describing
arteries in such a way as to create a pulsing vessel. Thus, this technique
could easily bring the ability to not only feel pulsing arteries but to see
them pulse in the ultrasound as well.
Real-time deformation of tissue in response to collision with tools is
more computationally demanding. The computational demand is
exacerbated by nonuniformity of stiffness of structures. We have FEM-
like methods that can convert collisions of a tool with exposed surfaces
and transform them into reasonable deformations throughout the virtual
body. However, unless constrained to very small volumes, they generally
cannot be run on commercially available computers. Fortunately, both
the first principles ultrasound simulation and the deformation of the
volumetrically described tissue are amenable to parallelization. With that
comes the ability to take advantage of the hundreds of graphical
processor units available on a standard graphics card. We expect this to
give us the ability to include deformation of tissues as a result of both
the ultrasound probe and the needle.
Interactive virtual reality simulators
Interactive virtual reality simulators usually employ a portable computer
and some type of joystick or artificial ultrasound probe that allows the
user to perform an ultrasound exam on a virtual human. Costs begin at
about $5,000. Most of these simulators do not allow the user to perform
realistic procedures on the virtual phantom, but they can provide good
image acquisition skills. The user will have to receive other types of
training to learn to image the needle and place it next to the nerve.
The SonoSim Ultrasound Training Solution combines the features of a
fully interactive ultrasound simulator with the scalability of online
learning and assessment (Figs. 9.12 and 9.13). Its core technology for
capturing scanned images with corresponding transducer movements in
three-dimensional space during original data capture and embedding the
acquired data sets into a laptop-based training interface is unique. It also
contains a cognitive task-training feature that enables ultrasound-guided
needle-based procedure training (e.g., regional nerve blocks).
Physical mannequins
Many mannequins have been developed that allow the user to obtain
some familiarity with ultrasound-guided nerve block. Blue Phantom and
MiniSim phantoms are commercially available at prices ranging from
$700 to $10,000. These phantoms provide the user an introduction to
standard nerve blocks. The simulators require the user to have access to
a functioning ultrasound platform, which adds to the cost of use if such a
platform is not already available to the trainee. These simulators are
most useful for early novices. The ultrasound images that can be
generated are not very realistic, but they do teach the relative locations
of bones, vessels, and nerves. Elementary probe manipulation and
scanning can be mastered as well as needle imaging and guidance. An
example of a femoral nerve mannequin is shown in Figure 9.14. The
yellow strip indicates the approximate location of the femoral nerve, and
medial to this lie the artery and vein, respectively. The black screen
signals the user with a flashing light and a tone when the user touches
the virtual nerve with an electrostimulation needle. The two ports at the
caudal end of the simulator allow the user to attach a syringe to the
femoral artery and vein. A separate user can cause these vessels to pulse
while the trainee scans the nerves and vessels. An example of the
ultrasound image is shown in Figure 9.15.
* Tetrahedra are three-dimensional volumetric constructions using triangle.
Reference
1. Spitzer VM, Ackerman MJ, Scherzinger AL, et al. The visible
human male: a technical report. J Am Med Inform Assoc.
1996;3:118–130.
10
Advanced
Ultrasound-Guided
Needle Technology
STEVEN L. OREBAUGH, ARVIND MURTHY,
KRISITIN ODENKO LIGDA, RICHARD ZHU,
PAUL E. BIGELEISEN, AND IWAN SOFJAN

Ultrasound-guided techniques have become the standard of practice in


regional anesthesia. Multiple studies have shown that using ultrasound
improves both the practicality and the efficacy of performing most types
of nerve blocks. One important feature that the ultrasound provides is
needle tip visibility, which may lead to fewer complications and a more
accurate anesthetic application than the traditional nerve stimulation
technique. As such, various innovations have been developed in the
hopes of maximizing needle visibility. These technologies, which will be
discussed in the following text, include echogenic needles, needle
guides, guidance positioning system (GPS)–guided needles, and digital-
enhancement software.
Echogenic needle General Concepts The visibility of an object imaged with
ultrasound depends on how well it reflects the incoming ultrasound wave back to
the probe sensors. Snell’s law requires that the angle of incidence (insonation)
equals the angle of reflection (backscattering) for a flat, hard surface. Thus,
maximal reflection occurs when the direction of wave is perpendicular to the
reflecting surface (Fig. 10.1A). By convention, this angle of insonation is
defined as zero. With linear probes, the needle shaft must be parallel to the
surface of the probe in order to get a strong reflection from the surface of a
smooth needle. In clinical practice, this does not occur because the angle of
insonation usually varies from 20 to 70 degrees when ultrasound-guided nerve
block is performed. In layman’s terms, when the probe is flat on the skin, the
needle usually makes an oblique angle with the surface of the probe (Fig.
10.1B). When this occurs, the sound wave is not reflected back to the probe.
In contrast, an echogenic needle has an engraved pattern in the needle
shaft that maximizes ultrasound wave reflection back toward the probe.
Two types of echogenic etchings are in common use. The first type is a
dihedral groove etched into the needle surface (Fig. 10.2A). This type of
etching works well if the incoming wave is a plane wave aligned with
the shaft of the needle. In most cases, the groove is etched continuously
around the shaft of the needle similar to that of the thread on a screw
(Fig. 10.3). In this manner, the dihedral reflector actually becomes a
universal three-dimensional surface reflector. The depth and period of
the groove can be manufactured to optimize the reflection from sound
waves of different frequencies. The other type of reflector is a
tetrahedral reflector that has a four-sided intermittent etching in the shaft
of the needle (Fig. 10.2B). In this case, the tetrahedral etching is the
universal three-dimensional reflector, and the needle has a triangular
pattern etched in its surface (Fig. 10.4).
Examples of Echogenic Needles 1. Life-Tech EchoBright The Life-Tech
EchoBright has a continuous dihedral V-shaped groove etched in its
surface (Fig. 10.3). This design has been shown to produce a constant,
even intensity of reflection when the angle of insonation varies over a
range from 0 to 80 degrees (Fig. 10.4). It is also available in an insulated
model so that it can be used with an electrostimulator.
2. Pajunk Sonoplex The Pajunk Sonoplex echogenic needle has a linear
pattern of tetrahedral reflectors that are spaced out circumferentially
around the shaft (Fig. 10.5). This arrangement allows reflection and
imaging at a wide range of insertion angles and bevel orientation (Fig.
10.6). Similar to a standard insulated block needle, the Sonoplex can be
connected to a nerve stimulator and can be used for a single injection
application or continuous infusion. Several different tip heads are also
available, as shown in Figure 10.5.

Clinical Data Clinical trials comparing echogenic needles against


standard needles are still few and far between. In general, studies show
that echogenic needles have better visibility,1,2 are less affected by the
angle of insertion,1,3 and can be utilized to more rapidly reach the target,
especially for a deeper nerve.4
In a trial comparing the Pajunk Sonoplex (echogenic) versus the Pajunk
Uniplex (standard), Hebard and Hocking1 found that the Sonoplex has a
better subjective and objective needle tip visibility. This study was a
randomized controlled trial performed with 60 peripheral nerve blocks
(half femoral, half sciatic) at four different needle insertion angles (5, 20,
35, and 50 degrees). The visibility of the needle was graded individually
using a 5-point scale by each anesthesiologist (the subjective portion)
and independently by a third person who recorded the percentage of the
time the needle tip was visible in the ultrasound monitor (the objective
portion). By these measures, the Sonoplex outperforms the Uniplex on
both femoral and sciatic approach at all insertion angles. Most
importantly, the Sonoplex maintains its objective visibility at all the
specified angles, whereas the Uniplex loses visibility as the angle
insertion increases.
In a separate study, Guo et al.2 compared the Pajunk Sonoplex, the
Braun Stimuplex, and the Braun standard needle. All needles were used
in-plane and out-of-plane at four different angles (30, 45, 60, and 75
degrees) in a cadaver model. The visibility was assessed by two
independent, blinded observers using a 5-point Likert ordinal scale.
Based on the measures, both the Sonoplex and the Stimuplex outperform
the standard needle in-plane and out-of-plane. Both echogenic needles
perform similarly in-plane, whereas the Sonoplex is more visible out-of-
plane.
In another study by Hebard et al.,3 the authors compared the visibility of
three echogenic needles (Pajunk Sonoplex, Life-Tech, Braun Stimuplex
D+) and a standard needle at different depths and insertion angles. The
ultrasound images were subjectively evaluated by 20 anesthesiologists in
terms of the needle tip location estimate, confidence in this estimate, and
the overall needle visibility. The actual tip position was blinded from the
evaluators and the tip estimate error was objectively measured. At
shallow angles, the echogenic needles did not significantly outperform
the standard needle in terms of confidence and needle visibility. At
steeper angles, however, the echogenic needle, especially the Sonoplex,
had significantly higher confidence and needle visibility. This pattern
was also found to be consistent in the objective tip error measurements.
These results suggest that, at steeper angles, the echogenic needles tend
to have less tip error. Of note, when the tip error is significant, the tip
tends to be deemed more superficial than it actually is, which can have
safety implications.
Some authors have also assessed whether using the echogenic needle
improves the speed of needle tip placement. Saurabh4 compared the
Braun Stimuplex, the Pajunk Sonoplex, and a standard needle in terms
of “needle-to-nerve time” to reach two phantom nerves at 10-and 20-mm
depth. Clinician experience level was also evaluated as a secondary
parameter. There was no reported difference in rate of needle placement
between the echogenic Stimuplex or Sonoplex as compared to the
standard needle at the shallow (10 mm) nerve depth. The level of
clinician experience, surprisingly, also did not affect the speed
significantly. For the deep (20 mm) nerve target, however, both
echogenic needles allowed faster placement than the standard needle,
regardless of the clinician’s experience level.
Even though most of these trials seem to support the use of echogenic
needles, there are still limitations that should be considered. For
example, most of the studies were done in nonliving models, such as
cadavers or phantoms, which may not translate directly to its use in a
living human. Furthermore, the sample sizes of these studies are small,
and none of the studies measured important clinical outcomes such as
the rate of successful blocks, the difference in complication rate, or the
cost efficiency. Nevertheless, echogenic needles appear to be a step
forward in regional anesthesia, and some clinical trials are currently in
progress to better assess these outcomes. One such trial is a randomized
clinical comparison evaluating the use of the Pajunk Sonoplex against a
standard Pajunk nonechogenic needle in sciatic nerve blocks for knee
replacements. As more of these trials are performed, we come to
understand more about the actual clinical efficacy of the echogenic
needle.
Needle guides General Concepts Needle guides come in three types.
1. Mechanical Needle Guide The first type is a mechanical guide that
forces the needle to remain aligned with the probe either using an in-
plane approach or an approach in which the needle is inserted
transverse to the long axis of the probe. For nerve block, the in-plane
guide is most useful. An example of the in-plane guide is shown in
Figure 10.7A and B. In Figure 10.7C, the reader can see that the guide
keeps the needle in line with the probe even when the user cannot see the
exact insertion point of the needle relative to the probe. Civco makes
these guides for many probes. Their cost is about $400 for the reusable
red bracket (Fig. 10.7A, B) that attaches to the probe and $15.00 for the
green disposable guide (Fig. 10.7A, B).
These guides can also be designed with simple CAD software and
manufactured in a fully disposable form using a three-dimensional
printer for about $4.00 for both the red bracket and green guide. An
example of a disposable guide manufactured with a three-dimensional
printer is shown in Figure 10.7C.
2. Combined Electromagnetic Needle Guides The second type of needle
guide tracks the needle in virtual space using a mechanical needle guide
combined with an electromagnetic guidance system. These needle guides
produce a virtual image of the needle on the ultrasound image, which
allows very accurate placement of the needle tip relative to the nerve.
Several mechanisms have been invented to achieve this, including two
types of these needle guides available in the market today: the Soma
AxoTrack and Civco eTrax.

Soma AxoTorack The Soma AxoTrack utilizes magnetic coupling to


detect the needle position. The system has two main components. The
first is a specially designed ultrasound probe that features a track of
magnetic sensors and a needle guide to hold the needle in place (Fig.
10.8A, B). The second component is a custom needle equipped with a
magnetic ring in its shaft (Fig. 10.8D). This magnetic ring will activate
different magnetic sensors in the probe as the needle moves forward and
backward, thus indirectly providing the location of the tip because the
distance between the magnetic ring and the tip is constant (Fig. 10.8C).
This system has been used for central venous access, cyst aspiration,
arthrocentesis, biopsy, and nerve blocks.
Civco eTrax

The Civco eTrax uses a different mechanism to track the needle tip. It
has an electromagnetic transmitter in the needle tip (Fig. 10.9A) and a
needle bracket. The transmitted signal from the tip is detected by a
sensor in the ultrasound probe and used to generate a real-time virtual
image of the needle on the screen (green line on Fig. 10.9B). The needle
bracket locks the needle in place to ensure proper in-plane alignment of
the needle. In addition to nerve blocks, this system has been used for
ablations, biopsy, drainage, aspiration, therapeutic delivery, and vascular
access.

GPS-guided needle placement The third type of needle guide uses a GPS to
produce a virtual image of the needle. It also uses an electromagnetic system to
produce the virtual needle image superimposed on the ultrasound image. It
consists of transmitting units and a receiver but does not require a mechanical
guide to keep the needle inline with the probe. For this reason, the user has much
more freedom in choosing how he or she wishes to insert the needle relative to
the probe. This type of unit usually costs about $10,000 for the transmitter and
receiver. Specialized needles costing about $50.00 per needle are required.
A basic GPS system is shown in Figure 10.10. By knowing the distance
from each transmitter (A, B, and C), the receiver (green dot) can deduce
its location using a process known as triangulation. Utilizing this
mechanism, one can build a GPS needle locator by embedding a sensor
in the needle and creating a transmitter to emit the signals. One such
system is the UltraSonix GPS, which is composed of a GPS transmitter
unit (the dark gray device in Fig. 10.11A) and GPS sensors on both the
needle (tip of the yellow cable in Fig. 10.11A and B) and the ultrasound
probe. Using both of these sensors, the needle location can be projected
in the screen, along with its position in relation to the probe (Fig. 10.12).
Clinical Data As with echogenic needles, not many studies have been
performed comparing the use of needle guides against standard methods.
One study by Ball et al.5 compared the use of a standard, physical needle
guide in placing central venous catheters against the standard freehand
technique in a specific central venous catheter placement mannequin.
Thirty anesthesiology residents participated in the study, and a needle
guide used in the long-axis approach was compared against the short-
axis freehand and the long-axis freehand techniques. Two objective
outcomes were measured, the fraction of time the needle is visible and
the time until the vessel puncture is made. The needle guide did
significantly improve the needle visualization compared with both
freehand techniques. However, the long-axis needle guide did not speed
up the puncture of the target vessel compared to the long-axis freehand
technique. On the contrary, it was even slower than the short-axis
freehand technique. This is somewhat expected because any long-axis
approach is technically more difficult than the short-axis approach.
Theoretically, these same benefits may apply to the use of this type of
needle guide to the performance of peripheral nerve blocks.
Needle enhancement software In addition to the hardware advancements
discussed so far, Sonosite has created a software algorithm that can digitally
enhance a standard needle image without affecting the appearance of
background structures. This software can be downloaded to some of the newer
Sonosite ultrasound machines, but the detail of the algorithm is proprietary.
Figure 10.13 shows how a standard needle is made more visible by using the
software.
References
1. Hebard S, Hocking G. Echogenic technology can improve needle
visibility during ultrasound-guided regional anesthesia. Reg Anesth
Pain Med. 2011;36:185–189.
2. Guo S, Schwab A, McLeod G, et al. Echogenic regional anaesthesia
needles: a comparison study in thiel cadavers. Ultrasound Med
Biol. 2012;38:702–707.
3. Hebard S, Hocking G, Murray K. Two-dimensional mapping to
assess direction and magnitude of needle tip error in ultrasound-
guided regional anaesthesia. Anaesth Intensive Care.
2011;39:1076–1081.
4. Saurabh RN. “Needle to nerve time” comparison of four different
echogenic ultrasound guided regional anaesthesia nerve block
needles. Reg Anesth Pain Med. 2011;36.
5. Ball, RD, Scouras NE, Orebaugh S, et al. Randomized, prospective,
observational simulation study comparing residents’ needle-guided
vs freehand ultrasound techniques for central venous catheter
access. Br J Anaesth. 2012;108:72–79.
11
From Paresthesia to
Neurostimulation and
Ultrasound-Guided
Regional Anesthesia
SYLVIA WILSON AND JACQUES E. CHELLY

In 1911, Kulenkampf introduced the concept of locating a nerve with a


paresthesia and blocking its conduction with local anesthetic. Paresthesia
ruled as the technique of choice for another 50 years. Although the
concept of neurostimulation was also introduced around 1910, it took
several decades for this technique to be recognized and used. Until
recently, neurostimulation has been the primary technique to perform
peripheral nerve blocks.
In the past 15 years, an increasing number of investigators have
demonstrated the usefulness of ultrasound-guided techniques to perform
regional anesthesia. In adults and children, the studies demonstrating the
usefulness of ultrasound-guided techniques were initially conducted in
patients undergoing upper extremity and superficial peripheral blocks.
The use of ultrasound in the pediatric population has gained rapid
recognition because of the quality of the images that can be generated in
this patient population. The use of ultrasound was eventually extended to
deeper peripheral blocks, especially sciatic and lumbar plexus blocks
when more powerful portable sonography platforms were developed.
The basis defining the use of ultrasound for the performance of neuraxial
blocks and for blocks traditionally used for the treatment of chronic pain
was also developed in this time frame.
In the past few years, studies have attempted to compare
neurostimulation to ultrasound-guided techniques. A systematic review
in 2010 examined block onset time, quality, and duration in 16
randomized controlled trials (RCTs) comparing ultrasound guidance
with nerve stimulation. Although one RCT found a slower onset time (2
minutes) with ultrasound guidance in upper extremity blocks, all of the
other RCTs reviewed found that ultrasound guidance was not different or
improved in terms of block onset time (60%), quality (25%), and
duration (17%). Although duration was not prolonged in lower extremity
blocks, RCTs reviewed showed ultrasound guidance was equivalent or
superior for block onset time (71%) and quality (63%). Multiple studies
have also demonstrated that ultrasound guidance decreases time to
perform peripheral nerve blocks. This results in fewer needle
manipulations and improves patient comfort and satisfaction with
regional anesthesia.
Utilization of ultrasound was hoped to decrease rare adverse events
associated with regional anesthesia. Whether or not ultrasound can
reduce the frequency of nerve injury associated with the performance of
regional anesthesia is an important but very complex question. Nerve
injury following regional anesthesia and surgery is very uncommon, and
the etiology is multifactorial. Many of the injuries occurring after
surgery have been shown to be related to the surgery (e.g., compression
of the tissues with a tourniquet, direct surgical injury of the nerve as well
as casting and positioning of the patient) rather than the performance of
a regional technique. Even when the nerve block is suspected as the
causative agent of injury, toxicity of the local anesthetic to the nerve
may be the etiology rather than the needle or injection per se. A
retrospective review of 5,436 patients described eight adverse events
(three nerve injuries and five seizures) associate with nerve-stimulated
guided blocks compared to no adverse events with ultrasound guidance.
It is not surprising that the incidence of seizure was decreased in the
ultrasound groups since ultrasound guidance has been repeatedly shown
to decrease the incidence of vascular puncture in upper and lower
extremity nerve blocks. Marhofer et al. and Mariano et al. examined
femoral nerve block placement and found a significant incidence of 15%
(3 of 20) and 20% (4 of 20), respectively, without ultrasound guidance
versus 0% with ultrasound guidance. Similarly, Mariano et al. reported
that 30% (6 of 20 patients) receiving infraclavicular blocks without
ultrasound experienced a vascular puncture compared to zero with
ultrasound guidance. Sauter et al. reported 33% of unintended blood
aspiration with neurostimulation versus 5% with the use of ultrasound
while performing infraclavicular blocks.
Studies have also compared the use of neurostimulation alone to the
combination of ultrasound and neurostimulation. Because the limits of
ultrasound imaging may be exceeded in some patients, a combined
approach can be useful in the case of deep peripheral blocks such as
proximal sciatic nerve blocks or lumbar plexus blocks. Another setting
where combining ultrasound with stimulation appears effective is when
the plexus is splayed out over a large area such as in the axilla. In these
cases, the combination of neurostimulation and ultrasound has been
proven to be more reliable than the use of the neurostimulation alone.
This is an exciting time for the use of ultrasound. The technology and
the equipment are constantly improving while declining in price and
improving access. Ultrasound training is widely available in most
residencies, and new techniques using ultrasound guidance are regularly
described in the literature. With ultrasound guidance, prohibitions
against the use of peripheral nerve block in anticoagulated patients have
decreased. Similarly, regional anesthesia utilization has greatly increased
in pediatric populations because of the ability to visualize needle
placement and decrease the local anesthetic dose injected. Nonetheless,
the experience and skill of the practitioner are still paramount with this
technique and will continue to be so for the near future. Advances in
multiplanar sonography (three-dimensional imaging), more echogenic
needles, better needle guides, and machine identification of nerves will
undoubtedly improve the efficiency of ultrasound in the future.
Bibliography
Brull R, Wijayatilake DS, Perlas A, et al. Practice patterns related to block selection, nerve
localization and risk disclosure: a survey of the American Society of Regional Anesthesia and
Pain Medicine. Reg Anesth Pain Med. 2008;33:395–403.
Chan VW, Perlas A, McCartney CJ, et al. Ultrasound guidance improves success rate of axillary
brachial plexus block. Can J Anaesth. 2007;54:176–182.
Kulenkampf D. Anesthesia of the brachial plexus [in German]. ZentralblChir. 1911;38:1337–
1350.
Liu SS, Ngeow J, John RS. Evidence basis for ultrasound-guided block characteristics: onset,
quality, and duration. Reg Anesth Pain Med. 2010;35:S26–S35.
Marhofer P, Schrogendorfer K, Koinig H, et al. Ultrasonographic guidance improves sensory
block and onset time of three-in-one blocks. Anesth Analg. 1997;85:854–857.
Marhofer P, Schrogendorfer K, Wallner T, et al. Ultrasonographic guidance reduces the amount of
local anesthetic for 3-in-1 blocks. Reg Anesth Pain Med. 1998;23:584–588.
Marhofer P, Sitzwohl C, Greher M, et al. Ultrasound guidance for infraclavicular brachial plexus
anaesthesia in children. Anaesthesia. 2004;59:642–646.
Mariano ER, Loland VJ, Sanhu NS, et al. Ultrasound guidance versus electrical stimulation for
femoral perineural catheter insertion. J Ultrasound Med. 2009;25:1453–1460.
Mariano ER, Loland VJ, Bellars RH, et al. Ultrasound guidance versus electrical stimulation for
infraclavicular brachial plexys perineural catheter insertion. J Ultrasound Med. 2009;28:1211–
1218.
Neal JM, Brull R, Chan VW, et al. The ASRA evidence-based assessment of ultrasound-guided
regional anesthesia and pain medicine: executive summary. Reg Anesth Pain Med.
2010;35:S1–S9.
Orebaugh SL, Williams BA, Kentor ML. Ultrasound guidance with nerve stimulation reduces the
time necessary for resident peripheral nerve blockade. Reg Anesth Pain Med. 2007;32:448–
454.
Orebaugh SL, Williams BA, Vallejo M, et al. Adverse outcomes associated with stimulator-based
peripheral nerve blocks with versus without ultrasound visualization. Reg Anesth Pain Med.
2009;34:251–255.
Sauter AR, Dodgson MS, Stubhaug A, et al. Electrical nerve stimulation or ultrasound guidance
for lateral sagittal infraclavicular blocks: a randomized, controlled, observer-blinded,
comparative study. Anesth Analg. 2008;106:1910–1915.
Sia S, Bartoli M. Selective ulnar nerve localization is not essential for axillary brachial plexus
block using a multiple nerve stimulation technique. Reg Anesth Pain Med. 2001;26:12–16.
12
Understanding
Needle-to-Nerve
Proximity in
Peripheral Nerve
Blocks
ALAN J. R. MACFARLANE AND RICHARD

BRULL

The most important goal of regional anesthesia is arguably to accurately


locate the nerve(s) and ensure a good quality block without causing
complications. Nerve damage is a potentially devastating but fortunately
rare complication of peripheral nerve blocks.1 The pathogenesis of such
nerve injury is complex and may be multifactorial. Suggested etiologies
include intraneural injection, direct mechanical injury, local anesthetic
(or additive) neurotoxicity, ischemia, pressure-related effects of the local
anesthetic, and edema/hematoma formation.2,3 Of these, both intraneural
injection and direct mechanical injury have long been recognized as
significant causes of nerve injury. The first—namely, intraneural
injection—appears to occur more commonly than previously thought,
but fortunately, this does not necessarily result in overt nerve injury.4–6
This is because evidence is now mounting that it is intrafascicular
intraneural injections in particular that are most harmful and a major risk
factor for nerve injury.7,8 The pathogenesis of the second cause, direct
mechanical injury, is similarly complex. Although a needle placed
deliberately intraneurally does not necessarily cause fascicular damage,2
even forced needle-to-nerve contact without penetration of the
epineurium increases the risk of histologic injury compared to resting
the needle against the nerve without any pressure.9 It is clear therefore
that in order to minimize nerve injury due to either intrafascicular
injection or direct traumatic mechanisms, there must be precise
knowledge of the needle tip position relative to the nerve during the
conduct of peripheral nerve blockade. Unfortunately, our ability to
monitor needle tip position with complete accuracy and certainty in
clinical practice is limited. Even the much-heralded advent of
ultrasound, which allows direct needle and nerve visualization, has been
unable to prevent intraneural injection.10,11 The purpose of this chapter
is to describe current clinically available nerve localization devices and
how these are able (or not) to detect needle-to-nerve proximity,
intraneural injection, and intrafascicular injection. The devices
specifically examined are nerve stimulators,12–19 injection pressure
monitors,18,20–23 and ultrasound.18,23 As much of the research is animal-
based, a short review is firstly provided on the major limitations of such
work.
In all but one of the peripheral nerve animal experiments described in
this chapter, the researchers chose pigs and dogs to study. The selection
of an animal model depends on many criteria, one of which clearly is
similarity to humans. For this reason, pigs—which have similar biology
and physiology as well as equivalent sizes of soft tissue, nerves,
epineurium, and surrounding vessels—are a common choice.24 In
addition, the risk of permanent nerve injury is reportedly greater in
larger animals, possibly due to reduced regenerative capacity.25 This
therefore further explains why pigs, along with primates, sheep, and
dogs, are commonly utilized in studies examining peripheral nerve
injury.24 Irrespective of any similarities with human nerves, however, it
must always be remembered that the major limitation shared by all
animal studies is that the results may not be directly transferable to
human clinical practice. A variety of other technical factors in the
experiments potentially further complicate the clinical applicability of
these animal studies. Firstly, some experiments were performed on
surgically exposed “open” nerves.13,14,19–23 Although in some cases the
tissue surrounding the nerve remained intact,19 it remains unclear
whether electrical conductivity, injection pressures, or ultrasound
imaging is comparable in these circumstances to the other “closed”
model studies where a percutaneous injection more similar to human
clinical practice was undertaken.12,15–18 Secondly, a number of studies
examined for either histologic signs of inflammation and
injury12,13,18–20,22,23 or utilized either ink or resin deposition as a marker
of needle-to-nerve proximity.15–17 Both of these techniques are subject
to postprocessing tissue distortion, and controls were not always used.
This could increase the risk of both false positives and inaccuracies
regarding the final needle tip location. In addition, the histologic
assessment of nerve injury was not always blinded. Finally, the
significance of any histologic injury was arguably unclear because most
studies did not examine animals for evidence of functional injury.
However, even when functional assessment of the animals was indeed
undertaken, there is a possibility that the clinical assessment tools used
were too insensitive to detect mild or even moderate functional
neurologic deficit, and so again any findings may not be transferable to
human nerve injury. With these limitations in mind, each of the
measurement tools are discussed in the following text. Further specific
limitations of the studies are discussed at the end of each section.
Electrical stimulation
As an alternative to using paresthesia to locate nerves, Greenblatt and
Denson26 introduced the portable nerve stimulator in 1962. This device
utilizes the principle of Coulomb’s law, which dictates that the threshold
current required to elicit a motor response (the minimum stimulating
current [MSC]) should exponentially decrease as the insulated needle tip
is advanced toward the nerve. Although this should in principal indicate
needle-to-nerve proximity, the absolute MSC is influenced by a number
of factors. These include pulse width, current polarity, bore size, and
geometry of the insulated needle tip, constitution of the needle-tissue
interface (e.g., dextrose versus saline), and indeed the specific peripheral
nerve being stimulated.27–29 A further confounding factor is the actual
precision and accuracy of the current delivered by the nerve stimulator.
This has been shown to vary considerably, particularly in the low-current
range, which is most important in nerve stimulation.30 Despite these
variables, convention states that an MSC of greater than 0.5 mA signifies
a lack of needle-to-nerve proximity and therefore is more likely to be
associated with block failure, whereas an MSC of less than 0.2 mA may
increase the risk of intraneural injection.31 However, evidence is now
emerging from recent animal studies that this relationship between
needle tip location, the target nerve, and MSC is not as clear.
A number of animal studies have examined MSC and intraneural needle
placement, with interesting results. By placing the tip of a stimulating
needle intentionally within the nerves of the brachial plexus in pigs,
Chan and colleagues13 demonstrated that the MSC required to achieve a
distal motor response was less than 0.2 mA in only one-third of cases.
Similarly, Tsai and colleagues14 found that the mean MSC required to
elicit a sciatic motor response in pigs with deliberate intraneural needle
tip placement was 0.56 mA, with a number of cases requiring currents
between 0.8 and 1.8 mA. Conversely, the MSC was less than 0.15 mA
only when the needle tip was intraneural. Steinfeldt and colleagues,16,17
experimenting on porcine femoral and brachial plexus nerves, also
elicited that an intraneural injection, as confirmed by resin injection,
could still occur with an MSC up to 1.0 mA. Finally, with a needle
inserted under ultrasound guidance within pig sciatic nerve
(subsequently demonstrated to be extrafascicular), Altermatt and
colleagues18 noted that the MSC needed to be as high as 3.3 mA in order
to elicit a motor response. Of these studies concentrating on deliberate
intraneural injection and MSC, only Chan and colleagues examined for
histologic evidence of nerve injury.13 Interestingly none was present,
although just 2 out of 24 injections in that study were confirmed with ink
to be intrafascicular. In 2005, Voelckel and colleagues12 did specifically
study the relationship between histologic nerve injury and MSC. Unlike
the other studies, the final needle position was not known, but nerve
injury was evident in 50% of the pig sciatic nerves when an MSC of less
than 0.2 mA was used to elicit a motor response, compared to no
histologic changes at an MSC between 0.3 and 0.5 mA.
Recent animal studies have also examined the relationship between
MSC and the distance from the needle tip to the nerve. Rigaud and
colleagues15 demonstrated that MSC levels within the wide range of
0.33 to 1.0 mA resulted in needle placement similarly close to the sciatic
nerve in dogs and that there was no correlation between MSC and
distance from the target nerve as measured by ink injection. Steinfeldt
and coworkers16 similarly found that the needle tip proximity to the
femoral or brachial plexus nerves of anesthetized pigs was equivalent
whether the motor response was elicited at either a low (0.01 to 0.3 mA)
or high (0.8 to 1.0 mA) MSC. This same group of investigators
performed a series of transcutaneous injections of dye at various current
outputs and demonstrated that the final dye position was always “close”
to the nerve when the MSC ranged between 0.3 and 1.4 mA. Above this
cutoff, however, the injectate was infrequently close.17 Conversely, Tsai
et al.14 did find a direct correlation between MSC and needle-to-nerve
proximity at distances of less than 0.1 cm. Also, in a further study by
Steinfeldt’s team19 again examining pig brachial plexus nerves, the final
needle position was closer to the nerve when the nerve was approached
with a constant MSC of 0.2 mA (mean distance 0.3 mm) compared to
1.0 mA (mean 2.9 mm). Interestingly, these distances, which do appear
to represent a correlation between MSC and needle-to-nerve proximity,
were within the range of those in Tsai and colleagues’14 study. Despite
finding a link between needle-to-nerve proximity close to the nerve, Tsai
and colleagues14 in the same study, however, found that with the needle
tip located on the surface of, but not penetrating, the epineurium, the
MSC ranged from 0.15 to 1.4 mA.
Some newer nerve stimulators have the ability to measure electrical
impedance. Electrical impedance is the resistance to flow of an
alternating current in an electrical circuit and it is highly sensitive to
changes in tissue composition, particularly water content. Tsui and
colleagues32 demonstrated that nerves have higher electrical impedance
than the surrounding extraneural tissue and that impedance increased
sharply upon moving from the extraneural to the intraneural
compartment. An absolute value, however, representing intraneural
needle placement could not be determined.
In addition to the generic issues previously described, there are a number
of specific technical and methodologic limitations with these studies.
For example, the pulse duration was not constant among the studies.
This may complicate direct comparison of absolute values because a
longer pulse width can theoretically reduce the MSC.33 Voelckel et al.,12
for example, used a pulse width of 0.3 msec instead of the conventional
0.1 msec, which in theory may have led to an underestimation of the
MSC associated with histologic inflammation nerve injury. In other
words, histologic evidence of nerve inflammation may only have been
detected at currents greater than 0.2 mA had a pulse duration of 0.1 msec
been chosen instead of 0.3 msec. In addition, the lack of an expected
motor response based on distance to the nerve in the work by Tsai and
colleagues14 may have resulted from the nerves being stimulated
initially with a high-intensity current before the amperage was reduced.
Such a protocol may have resulted in hyperpolarization of the nerves and
false-negative motor responses. In the work by Rigaud and colleagues,15
the animals had previously underwent coronary occlusion and infusion
of sildenafil for a separate experiment, and although it is unclear if this
would have influenced their results, it is known that ischemia is a risk
factor for nerve injury. Although no histologic examination for injury
occurred in this study, it is possible that the ischemia influenced the
motor response. This study was also underpowered, which could have
accounted for the lack of statistical difference in distance from the nerve
between the high and low MSC groups. Finally, the timing of assessment
varied between the studies from 0 to 7 days. In the original study by
Voelkel and colleagues,12 histologic assessment of injury was made at
only 6 hours postinsult, whereas peak inflammation is reported to occur
at up to 72 hours,34,35 thereby potentially underestimating the extent of
injury in this study.
In combination, this recent animal evidence suggests that nerve
stimulation alone does not reliably reflect the distance between the
needle tip and the nerve nor distinguish intraneural from extraneural
needle tip location. Furthermore, nerve stimulation cannot prevent
intraneural needle placement. Finally, no work has examined whether
the MSC varies with intrafascicular or extrafascicular needle placement.
Peripheral nerves consist of multiple sensory and motor fascicles, and
the anatomy can be complex, with multiple intercrossing fascicles.36 If
the motor fascicle lies to one side of the axon, it is conceivable that the
direction of approach and subsequent distance from the needle tip to the
motor fascicle may affect the MSC. This, in addition to all the other
factors that affect the absolute MSC, would be difficult to control for in
studies. Nevertheless, these animal results do suggest that the MSC is a
specific but not a sensitive detector of intraneural needle placement.
Therefore, although the minimum safety threshold currents that
represent intraneural and extraneural needle tip location have not been
conclusively established in animals, an MSC less than 0.2 mA appears
likely to indicate intraneural placement. Furthermore, this MSC in
animals was more likely to result in histologic injury. The insensitivity
of nerve stimulation, however, implies that the needle may be
intraneural, yet a high MSC is still required to elicit a motor response.
Interestingly, some of these findings have been mirrored in some recent
small clinical studies. Bigeleisen et al.,37 for example, investigated the
relationship between MSC and intraneural needle placement by
comparing intraneural and extraneural stimulation thresholds in
ultrasound-guided supraclavicular block. The MSC with intraneural
needle placement was indeed between 0.2 and 0.5 mA in 54% of patients
and even higher than 0.5 mA in a smaller percentage. The MSC,
however, was never less than 0.2 mA when the needle was located
extraneurally. Robards et al.38 also observed that intraneural injection
often occurred despite a lack of motor response in the sciatic popliteal
nerve. Finally, in agreement with the animal needle-to-nerve proximity
data, Perlas et al.39 demonstrated that despite visualizing the needle next
to but not inside the nerves in the axilla, a motor response was not
always obtained. It is conceivable in fact that in such circumstances
redirecting the needle to obtain a motor response may increase the risk
of nerve injury. Interestingly, in the animal study by Steinfeldt et al. 19
comparing approaching the nerve at either 0.2 or 1.0 mA, needle-to-
nerve contact was significantly greater in the low MSC group (94%
compared to 6%) and so was histologic nerve injury. Block success,
however, in this animal study was not examined, but this does perhaps
indicate that seeking a low MSC may in fact be harmful. Electrical
impedance monitoring appears promising, but at present, it requires that
nerve puncture must occur before a change in impedance is detected
(therefore risking direct mechanical injury), and it cannot detect needle-
to-nerve proximity. However, electrical impedance monitoring may at
least warn providers of intraneural needle tip placement before injecting,
thereby potentially avoiding further injury. Further studies and clinical
investigation of this tool are required, however.
Injection pressure
Intrafascicular injection appears to be the most harmful form of
intraneural injection. This is partly due to the risk of direct injury to
nerve fascicles by the needle but is also due to the risks associated with
high-pressure injections. The connective tissue perineurium surrounding
a nerve is less compliant than the looser epineurium.40 Injections into
this enclosed space that result in a high pressure can therefore either
rupture the perineurium or exceed capillary occlusion pressure within
the nerve, potentially leading to nerve ischemia. Interestingly, pioneering
animal studies several decades ago suggested a link between high
injection pressure and intraneural, particularly intrafascicular, injection,
although the relationship between high injection pressure and adverse
neurologic consequences at that time was unclear.40 The investigation
into whether any relationship exists between high injection pressure,
intrafascicular injection, and nerve injury has recently been addressed by
a number of researchers and is described in the following text.
In 2004, Hadzic and colleagues20 compared histologic injury between
deliberate intrafascicular and extrafascicular intraneural injections in
canine sciatic nerves. In an “open” model, they found that a high
opening injection pressure (25 psi) was associated with both histologic
and clinical evidence of nerve injury. Not all intrafascicular injections,
however, resulted in a high-pressure injection. Importantly, all
extrafascicular injections were associated with neither high injection
pressures nor evidence of histologic or clinical nerve injury. The same
group of investigators again demonstrated in an open canine sciatic
nerve model that high injection pressures (20 psi) correlated with both
histologic and prolonged clinical nerve injury (at 7 days postinsult)
following intrafascicular injection.22 As before, not all intrafascicular
injections resulted in a high pressure and in fact the proportion of
intrafascicular injections that were high pressure was lower (only 8 of 20
cases). Again, both low-pressure intrafascicular and extrafascicular
injections were not associated with prolonged injury or abnormal
histology. In both these studies, therefore, high-pressure injections only
occurred when the needle was intrafascicular, albeit not all
intrafascicular injections were high-pressure. There have been
contrasting results in other studies, however. For example, a wide range
of intraneural injection pressures (0.07 to 31.5 psi) were found when
porcine femoral and brachial plexus nerves were punctured using
ultrasound guidance in a closed model.18 Of note, the two injections
greater than 25 psi neither appeared to be intrafascicular nor result in
structural damage. Similarly, Lupu et al.23 recently found no clear
correlation between the maximum pressure generated during intraneural
injection and either histologic or clinical nerve injury when
experimenting on the median nerves of pigs in an open model. In this
study, the needle was inserted intraneurally, but it was unclear if the final
needle tip position was intrafascicular or extrafascicular. Nevertheless,
peak injection pressures were all well below 25 psi (ranging from 1.8 to
9.4 psi), yet 7 out of 10 specimens had evidence of axonal damage on
histologic examination. Furthermore, axonal damage ensued following a
maximum injection pressure of only 2.2 psi in one case. Reassuringly
however, functional deficits measured up to 7 days postinsult were
absent in all 10 pigs studied.
The inconsistency between the injection pressure findings may in part be
related to the gross methodologic heterogeneity among the studies. Most
studies were on exposed open nerves,20,22,23 the limitations of which
have previously been discussed. In addition, the “true” intraneural
pressure was never measured, making comparison of absolute values
difficult. Instead, in each study, a laboratory manometer was used to
measure proximal pressure as a surrogate marker of intraneural pressure.
Injection rate, volume of injectate, and needle type may all affect the
proximal pressure generated. These variables differed between studies
and, in addition, deviated from human clinical practice. A slow injection
rate20 or small injectate volume,18,20,22 for example, may underestimate
the true intraneural pressure. Lupu and colleagues23 did inject at a faster
rate, but a shallow needle angle was utilized. This, combined with
extraneural leakage, could perhaps have resulted in longitudinal spread
of the local anesthetic, blunting the maximum pressure measured. In
another study,18 pressure measurement was not continuous, and so high
pressures may have been missed. Nevertheless, it was in this study
where high pressures that did occur were not consistently associated
with injury. Additional confounders such as the use of adrenaline mixed
with local anesthetic in some studies,20,23 as well as using long-bevel
hypodermic needles in others,22 may each independently increase the
risk of nerve injury compared to plain local anesthetic and short-bevel
needles.41,42 Finally, the exact location and accuracy of needle tip and
injectate were somewhat inconsistent. For example, Lupu and
colleagues23 used only direct vision without any additional histologic
confirmation, whereas Altermatt et al. placed the needle intraneurally
under ultrasound guidance but subsequently used histology and ink
injection to confirm the final needle tip position.18 Although Hadzic and
Kapur used a microscope to confirm either intrafascicular or
extrafascicular needle placement, the authors still speculated that not all
“intrafascicular” injections were actually inside the fascicle.20,22 Indeed,
the size of a human nerve fascicle is reported to be between 100 and
1,000 μm, depending on the specific nerve.36,43 As 22G and 25G needles
have approximate diameters of 700 and 500 μm, respectively,23,20,22
intentional and complete intrafascicular needle tip placement is arguably
extremely challenging, even with the benefit of microscopy. This may
explain why Chan and colleagues found only two cases of intrafascicular
staining out of 24 intraneural ink injections despite puncturing all nerves
under direct vision, albeit with a shallow angle of approach.13
The available animal evidence suggests that high injection pressure
appears to be neither a sensitive nor a specific detector of intraneural or,
in particular, intrafascicular injection. High injection pressures, for
example, may also occur when the needle is obstructed in a tendon or in
tissue compressed by the ultrasound transducer.44 Furthermore,
histologic injury has been reported in animals even after a low-pressure
injection.23 Nevertheless, traditionally, anesthesiologists have avoided
high-pressure injections, relying on the subjective “feel” of resistance
during injection to do so. Because the only prolonged clinical injury in
animals occurred with high-pressure injections, avoiding high pressures
remains logical.20 Unfortunately, “tactile” methods are wholly unreliable
in accurately detecting both injection pressure and the exact tissue
location of the needle.45 A simple, commercially available pressure-
monitoring device now exists, but as yet, there have been no large
clinical trials proving its efficacy. As with any such device, however,
adequately powering such a trial would be extremely difficult due to the
huge patient numbers required to detect a difference in the already small
incidence of nerve injury. Nevertheless, although further such work is
required to establish the relationship between injection pressures and
nerve injury, avoiding “high” injection pressures in the meantime, using
an objective measurement device, would seem prudent practice.
Ultrasound
Ultrasound is the latest major advance in nerve localization and it allows
real-time, direct visualization of the needle, nerve, and surrounding
structures.46 In principle, therefore, ultrasound should reduce both
needle-to-nerve contact and intraneural injection, but although it appears
to increase block success,47 large studies have not yet demonstrated a
superior safety profile of ultrasound compared to other traditional
methods of nerve localization.48–50 A small number of animal studies
have primarily studied the ability of ultrasound to detect intraneural
injection.13,18,23
In an open pig brachial plexus model, Chan et al.13 deliberately injected
ink intraneurally and demonstrated that real-time ultrasound
visualization can sensitively detect a change in nerve diameter, even
with as little as 1 mL of injectate. Where limited or no swelling was seen
on ultrasound, histology confirmed scant or no intraneural injection.
Despite intraneural injection, however, no histologic evidence of injury
was found in any of the 24 nerve specimens, although only two of the
intraneural injections subsequently were confirmed to have
intrafascicular ink staining. In a closed porcine model, Altermatt and
colleagues18 similarly were able to detect swelling during a series of
small volume intraneural injections of ink, with nerve diameters
increasing by up to 50%. All injections were confirmed as being
intraneural, but none were intrafascicular. Most recently, Lupu and
coworkers23 undertook a randomized trial using an open porcine median
nerve model and injected a larger, more clinically relevant volume
intraneurally. An increase in cross-sectional area ranging from 158% to
706% was witnessed with ultrasound, but there was no correlation,
however, between nerve expansion and inflammation or injury. Despite
evidence of inflammation in all 10 cases and injury in 6 of these, there
were no signs of functional nerve injury in any of the pigs up to 7 days
after injection. As with the previous studies on nerve stimulation and
injection pressure, similar limitations apply to the ultrasound studies. In
particular, the observed increase in nerve diameter on the studies on
surgically exposed nerves13,23 may not reflect that seen in intact
conditions. Secondly, both Chan and colleagues13 and Altermatt and
colleagues18 injected only small volumes of ink (up to 5 mL), which
may have accounted for the lack of histologic injury. When one of these
groups addressed this limitation in their more recent randomized trial by
injecting a more clinically relevant volume (up to 20 mL), all nerves did
show histologic evidence of inflammation.23 In this study, however, it
was not known if the final needle position was intrafascicular. Unlike
electrical impedance and injection pressure monitoring, which are only
abnormal when the needle tip is already inside the nerve, and nerve
stimulation, which does not reliably reflect needle-to-nerve proximity,
the major advantage of ultrasound is that direct visualization of the
needle tip and target should facilitate prevention of intraneural injection.
Unfortunately, this depends on the operator maintaining the needle tip in
view at all times, which can be difficult, particularly in deeper blocks.51
In fact, inadvertent intraneural injection during real-time ultrasound-
guided blocks has now been frequently described in the human
literature.10,52–54 Echogenic needle tips may help reduce this,
particularly for deeper blocks with steeper needle angles, but no large
studies have yet examined this topic.55 Similarly, guidance positioning
system tracking and needle guidance is now available, which again in
theory may improve the accuracy of needle tip positioning with respect
to the nerve. The animal studies discussed describe the other function of
ultrasound should inadvertent intraneural injection occur, namely, the
ability to observe for nerve expansion. Given that ultrasound can detect
very small volumes of injectate, ultrasound may in theory limit the
extent of intraneural injection, which may in turn reduce intrafascicular
pressure, although whether such early detection could reduce subsequent
nerve injury is unknown. Detecting subtle nerve swelling, however, may
not be straightforward, given that in a recent case series of ultrasound-
guided interscalene and supraclavicular blocks, the incidence of
unintentional intraneural injection was as 17%.5 In this study, only half
of these were actually recognized by the operator, and the injection
stopped with the other half noted only during retrospective video review.
As with other emerging small series of clinical data, none had a
neurologic complication. A further important drawback of ultrasound is
that the resolution of modern portable machines is not high enough to
differentiate between potentially hazardous intrafascicular injection and
the potentially less dangerous extrafascicular compartment. It is
conceivable that the lower compliance intrafascicular compartment
would not swell to the same extent as an extrafascicular injection, but
there have not been ultrasound descriptions of such injections to date.
Conclusions
Although intrafascicular intraneural injection appears to be most
harmful, the fact that extrafascicular injection and even forced needle-to-
nerve contact are also associated with histologic nerve injury means that
it is imperative that needle position is known throughout the
performance of a nerve block. Unfortunately, however, recent animal
studies, as described earlier, and some human studies suggest that both
traditional and novel markers of needle-to-nerve proximity are
insensitive techniques of detecting both intraneural injection and needle-
to-nerve contact. Firstly, regarding nerve stimulation, there is no
absolute MSC that sensitively and specifically excludes intraneural
needle placement, although animal and some preliminary human studies
suggest that an MSC of less than 0.2 mA is associated with both
intraneural injection and a higher risk of histologic nerve injury.
Electrical impedance monitoring may be useful in detecting intraneural
injection but only once the nerve is entered; therefore, this tool still risks
mechanical nerve damage and cannot prevent intraneural injection. Both
nerve stimulation and electrical impedance do not reliably inform the
operator of needle-to-nerve distance or prevent needle to nerve contact.
Once inside the nerve, high injection pressures appear suggestive of
intrafascicular needle placement, but again, this measurement does not
always reliably correlate with intraneural or intrafascicular injection.
Injection pressure monitoring again clearly provides no information on
needle-to-nerve distance. Nevertheless, avoiding high-pressure
injections, which have been associated with histologic nerve injury,
would appear sensible. Ultrasound guidance is best able to detect both
intraneural injection and needle-to-nerve proximity but cannot prevent
needle-nerve puncture. Ultimately, safety depends not only on the tools
used but also the clinical ability and skill of the operator. Although nerve
stimulation, electrical impedance, and pressure monitoring all provide
objective values, the overriding benefit and theoretical advantages of
ultrasound are limited by operator inadequacies, even in experienced
individuals. Using a combination of some or all of the aforementioned
techniques may prove important in achieving the ultimate clinically
relevant goal, avoiding nerve injury. The relative contribution of each of
these modalities toward achieving this goal, however, remains unknown,
and much further work in humans is required before stringent guidance
can be provided.
References
1. Brull R, McCartney CJ, Chan VW, et al. Neurological
complications after regional anesthesia: contemporary estimates of
risk. Anesth Analg. 2007;104:965–974.
2. Hogan QH. Pathophysiology of peripheral nerve injury during
regional anesthesia. Reg Anesth Pain Med. 2008;33:435–441.
3. Neal JM, Bernards CM, Hadzic A, et al. ASRA Practice Advisory
on Neurologic Complications in Regional Anesthesia and Pain
Medicine. Reg Anesth Pain Med. 2008;33:404–415.
4. Hara K, Sakura S, Yokokawa N, et al. Incidence and effects of
unintentional intraneural injection during ultrasound-guided
subgluteal sciatic nerve block. Reg Anesth Pain Med. 2012;37:289–
293.
5. Liu SS, YaDeau JT, Shaw PM, et al. Incidence of unintentional
intraneural injection and postoperative neurological complications
with ultrasound-guided interscalene and supraclavicular nerve
blocks. Anaesthesia. 2011;66:168–174.
6. Bigeleisen PE. Nerve puncture and apparent intraneural injection
during ultrasound-guided axillary block does not invariably result
in neurologic injury. Anesthesiology. 2006;105:779–783.
7. Macfarlane AJ, Bhatia A, Brull R. Needle to nerve proximity: what
do the animal studies tell us? Reg Anesth Pain Med. 2011;36:290–
302.
8. Whitlock EL, Brenner MJ, Fox IK, et al. Ropivacaine-induced
peripheral nerve injection injury in the rodent model. Anesth Analg.
2010;111:214–220.
9. Steinfeldt T, Poeschl S, Nimphius W, et al. Forced needle
advancement during needle-nerve contact in a porcine model:
histological outcome. Anesth Analg. 2011;113:417–420.
10. Schafhalter-Zoppoth I, Zeitz ID, Gray AT. Inadvertent femoral
nerve impalement and intraneural injection visualized by
ultrasound. Anesth Analg. 2004;99:627–628.
11. Neal JM, Wedel DJ. Ultrasound guidance and peripheral nerve
injury: is our vision as sharp as we think it is? Reg Anesth Pain
Med. 2010;35:335–337.
12. Voelckel WG, Klima G, Krismer AC, et al. Signs of inflammation
after sciatic nerve block in pigs. Anesth Analg. 2005;101:1844–
1846.
13. Chan VW, Brull R, McCartney CJ, et al. An ultrasonographic and
histological study of intraneural injection and electrical stimulation
in pigs. Anesth Analg. 2007;104:1281–1284.
14. Tsai TP, Vuckovic I, Dilberovic F, et al. Intensity of the stimulating
current may not be a reliable indicator of intraneural needle
placement. Reg Anesth Pain Med. 2008;33:207–210.
15. Rigaud M, Filip P, Lirk P, et al. Guidance of block needle insertion
by electrical nerve stimulation: a pilot study of the resulting
distribution of injected solution in dogs. Anesthesiology.
2008;109:473–478.
16. Steinfeldt T, Graf J, Vassiliou T, et al. High or low current threshold
for nerve stimulation for regional anaesthesia. Acta Anaesthesiol
Scand. 2009;53:1275–1281.
17. Steinfeldt T, Graf J, Vassiliou T, et al. Systematic evaluation of the
highest current threshold for regional anaesthesia in a porcine
model. Acta Anaesthesiol Scand. 2010;54:770–776.
18. Altermatt FR, Cummings TJ, Auten KM, et al. Ultrasonographic
appearance of intraneural injections in the porcine model. Reg
Anesth Pain Med. 2010;35:203–206.
19. Steinfeldt T, Graf J, Schneider J, et al. Histological consequences of
needle-nerve contact following nerve stimulation in a pig model.
Anesthesiol Res Pract. 2011;2011:591851.
20. Hadzic A, Dilberovic F, Shah S, et al. Combination of intraneural
injection and high injection pressure leads to fascicular injury and
neurologic deficits in dogs. Reg Anesth Pain Med. 2004;29:417–
423.
21. Vuckovic I, Hadzic A, Dilberovic F, et al. Detection of
neurovascular structures using injection pressure in blockade of
brachial plexus in rat. Bosn J Basic Med Sci. 2005;5:79–85.
22. Kapur E, Vuckovic I, Dilberovic F, et al. Neurologic and histologic
outcome after intraneural injections of lidocaine in canine sciatic
nerves. Acta Anaesthesiol Scand. 2007;51:101–107.
23. Lupu CM, Kiehl TR, Chan VW, et al. Nerve expansion seen on
ultrasound predicts histologic but not functional nerve injury after
intraneural injection in pigs. Reg Anesth Pain Med. 2010;35:132–
139.
24. Scholz T, Pharaon M, Evans GR. Peripheral nerve anatomy for
regeneration studies in pigs: feasibility of large animal models. Ann
Plast Surg. 2010;65:43–47.
25. Mackinnon SE, Hudson AR, Llamas F, et al. Peripheral nerve injury
by chymopapain injection. J Neurosurg. 1984;61:1–8.
26. Greenblatt GM, Denson JS. Needle nerve stimulator-locator: nerve
blocks with a new instrument for locating nerves. Anesth Analg.
1962;41:599–602.
27. Tsui B, Hadzic A. Peripheral nerve stimulators and
electrophysiology of nerve stimulation. In: Hadzic A, ed. Textbook
of Regional Anesthesia and Acute Pain Management. New York,
NY: McGraw Hill; 2007:93–104.
28. Sauter AR, Dodgson MS, Stubhaug A, et al. Ultrasound controlled
nerve stimulation in the elbow region: high currents and short
distances needed to obtain motor responses. Acta Anaesthesiol
Scand. 2007;51:942–948.
29. Li J, Kong X, Gozani SN, et al. Current-distance relationships for
peripheral nerve stimulation localization. Anesth Analg.
2011;112:236–241.
30. Hadzic A, Vloka J, Hadzic N, et al. Nerve stimulators used for
peripheral nerve blocks vary in their electrical characteristics.
Anesthesiology. 2003;98:969–974.
31. Hadzic A. Peripheral nerve stimulators: cracking the code—one at a
time. Reg Anesth Pain Med. 2004;29:185–188.
32. Tsui BC, Pillay JJ, Chu KT, et al. Electrical impedance to
distinguish intraneural from extraneural needle placement in
porcine nerves during direct exposure and ultrasound guidance.
Anesthesiology. 2008;109:479–483.
33. Sung DH. Locating the target nerve and injectate spread in rabbit
sciatic nerve block. Reg Anesth Pain Med. 2004;29:194–200.
34. Mueller M, Leonhard C, Wacker K, et al. Macrophage response to
peripheral nerve injury: the quantitative contribution of resident and
hematogenous macrophages. Lab Invest. 2003;83:175–185.
35. Mueller M, Wacker K, Ringelstein EB, et al. Rapid response of
identified resident endoneurial macrophages to nerve injury. Am J
Pathol. 2001;159:2187–2197.
36. Stewart JD. Peripheral nerve fascicles: anatomy and clinical
relevance. Muscle Nerve. 2003;28:525–541.
37. Bigeleisen PE, Moayeri N, Groen GJ. Extraneural versus intraneural
stimulation thresholds during ultrasound-guided supraclavicular
block. Anesthesiology. 2009;110:1235–1243.
38. Robards C, Hadzic A, Somasundaram L, et al. Intraneural injection
with low-current stimulation during popliteal sciatic nerve block.
Anesth Analg. 2009;109:673–677.
39. Perlas A, Niazi A, McCartney C, et al. The sensitivity of motor
response to nerve stimulation and paresthesia for nerve localization
as evaluated by ultrasound. Reg Anesth Pain Med. 2006;31:445–
450.
40. Selander D, Sjostrand J. Longitudinal spread of intraneurally
injected local anesthetics. An experimental study of the initial
neural distribution following intraneural injections. Acta
Anaesthesiol Scand. 1978;22:622–634.
41. Selander D, Brattsand R, Lundborg G, et al. Local anesthetics:
importance of mode of application, concentration and adrenaline
for the appearance of nerve lesions. An experimental study of
axonal degeneration and barrier damage after intrafascicular
injection or topical application of bupivacaine (Marcain). Acta
Anaesthesiol Scand. 1979;23:127–136.
42. Selander D, Dhuner KG, Lundborg G. Peripheral nerve injury due
to injection needles used for regional anesthesia. An experimental
study of the acute effects of needle point trauma. Acta Anaesthesiol
Scand. 1977;21:182–188.
43. Sunderland S, Bradley KC. The perineurium of peripheral nerves.
Anat Rec. 1952;113:125–141.
44. Gadsden J, McCally C, Hadzic A. Monitoring during peripheral
nerve blockade. Curr Opin Anaesthesiol. 2010;23:656–661.
45. Theron PS, Mackay Z, Gonzalez JG, et al. An animal model of
“syringe feel” during peripheral nerve block. Reg Anesth Pain Med.
2009;34:330–332.
46. Chin KJ, Chan V. Ultrasound-guided peripheral nerve blockade.
Curr Opin Anaesthesiol. 2008;21:624–631.
47. Abrahams MS, Aziz MF, Fu RF, et al. Ultrasound guidance
compared with electrical neurostimulation for peripheral nerve
block: a systematic review and meta-analysis of randomized
controlled trials. Br J Anaesth. 2009;102:408–417.
48. Barrington MJ, Watts SA, Gledhill SR, et al. Preliminary results of
the Australasian Regional Anaesthesia Collaboration: a prospective
audit of more than 7000 peripheral nerve and plexus blocks for
neurologic and other complications. Reg Anesth Pain Med.
2009;34:534–541.
49. Liu SS, Ngeow JE, Yadeau JT. Ultrasound-guided regional
anesthesia and analgesia: a qualitative systematic review. Reg
Anesth Pain Med. 2009;34:47–59.
50. Fredrickson MJ, Kilfoyle DH. Neurological complication analysis
of 1000 ultrasound guided peripheral nerve blocks for elective
orthopaedic surgery: a prospective study. Anaesthesia.
2009;64:836–844.
51. Chin KJ, Perlas A, Chan VW, et al. Needle visualization in
ultrasound-guided regional anesthesia: challenges and solutions.
Reg Anesth Pain Med. 2008;33:532–544.
52. Cohen JM, Gray AT. Functional deficits after intraneural injection
during interscalene block. Reg Anesth Pain Med. 2010;35:397–399.
53. Reiss W, Kurapati S, Shariat A, et al. Nerve injury complicating
ultrasound/electrostimulation-guided supraclavicular brachial
plexus block. Reg Anesth Pain Med. 2010;35:400–401.
54. Russon K, Blanco R. Accidental intraneural injection into the
musculocutaneous nerve visualized with ultrasound. Anesth Analg.
2007;105:1504–1505.
55. Hebard S, Hocking G. Echogenic technology can improve needle
visibility during ultrasound-guided regional anesthesia. Reg Anesth
Pain Med. 2011;36:185–189.
13
Ultrasound-Guided
Intraneural Injection
—The Human Data
MEG A. ROSENBLATT AND PAUL E.

BIGELEISEN

Intraneural injections were once considered harbingers for neural injury,


and the practice of regional anesthesia focused on their avoidance.
Stimulation techniques centered on maximizing proximity of needles
and catheters to nerves without piercing them,1,2 and studies examined
the risks to nerves involved with elicitation of paresthesias.3 The use of
ultrasound for regional anesthesia has offered new insights into the
relationship between intraneural (into the epineurium) and
intrafascicular (within the perineurium) injections of local anesthesia
and their relationship with postoperative neurologic complications.
Initial studies looked at the effect of intraneural injections in animal
models, but more recently, their effect in humans has been described.
Bigeleisen4 promoted the idea that intraneural injections of local
anesthesia did not necessarily yield postoperative neural dysfunction. He
performed ultrasound-guided axillary nerve blocks on 26 patients
undergoing surgery on the base of the thumb. Using a 22G short-beveled
needle under direct visualization, he attempted to inject each of the four
nerves (radial, median, ulnar, musculocutaneous) with 2 to 3 mL of local
anesthetic. Nerve swelling was considered evidence of an intraneural
injection and was observed in 72 of the 104 injections, whereas the
remaining injections occurred immediately outside the epineurium (Fig.
13.1). Complete surgical anesthesia was achieved in 100% of patients.
Postoperatively and at 6-month follow-up examinations, none of the
patients reported paresthesias or dysesthesias in the distribution of the
four injected nerves. Bigeleisen offers the explanation that fascicles in
the axillary nerves are separated by large amounts of stroma between the
fascicles. Intraneural injections performed with blunt needles then do not
result in neurologic damage because the nerves are able to swell and
because the needles do not readily penetrate the perineurium, which is
much stronger than the epineurium (Fig. 13.2).5 There are other reports
of intraneural injections without neurologic sequelae. Upon review of
video ultrasound images of an axillary block, Russon and Blanco noted
that an intraneural injection of 7 mL of levobupivacaine into the
musculocutaneous nerve had occurred.6 That patient reported no
subjective or objective neurologic deficit 6 months after the event. An
inadvertent intraneural injection of 35 mL of local anesthesia into the
femoral nerve, with no adverse sequelae, has also been published.7 In a
separate study, 34 patients received ultrasound-guided intraneural
supraclavicular blocks. Included in this cohort were seven patients with
long-standing diabetes, three of whom had polyneuropathy. All of these
patients also had successful blocks without any measurable neurologic
injuries at 6 months.8
Although a “donut” sign may represent adequate spread of local
anesthesia for many nerve blocks, this is not true for an interscalene
block. In an attempt to define what constitutes adequate local anesthesia
spread for an interscalene block, Spence et al.9 compared periplexus
injection (in the potential space between the middle scalene muscle and
the brachial plexus sheath) to an injection into the plexus (between the
two most superficial large hypoechoic circles on ultrasound). One
hundred sixty-six patients were studied and followed until the resolution
of the block with confirmation that there were no persistent sensory or
motor deficits. They had a 95% success rate in the intraplexus and 93%
success rate in the periplexus groups and no persistent deficits in either
group. Interestingly, there was a 16% longer duration of action of the 30
mL 0.5% bupivacaine blocks seen in the intraplexus group. The authors
hypothesize that this may be secondary to the sheath serving as a
reservoir for the volume of local anesthetic.9 Some believe that
intraplexus injection during interscalene block is inevitable. Orebaugh
performed a routine ultrasound-guided interscalene injections using
India ink in 10 preserved shoulders, then dissected and examined the
results. In five of the specimens, there was histologic evidence of ink
below the epineurium.10 In a single-blind observational study of 257
patients who underwent shoulder surgeries with either interscalene or
supraclavicular blocks, anesthesiologists reviewing the videos of the
blocks found that 42 patients (17%) had intraneural injections. None of
these patients was found to have a postoperative neurologic
complication at a 4-to-6-week follow-up.11
In an interesting study, Robards et al.12 used ultrasound to look at the
incidence of intraneural injection using low-current stimulation. They
took 24 subjects undergoing surgeries of the foot or ankle and performed
an ultrasound-guided popliteal block with the end point to initiate the
injection being either a motor response <0.2 × <0.5 mA or the needle
being visualized within the nerve. In 20 patients, stimulation was only
achieved once the needle was below the epineurium, whereas in the
remaining patients, the needle was seen within the nerve, yet no
response to stimulation was elicited with the stimulator at 1.5 mA. No
neurologic complications occurred.12 Forty-two patients who were
undergoing hallux valgus repair under popliteal nerve block were
studied by Sala Blanch et al.,13 who assessed the incidence of
subepineural injection. An intraneural injection was defined as a 15% or
greater increase in nerve area and having ultrasonographic evidence of
either nerve swelling or proximal–distal diffusion within epineural tissue
after the performance of a nerve block with a stimulatory technique.
These criteria were met by 28 (66%) patients in the study, none of whom
had any evidence of neurologic complications.13 Most recently, this
research group looked at the rate of onset of the popliteal block,
comparing blocks performed using an ultrasound and placed between the
tibial and common peroneal (fibular) nerves at the level of their division,
to a traditional nerve stimulator block, performed 7 cm above the
popliteal fossa crease. All 52 patients had a successful block at 30
minutes, but those in the ultrasound-intraneural injection group had a
statistically significantly shorter onset, with 80% versus 4% having a
sensory block and 60% versus 8% having a motor block 15 minutes after
injection.14
Whether anesthesiologists are advocating the use of intraneural nerve
blocks or eschewing the practice, videos promoting its use are
admittedly or unknowingly readily available throughout the Internet.
Examples of intraneural placement of local anesthesia during
interscalene,15 supraclavicular,16 femoral,17 and popliteal18 blocks are
often what the novice ultrasonographer is accessing when attempting to
learning these techniques, and thus, this practice may be more
widespread than acknowledged.
The use of ultrasound offers the ability to visualize what has historically
been performed blindly but with a low incidence of complications.
Although it is premature to conclude that intraneural injections are
without neurologic sequelae, ultrasound is helping to elucidate the
multifactorial etiology of postoperative neurologic dysfunction. Much
larger prospective studies will be needed to determine an outcome-based
approach to ultrasound-guided intraneural injection.

References 1. Gurnaney H, Ganesh A, Cucchiaro G. The


relationship between current intensity for nerve stimulation and
success of peripheral nerve blocks performed in pediatric patients
under general anesthesia. Anesth Analg. 2007;105:1605–1609.
2. Pham-Dang C, Kick O, Collet T, et al. Continuous peripheral nerve
blocks with stimulating catheters. Reg Anesth Pain Med.
2003;28:83–88.
3. Liguori GA, Zayas VM, YaDeau JT, et al. Nerve localization
techniques for interscalene brachial plexus blockade: a prospective,
randomized comparison of mechanical paresthesia versus electrical
stimulation. Anesth Analg. 2006;103:761–767.
4. Bigeleisen PE. Nerve puncture and apparent intraneural injection
during ultrasound-guided axillary block does not invariably result
in neurologic injury. Anesthesiology. 2006;105:779–783.
5. Moayeri N, Bigeleisen PE, Groen G. Quantitative architecture of
the brachial plexus and surrounding compartments: their possible
implications for plexus blocks. Anesthesiology. 2008;108:299–305.
6. Russon K, Blanco R. Accidental intraneural injection into the
musculocutaneous nerve visualized with ultrasound. Anesth Analg.
2007;105:1504–1505.
7. Schafhalter-Zoppoth I, Zeitz ID, Gray AT. Inadvertent femoral
nerve impalement and intraneural injections visualized by
ultrasound. Anesth Analg. 2004;99:627–628.
8. Bigeleisen PE, Moayeri N, Groen GJ. Extreneural versus intraneural
stimulation thresholds during ultrasound-guided supraclavicular
block. Anesthesiology. 2009;110:1229–1234.
9. Spence BC, Beach ML, Gallagher JD, et al. Ultrasound-guided
interscalene blocks: understanding where to inject the local
anesthetic. Anaesthesia. 2011;66:509–514.
10. Orebaugh SL, McFadden K, Skorupan H, et al. Subepineurial
injection in ultrasound-guided interscalene needle tip placement.
Reg Anesth Pain Med. 2010;35:450–454.
11. Liu SS, YaDeau JT, Shaw PM, et al. Incidence of unintentional
intraneural injection and postoperative neurological complications
with ultrasound-guided interscalene and supraclavicular nerve
blocks. Anaesthesia. 2011;66:168–174.
12. Robards C, Hadzic A, Somasundaram L, et al. Intraneural injection
with low-current stimulation during popliteal sciatic nerve block.
Anesth Analg. 2009;109:673–677.
13. Sala Blanch X, Lopez AM, Carazo J, et al. Intraneural injection
during nerve-stimulator guided sciatic nerve block at the popliteal
fossa. Br J Anaesth. 2009;102:855–861.
14. Sala-Blanch X, deRiva N, Carrera A, et al. Ultrasound-guided
popliteal sciatic block with a single injection at the sciatic division
results in faster block onset than the classical nerve stimulator
technique. Anesth Analg. 2012;114:1121–1127.
15. Shibata Y. Ultrasound-guided interscalene brachial plexus block.
August 2008. Available at: http://www.youtube.com/watch?
v=w0Nwm687hf0&feature=fvst. Accessed October 22, 2014.
16. Shibata Y. Ultrasound-guided supraclavicular brachial plexus block.
July 2008. Available at: http://www.youtube.com/watch?
v=YPRSDElCcqA. Accessed October 22, 2014.
17. Shibata Y, Horato M. Ultrasound-guided femoral nerve block. June
2008. Available at: http://www.youtube.com/watch?
v=Bxj0LyQdy_M. Accessed October 22, 2014.
18. Sites M. Ultrasound guided sciatic nerve block: the popliteal
approach. December 2011. Available at:
http://www.youtube.com/watch?v=2pzsKjwRpr0. Accessed
October 22, 2014.
14
Use of Ultrasound for
Placement of
Peripheral
NerveBlock
Catheters
R. SCOTT LANG, PAUL E. BIGELEISEN, AND

STEVEN L. OREBAUGH

The insertion of catheters percutaneously to allow the infusion of local


anesthetic solutions was initially described by Ansbro in 1946.1
Techniques for placement of catheters have evolved from elicitation of
paresthesias, to fascial clicks,2 and subsequently to fluoroscopy,3 nerve
stimulator, and ultrasound imaging. The most popular method for
inserting catheters is a through-the-needle technique (Fig. 14.1),
although catheter-over-the-needle techniques have been reported as well.
Placement of a continuous catheter offers some distinct advantages
compared to single-injection techniques. In particular, the duration of
nerve blockade is a major advantage. Although there are many adjuncts
that have been studied to increase the duration of single-injection
peripheral nerve blockade, none have been shown to consistently result
in analgesia for multiple days to weeks as can catheter techniques.4,5 A
catheter can also permit titration of the dose of local anesthetic to
increase analgesia or decrease motor blockade to encourage participation
in physical therapy. Catheters also provide flexibility in discontinuation
and continuation of infusion as dictated by patient care. These benefits
have encouraged the use of catheters throughout the body.

Continuous nerve blockade is used to provide analgesia for surgeries of


the upper extremity (Fig. 14.2), thoracic and abdominal areas (Fig.
14.3), and lower extremity (Fig. 14.4). Upper extremity applications
include blocks at the interscalene, supraclavicular, cervical
paravertebral, suprascapular, infraclavicular, and axillary sites. Thoracic
and abdominal catheter techniques include paravertebral, intercostal,
epidural, transversus abdominis plane, rectus sheath, and
ilioinguinal/iliohypogastric blocks. Finally, catheters for lower extremity
surgeries may be inserted at the lumbar plexus, femoral, parasacral,
paravertebral, lateral femoral cutaneous, obturator, fascia iliaca, sciatic,
and saphenous regions. The use of ultrasound helps identify the nerves
in many of these techniques and allows flexibility in the approach to
each. As experience with continuous catheter techniques in a myriad of
anatomic regions has accumulated, so has the literature on proper
placement, use, and complications.
Peripheral nerve blockade has been associated with neurologic injuries;
catheter knotting6–8; catheter migration and dislodgement; catheter
obstruction; fluid leakage; neural irritation9; inflammation; colonization;
abscess formation; sepsis; hematoma formation10,11; myotoxocity12; and
catheter shearing, breakage, and retention. Inaccurate placement of
catheters has also been reported, with catheter tips too far from the
nerve, intraneural placement, intravascular insertion, inadvertent
intrathecal or epidural penetration, and intrapleural placement.
Considering these risks and complications, it is essential to ensure
proper placement of the catheter and confirm its position, especially if
they are to be used in the ambulatory setting.
Benefits of ultrasound Ultrasound technology has enabled the regional
anesthesiologist to place catheters under real-time imaging and to assess position
after successful placement (Fig. 14.5). This has become the standard of care in
many institutions for many peripheral nerve block techniques, especially those
areas in close proximity to the neuraxis or blood vessels. Ultrasound guidance
and nerve stimulation techniques remain popular techniques to place continuous
catheters efficiently and effectively. Although nerve stimulation techniques
continue to be utilized, either alone or in conjunction with ultrasound techniques,
there are times when nerve stimulation may be undesirable due to discomfort
from local muscle contraction. Compared to nerve stimulation techniques,
ultrasound guidance has been shown to improve success rates,13–16 onset,17,18
and quality17,19 and also to decrease needle passes/attempts20 and time to block
completion.15,19 In addition, increased duration of block has been described,21
along with improved patient satisfaction,18,20 reduced dose of local
anesthetic,21–24 and reduced opioid consumption.20

Some evidence has accumulated suggesting that ultrasound guidance


facilitates more effective catheter placement. Mariano et al.25
demonstrated more rapid insertion of the popliteal-sciatic catheters, with
a significantly lower failure rate among 40 patients undergoing foot–
ankle surgery. The results of another study involving 40 patients
receiving infraclavicular brachial plexus catheters also demonstrated
shorter time to block completion, higher success rate, and fewer vascular
punctures.26 Bendsten and colleagues20 examined sciatic nerve block
catheters, utilizing either ultrasound or electrical stimulation for
guidance in 100 patients. The results were similar, showing an increased
success rate in sensory block, fewer needle passes, decreased opioid
consumption, and increased patient satisfaction. Fredrickson et al.27
examined 82 patients undergoing shoulder surgery for which
interscalene brachial plexus block was utilized. Between the groups
receiving ultrasound guidance and nerve stimulation for placement,
ultrasound imaging provided an increase in effectiveness during the first
24 hours, with fewer needle passes/manipulation and less patient
discomfort during insertion. As data continues to be gathered, the
benefits of the use of ultrasound to place perineural catheters are
increasingly highlighted, potentially making it the standard of practice
for the placement of most peripheral nerve catheters.
Although the cost of ultrasound equipment is high, studies have shown a
cost–benefit in its utilization.28 Further, appropriate billing in many
regions allows additional remuneration for ultrasound-guided
techniques, which may offset or negate the cost of ultrasound equipment.
Sandhu and colleagues29 described a significant decrement in cost when
using an ultrasound approach instead of a nerve-stimulator approach.
Some of the decrease in cost stems from reduced operating room or
block time as well as the cost of the disposable equipment (the authors
did not use stimulating needles for ultrasound blocks). In addition, the
cost of ultrasound use was mitigated by suggesting its versatility
regarding vascular access in addition to use with regional anesthesia.
Studies that examine several cost factors have been published that help
support the use of ultrasound for regional anesthesia in certain common
clinical scenarios.30
The use of ambulatory catheters, placed with ultrasound guidance, has
increased in many institutions and improves patient satisfaction after
ambulatory surgery,31 with very effective pain control. Patients are able
to go home with continuous peripheral nerve block catheters and pumps
while successfully managing and removing them with minimal contact
with anesthesiologists via telephone (Fig. 14.6). It is likely more
important to ensure proper placement when a patient is to be sent home
with a low likelihood of health care professional supervision on a daily
basis as compared to managing inpatients. Ultrasound placement and
confirmation helps establish functioning catheters and adequate pain
control prior to being discharged home.
Regional anesthesia in the pediatric patient is growing in popularity as
more studies are looking at the effect that general anesthetics have on the
developing mind. Because many of the landmark techniques used with
nerve stimulation have been developed in adults, many of the
generalized guidelines on insertion sites are not applicable to pediatric
patients. Using a standardized measure of distance as established in adult
patients may cause significant error in technique and place vulnerable
tissue in the pediatric patient at risk. However, use of ultrasound
guidance to visualize the nerve or associated landmarks helps to ensure
more accurate execution. Ultrasound has been used to place catheters
successfully and safely in pediatric patients for years and plays a unique
role in confirming the positioning of epidural catheters placed through
the caudal canal, which can be inserted up to lumbar and even thoracic
levels. Accumulation of fluid in the epidural space, or the structure of
the catheter itself, may be visualized to ensure accurate placement.
Ultrasound imaging also permits estimation of tissue depths for deeper
and more complex nerve blocks, such as the distance to the transverse
processes for lumbar plexus catheter placement. Catheters placed under
ultrasound guidance allow the use of less systemic anesthetic during
surgical procedures as well as help limit the need for narcotic use in the
postoperative period.
References
1. Ansbro FP. A method of continuous brachial plexus block. Am J
Surg. 1946;71:716–722.
2. Selander D. Catheter technique in axillary plexus block.
Presentation of a new method. Acta Anaesthesiol Scand.
1977;21:324–329.
3. Pham-Dang C, Meunier JF, Poirier P, et al. A new axillary approach
for continuous brachial plexus block. A clinical and anatomic study.
Anesth Analg. 1995;81:686–693.
4. Burgoyne LL, Pereiras LA, Bertani LA, et al. Long-term use of
nerve block catheters in paediatric patients with cancer related
pathologic fractures. Anaesth Intensive Care. 2012;40:710–713.
5. Borghi B, D’Addabbo M, White PF, et al. The use of prolonged
peripheral neural blockade after lower extremity amputation: the
effect on symptoms associated with phantom limb syndrome.
Anesth Analg. 2010;111:1308–1315.
6. Burgher AH, Hebl JR. Minimally invasive retrieval of knotted
nonstimulating peripheral nerve catheters. Reg Anesth Pain Med.
2007;32:162–166.
7. Offerdahl MR, Lennon RL, Horlocker TT. Successful removal of a
knotted fascia iliaca catheter: principles of patient positioning for
peripheral nerve catheter extraction. Anesth Analg. 2004;99:1550–
1552.
8. David M. Knotted peripheral nerve catheter. Reg Anesth Pain Med.
2003;28:487–488.
9. Ribeiro FC, Georgousis H, Bertram R, et al. Plexus irritation caused
by interscalene brachial plexus catheter for shoulder surgery.
Anesth Analg. 1996;82:870–872.
10. Wiegel M, Gottschaldt U, Hennebach R, et al. Complications and
adverse effects associated with continuous peripheral nerve blocks
in orthopedic patients. Anesth Analg. 2007;104:1578–1582.
11. Jöhr M. A complication of continuous blockade of the femoral
nerve. Reg Anaesth. 1987;10:37–38.
12. Hogan Q, Dotson R, Erickson S, et al. Local anesthetic
myotoxicity: a case and review. Anesthesiology. 1994;80:942–947.
13. Bendtsen TF, Nielsen TD, Rohde CV, et al. Ultrasound guidance
improves a continuous popliteal sciatic nerve block when compared
with nerve stimulation. Reg Anesth Pain Med. 2011;36:181–184.
14. Sites BD, Beach ML, Spence BC, et al. Ultrasound guidance
improves the success rate of a perivascular axillary plexus block.
Acta Anaesthesiol Scand. 2006;50:678–684.
15. Dingemans E, Williams SR, Arcand G, et al. Neurostimulation in
ultrasound-guided infraclavicular block: a prospective randomized
trial. Anesth Analg. 2007;104:1275–1280.
16. Chan VW, Perlas A, McCartney CJ, et al. Ultrasound guidance
improves success rate of axillary brachial plexus block. Can J
Anaesth. 2007;54:176–182.
17. Marhofer P, Schrögendorfer K, Koinig H, et al. Ultrasonographic
guidance improves sensory block and onset time of three-in-one
blocks. Anesth Analg. 1997;85:854–857.
18. Marhofer P, Sitzwohl C, Greher M, et al. Ultrasound guidance for
infraclavicular brachial plexus anaesthesia in children. Anaesthesia.
2004;59:642–646.
19. Williams SR, Chouinard P, Arcand G, et al. Ultrasound guidance
speeds execution and improves the quality of supraclavicular block.
Anesth Analg. 2003;97:1518–1523.
20. Bendtsen TF, Nielsen TD, Rohde CV, et al. Ultrasound guidance
improves a continuous popliteal sciatic nerve block when compared
with nerve stimulation. Reg Anesth Pain Med. 2011;36:181–184.
21. Oberndorfer U, Marhofer P, Bösenberg A, et al. Ultrasonographic
guidance for sciatic and femoral nerve blocks in children. Br J
Anaesth. 2007;98:797–801.
22. Marhofer P, Schrögendorfer K, Wallner T, et al. Ultrasonographic
guidance reduces the amount of local anesthetic for 3-in-1 blocks.
Reg Anesth Pain Med. 1998;23:584–588.
23. Willschke H, Marhofer P, Bösenberg A, et al. Ultrasonography for
ilioinguinal/iliohypogastric nerve blocks in children. Br J Anaesth.
2005;95:226–230.
24. Casati A, Baciarello M, Di Cianni S, et al. Effects of ultrasound
guidance on the minimum effective anaesthetic volume required to
block the femoral nerve. Br J Anaesth. 2007;98:823–827.
25. Mariano ER, Cheng GS, Choy LP, et al. Electrical stimulation
versus ultrasound guidance for popliteal-sciatic perineural catheter
insertion: a randomized controlled trial. Reg Anesth Pain Med.
2009;34:480–485.
26. Mariano ER, Loland VJ, Bellars RH, et al. Ultrasound guidance
versus electrical stimulation for infraclavicular brachial plexus
perineural catheter insertion. J Ultrasound Med. 2009;28:1211–
1218.
27. Fredrickson MJ, Ball CM, Dalgleish AJ. A prospective randomized
comparison of ultrasound guidance versus neurostimulation for
interscalene catheter placement. Reg Anesth Pain Med.
2009;34:590–594.
28. Ehlers L, Jensen JM, Bendtsen TF. Cost-effectiveness of ultrasound
vs nerve stimulation guidance for continuous sciatic nerve block.
Br J Anaesth. 2012;109:804–808.
29. Sandhu NS, Sidhu DS, Capan LM. The cost comparison of
infraclavicular brachial plexus block by nerve stimulator and
ultrasound guidance. Anesth Analg. 2004;98:267–268.
30. Liu SS, John RS. Modeling cost of ultrasound versus nerve
stimulator guidance for nerve blocks with sensitivity analysis. Reg
Anesth Pain Med. 2010;35:57–63.
31. Swenson JD, Bay N, Loose E, et al. Outpatient management of
continuous peripheral nerve catheters placed using ultrasound
guidance: an experience in 620 patients. Anesth Analg.
2006;103:1436–1443.
Section II
Upper Extremity
15 Ultrasound-Guided Interscalene Block Using the
Posterior Approach

16 Ultrasound-Guided Interscalene Block

17 Ultrasound-Guided Supraclavicular Block

18 Ultrasound-Guided Suprascapular and Axillary


Nerve Block
Part 1: Suprascapular Nerve Block
Part 2: Ultrasound-Guided Selective Axillary
Nerve Block

19 Ultrasound-Guided Infraclavicular Block

20 Ultrasound-Guided Axillary Block

21 Ultrasound-Guided Blocks at the Elbow and


Forearm


15
Ultrasound-Guided
Interscalene Block
Using the Posterior
Approach
DAVID BYRNES AND PATRIK
FILIP

Introduction and indications: The posterior approach to the roots of the


brachial plexus was originally described by Kappis1 in 1923, who
argued that a posterior approach was safest because the carotid, internal
jugular, vertebral vessels, sympathetic ganglia, phrenic, and recurrent
laryngeal nerves are located anterior to the brachial plexus. Studies by
Sandefo et al.2 and Pere et al.3 showed an overall success rate of 97%,
including 100% block of the axillary and radial nerves, 97% block of the
median and musculocutaneous nerves, and 68% block of the ulnar nerve.
They also showed a 0.5% phrenic nerve block compared to the near
100% rate of phrenic nerve block with the lateral approach and only a
7% superior sympathetic ganglion block. The technique was little used
before it was revised by Pippa et al.4 in 1990 and again by Boezaart5 in
2003 with modifications to make it less painful by directing the needle
between muscle groups. The block is used for surgery of the shoulder
and proximal arm when an indwelling catheter distant from the surgical
site is desired.
Anatomy: The brachial plexus is composed of the ventral roots of spinal
nerves C5–T1. The roots exit the lateral foraminal spaces and pass
between the anterior and middle scalene muscles to innervate the upper
limb (Fig. 15.1). At the root level, the fascicles are surrounded by
dura/perineurium. Within the perineurium, there is little or no stroma, so
care must be taken to not position the needle within the nerve root itself.
For this reason, the authors recommend using an 18G Tuohy needle even
for single-injection techniques. The carotid artery and jugular vein
should be identified as well to prevent their puncture (Fig. 15.2).
Patient position: Sitting or lateral decubitus.
Transducer: Linear (25 or 38 mm) oscillating at 13 MHz. Curved (11
mm) oscillating at 10 MHz.
Transducer orientation: Transverse, over the sternocleidomastoid at the
level of the thyroid cartilage.

Needle: 5 or 10 cm, 18G Tuohy needle with a 20G catheter.


Local anesthetic: 10 mL of 0.2% ropivacaine or 0.25% bupivacaine.
Technique: The patient is positioned in either the lateral or sitting
position with the neck flexed and the face turned away from the block
side. Some practitioners choose to use electrostimulation along with
ultrasound. The anesthesiologist stands behind the patient with the
ultrasound machine placed in front of the patient. The interscalene
brachial plexus is located with the ultrasound probe and the needle entry
site identified. Historically, the posterior spinous process of C8 was
identified by palpation, and the space between the trapezius and levator
scapulae muscles was used as the point of entry (Fig. 15.3A). With the
advent of ultrasound, it is usually easier to place the probe in the
supraclavicular fossa and then trace the plexus cephalad to the level
where the roots are readily visible (Fig. 15.3B).
The skin over the supraclavicular fossa and neck are washed with sterile
solution. A 25-or 38-mm, 13-MHz probe covered with a sterile sheath is
placed in the supraclavicular fossa and the subclavian artery and brachial
plexus are identified. The probe is then moved in a cephalad direction
until the roots of C5, C6, and C7 are identified as three round
hypoechoic spheres (Fig. 15.4). In most cases, the carotid artery and
jugular vein can also be seen. In some cases, the roots C8 and T1 may
also be identified in the same image. With proper toggling of the probe,
the sulcus of the transverse process can also be identified (Fig. 15.4). A
25G, 4-cm needle is used to anesthetize the skin as well as the path to
the brachial plexus under ultrasound guidance using an entrance point
immediately posterior to the probe. This makes the subsequent
placement of the 18G Tuohy needle painless. The infiltration needle is
removed and a 5-cm, 18G Tuohy needle is advanced along the same
path until the tip of the needle lies immediately posterior to the fifth and
sixth nerve roots (Fig. 15.5). Ten to 25 mL of local anesthetic is then
injected in aliquots of 5 mL with care taken to make sure that the needle
tip is neither intraneural or intravascular. At the end of the injection, the
roots should be surrounded with local anesthetic. Once the injection is
complete, the catheter is inserted through the needle and then the needle
is withdrawn over the catheter (Fig. 15.6). The catheter is then fixed in
place with adhesive bandages and bio-occlusive dressings. Some
catheters may be visible on ultrasound (Fig. 15.6). Some practitioners
prefer to insert the catheter next to the plexus before injecting any local
anesthetic around the nerve roots. In some cases, the catheter may be
visualized on ultrasound. If the catheter cannot be visualized, local
anesthetic is injected through the catheter to ensure proper location. As
local anesthesia is injected, the pooling of local anesthesia around the
nerve roots can be visualized on color-flow Doppler imaging. The
catheter is secured in place and covered with a sterile occlusive dressing.
The catheter is then connected to a pump for continuous postoperative
infusion. For most brachial plexus catheters, an infusion of 8 mL/hr
results in adequate spread of local anesthetic. An infusion of lidocaine
0.5% or ropivacaine 0.2% provides adequate analgesia with a minimum
of motor blockade. Patient-controlled regional anesthesia (PCRA) is a
very safe and effective means of providing adequate analgesia and
excellent patient satisfaction through a peripheral nerve catheter. In this
case, the infusion is run at 6 mL/hr with a 2 mL bolus available every 30
to 60 minutes. Anesthesia to the phrenic nerve; recurrent laryngeal
nerve; and the sympathetic ganglia resulting in dyspnea, hoarseness, and
Horner syndrome may occur as side effects. If the catheter is used to
deliver the dose of local anesthetic as described earlier, instead of a large
bolus through the needle, the incidence of phrenic nerve, recurrent
laryngeal nerve, and the sympathetic ganglia being anesthetized is
greatly diminished. Occasionally, the catheter may migrate into a vessel,
and the staff must remain vigilant for this complication.
Tips 1. Some practitioners prefer to insert the catheter into the Tuohy
needle before inserting the needle into the skin. Once the needle is in the
proper position, the catheter can then be inserted beyond the end of the
needle after which the needle is removed.
2. Spiral-bound or flex-tip catheters can usually be visualized on
ultrasound.
3. For surgery of the shoulder, it is best to position the catheter next to
C5. When surgery or analgesia of the forearm is desired, it is best to
position the catheter at the level of C7.

References 1. Kappis M. Weitere Erfahrungen mit der


Sympathektomie. Klin Wehr. 1923;2:1441.
2. Sandefo I, Iohom G, Van Elstraete A, et al. Clinical efficacy of the
brachial plexus block via the posterior approach. Reg Anesth Pain
Med. 2005;30:238–242.
3. Pere P, Pitkänen M, Rosenberg PH, et al. Effect of continuous
interscalene brachial plexus block on diaphragm motion and on
ventilatory function. Acta Anaesthesiol Scand. 1992;36:53–57.
4. Pippa P, Cominelli E, Marinelli C, et al. Brachial plexus block using
the posterior approach. Eur J Anaesthesiol. 1990;7:411–420.
5. Boezaart A. Paravertebral approach to the brachial plexus: an
anatomic improvement in technique. Reg Anesth Pain Med.
2003;28:241–244.
16
Ultrasound-Guided
Interscalene Block
STEVEN L. OREBAUGH, RYAN GUFFEY, NIZAR

MOAYERI, AND PAUL E. BIGELEISEN

Background and indications: Winnie1 first described the interscalene


block in 1970. This block is ideal for surgery of the shoulder, distal
clavicle, or proximal humerus. It provides anesthesia to the superior
elements of the brachial plexus, including the suprascapular nerve and
the supraclavicular nerve (C4). It frequently spares the ulnar nerve
completely and preserves some motor and sensory function of the radial
and median nerves of the forearm and hand. The interscalene block may
not cover the superficial cervical plexus completely. This can be
supplemented as necessary for surgery involving the lower neck. If
surgery includes the anteroinferior or posteroinferior surface of the
upper arm, a separate block of the intercostal brachial and medial
cutaneous nerves (T1–T2) may be required. Surgery involving the elbow
may be accomplished with the interscalene nerve block as described in
the following text with the addition of a second injection of local
anesthetic to the inferior trunk of the brachial plexus.
Anatomy: Block of the brachial plexus at this level requires placement of
local anesthetic in the interscalene groove, a potential space between the
anterior and middle scalene muscles, which is occupied by the nerve
elements of the plexus, as well as the subclavian artery. This space
typically lies just posterior to the lateralmost extent of the
sternocleidomastoid muscle and is exposed when the patient turns the
head to the contralateral side (Fig. 16.1). The groove between the
anterior and middle scalene muscles can be palpated in most thin
patients. It may be difficult to palpate in obese patients and in those
patients with limited neck mobility.

In the interscalene space, the roots of C5–T1 coalesce to form the


superior (C5–C6), middle (C7), and inferior (C8–T1) trunks, which
proceed laterally and inferiorly toward the space between the clavicle
and first rib and then into the axilla (Fig. 16.2). Several important
branches are released from the brachial plexus at this level, including the
suprascapular nerve, the dorsal scapular nerve, and the long thoracic
nerve.
On ultrasound, an appreciation of the anatomy at the level of the roots
begins with imaging of a reliable landmark at the base of the neck—that
is, the subclavian artery and the brachial plexus, which lies dorsolateral
to the artery and superior to the rib (Fig. 16.3). This requires placement
of the transducer first in the supraclavicular fossa, in a sagittal oblique
orientation (Fig. 16.3). At this level, the nerves appear as a cluster of
grapes dorsal and lateral (posterior) to the artery. Upon establishing
these landmarks, the transducer should be moved slowly cephalad,
tilting the probe more horizontally and following the nerve plexus
proximally. This position will reveal the nerve elements (roots) aligning
vertically between the scalene muscles (Figs. 16.4 through 16.6). The
sternocleidomastoid muscle at this position is usually very attenuated
and is visible as a triangular slip of muscle lying superficial to the plexus
and scalene muscles. At this level, the practitioner may be imaging
trunks (Fig. 16.4) or roots (Figs. 16.5 and 16.6). In a few patients, the C5
nerve root may pass anterior to or directly through the anterior scalene
muscle (Fig. 16.7).
The interscalene space is lined by evaginated prevertebral fascia, which
proceeds distally with the plexus as a discontinuous sheath (Fig. 16.1).1,4
This fascia is not typically evident on ultrasound imaging; rather, the
nerve roots and trunks are sandwiched between the scalene muscles and
appear as hypoechoic nodules (Fig. 16.6).2 In most patients, roots C5,
C6, and C7 are readily visible in the same image. In some patients, all
five roots are visible (Fig. 16.6). The nerves, which are separated from
the carotid sheath by the anterior scalene muscle, are relatively
superficial in most subjects and lie at a mean depth of 5.5 mm from the
skin surface (Fig. 16.1).3
Superficial to the muscles, just beneath the skin, the external jugular
vein is often apparent but is easily compressed. Gentle “bobbing” up and
down with the transducer may be required to demonstrate this small
vein. Fortunately, it is seldom in the path of the blocking needle with this
technique. Another vessel of importance in this region, coursing
transversely across the space, is the transverse cervical artery and
associated vein (Fig. 16.1).4 The dorsal scapular artery, a branch of the
transverse cervical artery, can also often be found coursing between the
individual trunks of the brachial plexus. A muscular structure in this
space, which may be confused with a vessel, is the inferior belly of the
omohyoid muscle.
Patient position: Supine with the head turned away from the practitioner,
lateral decubitus position.
Transducer type: A 2-to 4-cm linear probe oscillating at 6 to 13 MHz.
Transducer orientation: Transverse over the sternocleidomastoid
muscle at the C6 level or lateral to the thyroid cartilage.

Needle: The authors recommend a 22G short-bevel needle for single-shot


blockade or an 18G Tuohy needle for continuous blockade. A 5-cm
needle is appropriate for a posterior approach with needle insertion next
to the probe, but a 9-cm or longer needle may be required for a more
posterior needle insertion or if a 4-cm ultrasound probe is utilized. A
longer needle may improve needle visualization by allowing a shallower
needle angle but in inexperienced hands can also increase risk by
allowing deeper penetration to critical structures.
Local anesthetic: 10 to 20 mL of 0.2% to 0.75% ropivacaine or 0.125%
to 0.5% bupivacaine. Smaller concentrations are associated with less
phrenic nerve blockade but delayed onset and decreased length of
shoulder analgesia.5 This can be ideal for loading a catheter in a patient
that will have general anesthesia. Higher concentrations are more
appropriate for single-shot or surgical blocks.
Technique: Oxygen by face mask is initiated. The patient monitors are
placed and a brief preprocedural “time-out” is conducted at the bedside.
The anesthesiologist marks the patient’s operative extremity and
sedation is administered. Antiseptic is used to prepare the skin of the
block area.
Given the relatively superficial nature of the nerves to be imaged, a
high-frequency transducer (6 to 13 MHz) is desired. The transducer is
placed in the supraclavicular fossa in the coronal oblique position. The
subclavian artery, brachial plexus, and first rib are identified, and the
transducer is used to follow the nerve plexus proximally up the neck
until it lies in an axial oblique position. An optimal image of the brachial
plexus roots is obtained by toggling the transducer, usually with slight
caudad tilt (Fig. 16.4).
Ultrasound-guided interscalene block may be conducted with or without
nerve stimulator confirmation. The needle may be inserted either
posterior (Fig. 16.4) or anterior (Fig. 16.8) to the transducer using an in-
plane approach. One may also use an out-of-plane approach in which the
transducer position is identical but the needle is introduced from a
position superior or inferior to the midpoint of the transducer.

After establishing an appropriate image of the nerve roots, the skin is


anesthetized, along with the subcutaneous tissues. The authors prefer a
5-cm needle using the posterior in-plane approach. The block needle is
then introduced along the predicted pathway, with subtle transducer
adjustments to maintain an image of both the nerve and needle. The
nerve stimulator, if used, is turned on before needle introduction, or
later, as the needle approaches the nerves. The threshold for stimulation
has not been found to be of importance in this technique.6 However,
lower currents are associated with inadvertent intraneural injection.7
Also, a deltoid or triceps response is associated with a higher success
rate than biceps stimulation alone.8 After appropriate positioning of the
needle between the C5 and C6 nerve roots, or obtaining the desired
motor response, local anesthetic solution is injected, typically 20 mL,
ensuring that the injectate surrounds the visible nerve elements. If
necessary, the needle position may be shifted to ensure that all of the
nerve roots are surrounded by local anesthetic.
For indwelling catheter placement, a similar technique is commonly
used. In addition, a sterile sheath should be used for placement of an
indwelling device, along with sterile gloves, cap, mask, and sterile
towels for drapes. The needle insertion point should be as far from the
surgical site as possible while still in plane. This is often possible just
below the lateral edge of the hairline. A 4-cm ultrasound probe and a 10-
cm needle make this technique easier. After visualization of the nerve
roots and positioning of the probe so that the desired needle insertion
point is in plane with the transducer, the skin should be anesthetized.
The needle should be advanced as above and a bolus of local anesthetic,
saline, or dextrose solution (3 to 20 mL) can be used to confirm accurate
needle tip placement and create space for the catheter to thread. If the
catheter to be placed is of the stimulating variety, then the stimulator is
attached to the catheter after needle localization, and placement is
confirmed by continued stimulation as the catheter is inserted. By either
technique, after catheter placement, ultrasound imaging can be used to
confirm that the solution injected through the catheter is deposited in
juxtaposition to the nerves of interest. Color-flow Doppler may improve
the ability to image in this situation. Alternatively, a small bubble of air
(less than 1 mL) can be injected through the catheter slowly. One can
follow the bubble under ultrasound and visualize the entire course of the
catheter. An out-of-plane technique can also be used to place the catheter
insertion point below the ear. If the nerves lie at a depth of 2 cm, then a
needle introduced 2 cm proximal to the midpoint of the transducer will
require an angle of insertion of 45 degrees to encounter the target. For
deeper targets, a steeper angle or more superior insertion point is
required. The needle is then guided to the nerve roots, and nerve
stimulation is used for confirmation if desired.
To secure the catheter, multiple techniques are possible. The author
prefers to use benzoin and two adhesive strips to sandwich the catheter
laterally at the insertion point. This is followed by small tegaderms that
extend around the neck to the opposite side. There the catheter is
labeled, coiled, and fixed with a medium-sized tegaderm. Possible
modifications include tunneling the catheter with a 16G angiocath or
using dermabond to seal the skin insertion point to the catheter. Aquacel
can also be placed over the steri-strips, but under the tegaderm, to
decrease leakage.
Anatomical variance: Kessler and Gray9 conducted a study of
volunteers in which the plexus and its surrounding anatomy were
visualized in the interscalene region with ultrasonography. In 13% of
plexuses, variations from the typical relationship of the scalene muscles
and brachial plexus roots were present, the most common being the C5
nerve root running through or outside the anterior scalene muscle. The
authors postulate that this arrangement may be responsible for
occasional incomplete interscalene blocks. In case of block failure, the
C5 root can be identified by first identifying roots C6 and C7. The C7
root (Fig.16.9A) has a single posterior tubercle. The C6 root often has a
double head and there is both anterior and posterior tubercle
(Chassaignac tubercle) at the C6 level. Once C6 and C7 have been
identified, then the proceduralist can scan up the neck to the C5 root.
The authors recommend blockade of the nerve root as far from its
emergence from the vertebral foramen as possible to prevent inadvertent
epidural spread.
Complications of interscalene block: The interscalene nerve block is
very safe when performed by an experienced practitioner, but it has been
described as the superficial peripheral nerve block associated with the
highest risk of complications. It is thought that ultrasound guidance
decreases the incidence of rare severe complications, but studies large
enough to evaluate this have not been performed. Common
complications include neck hematoma and Horner syndrome. Rare
complications previously reported include pneumothorax,10 pulmonary
edema,11 superficial cervical plexus neuropathy,12 brachial plexus
entanglement (by catheter),13 epidural spread,14 spinal injection,15 and
syrinx formation. The latter may result in permanent quadriplegia.16 It is
thought that pneumothorax could be completely preventable with the use
of ultrasound, but inability to completely visualize the needle can result
in pleural puncture.10 Phrenic nerve paralysis has been previously
described as 100%, but modifications to the procedure such as lowering
the local anesthetic concentration, decreasing the dose volume, or
targeting the C7 instead of the C5 and C6 nerve roots have resulted in
lower incidences. The rate of temporary neuropathy approaches 3%.
This is higher than other commonly performed nerve blocks including
femoral and axillary.17
References
1. Winnie AP. Interscalene brachial plexus block. Anesth Analg.
1970;49:455–466.
2. Perlas A, Chan VW, Simons M. Brachial plexus examination and
localization using ultrasound and electrical stimulation.
Anesthesiology. 2003;99:429–435.
3. Yang WT, Chui PT, Metreweli C. Anatomy of the normal brachial
plexus revealed by sonography and the role of sonographic
guidance in anesthesia of the brachial plexus. AJR Am J
Roentgenol. 1998;171:1631–1636.
4. Demondion X, Herbinet P, Boutry N, et al. Sonographic mapping of
the normal brachial plexus. AJNR Am J Neuroradiol.
2003;24:1303–1309.
5. Thackeray EM, Swenson JD, Gertsch MC, et al. Diaphragm
function after interscalene brachial plexus block: a double-blind,
randomized comparison of 0.25% and 0.125% bupivacaine. J
Shoulder Elbow Surg. 2013;22:381–386.
6. Sinha SK, Abrams JH, Weller RS. Ultrasound-guided interscalene
needle placement produces successful anesthesia regardless of
motor stimulation above or below 0.5 mA. Anesth Analg.
2007;105:848–852.
7. Bigeleisen PE, Moayeri N, Groen GJ. Extraneural versus intraneural
stimulation thresholds during ultrasound-guided supraclavicular
block. Anesthesiology. 2009;110:1235–1243.
8. Borgeat A, Ekatodramis G, Guzzella S, et al. Deltoid, triceps, or
both responses improve the success rate of the interscalene catheter
surgical block compared with the biceps response. Br J Anaesth.
2012;109:975–980.
9. Kessler J, Gray AT. Sonography of scalene muscle anomalies for
brachial plexus block. Reg Anesth Pain Med 2007;32:172–173.
10. Mandim BL, Alves RR, Almeida R, et al. Pneumothorax post
brachial plexus block guided by ultrasound: a case report. Rev Bras
Anestesiol. 2012;62:741–747.
11. Betts A, Eggan JR. Unilateral pulmonary edema with interscalene
block. Anesthesiology. 1998;88:1113–1114.
12. Fredrickson MJ. Superficial cervical plexus neuropathy with
chronic pain after superficial cervical plexus block and interscalene
catheter placement. Reg Anesth Pain Med. 2011;36:206.
13. Bowens C Jr, Briggs ER, Malchow RJ. Brachial plexus entrapment
of interscalene nerve catheter after uncomplicated ultrasound-
guided placement. Pain Med. 2011;12:1117–1120.
14. Scammell SJ. Case report: inadvertent epidural anaesthesia as a
complication of interscalene brachial plexus block. Anaesth
Intensive Care. 1979;7:56–57.
15. Yanovski B, Gaitini L, Volodarski D, et al. Catastrophic
complication of an interscalene catheter for continuous peripheral
nerve block analgesia. Anaesthesia. 2012;67:1166–1169.
16. Benumof JL. Permanent loss of cervical spinal cord function
associated with interscalene block performed under general
anesthesia. Anesthesiology. 2000;93:1541–1544.
17. Brull R, McCartney CJ, Chan VW, et al. Neurological
complications after regional anesthesia: contemporary estimates of
risk. Anesth Analg. 2007;104(4):965–974.
17
Ultrasound-Guided
Supraclavicular
Block
STEVEN L. OREBAUGH AND PAUL E.

BIGELEISEN

Background and indications: Supraclavicular brachial plexus block was


originally described by Kulenkampff in 1911.1 After initial popularity,
the block fell into disfavor due to the relatively high risk of
pneumothorax. The compact structure of the plexus is an advantage to
nerve block at this level. With the advent of ultrasound, supraclavicular
block has become the most popular and common technique to provide
anesthesia and analgesia to the entire upper extremity distal to the
shoulder.
Anatomy: After traversing the interscalene groove, the trunks of the
brachial plexus cross the supraclavicular fossa and divide into anterior
and posterior divisions at or distal to the level of the clavicle (Fig. 17.1).
With ultrasound, the most important landmark for the supraclavicular
block is the subclavian artery, which is readily imaged in cross section as
it lies atop the bright, hyperechoic first rib (Fig. 17.2).
The brachial plexus has a fascial investment as it courses laterally to the
arm. Winnie and DiFranco referred to this structure as a “sheath,”
although its existence as such has been challenged.2–4 Nevertheless, a
connective tissue investment around the plexus is common and has been
described as an evagination of the prevertebral fascia (deep cervical
fascia) at cervical levels.4 This connective tissue covering is not a
simple, continuous tube, nor is it impermeable to leakage of contrast or
local anesthetic.5
At this level, the deep cervical fascia is adherent to the epineurium on
the superior surface of the plexus. Winnie4 and DiFranco referred to this
connective tissue as the plexus sheath. Puncture of the sheath (long
arrows in Fig. 17.3) in this location results an intraneural injection with
a very rapid onset. Inferior to the plexus, the plexus is covered only by
the epineurium (arrowheads in Fig. 17.3). In this location, injection
between the rib and the plexus (eight ball corner pocket) is an
extraneural injection with a slower onset. Below the rib, a lack of echoes
represents the air-filled lung. In some patients, the hyperechoic line of
the pleura may be seen as well (Fig. 17.4). Adjacent to the subclavian
artery are the elements of the brachial plexus, arrayed as a group of
hypoechoic nodules posterior to the vessel, frequently described as a
“cluster of grapes.” Although the fascicles may be distinct, the
epineurium (Fig. 17.4), which lines and separates the trunks or divisions,
is not readily discernable. In some patients, it may be difficult to image
the inferior trunk, which lies immediately superior to first rib.
Just anterior to the artery, a hypoechoic structure is seen, which may be
mistaken for a vessel but does contain some internal echoes. This
represents the anterior scalene muscle, which descends from the neck to
insert on the first rib (Fig. 17.3). Further anteriorly, the subclavian vein
is seen at its confluence with the internal jugular vein. This vein can be
difficult to visualize with a linear probe, as it lies anterior to the artery
and deep to the clavicle. If the practitioner wishes to visualize the vein, a
small curved transducer is the best choice. In some patients, valves can
be imaged within the vein. In some patients, the omohyoid muscle can
be visualized as well (Fig. 17.2).
Patient position: Supine with the head turned away from the patient.
Transducer: 25 mm or 38 mm linear transducer oscillating at 13 MHz;
11 mm curved transducer oscillating at 10 MHz.
Transducer orientation: Coronal oblique position just anterior to the
clavicle.

Needle: 50-mm, 22G blunt needle.


Local anesthetic: 15 to 25 mL of 0.5% ropivacaine or 0.5% bupivacaine
with epinephrine.
Technique: The patient monitors are placed, and oxygen and sedation are
administered to the patient. The operative arm is marked by the
anesthesia team after an appropriate time-out to identify the correct
surgical site. An antiseptic is utilized to prepare the skin of the block
area.
For supraclavicular block, the transducer is placed in a coronal oblique
position just anterior to the clavicle (Fig. 17.2). Usually, the transducer
must be held relatively vertically in order to obtain transverse images of
the subclavian artery and optimal images of the nearby nerve plexus.
Due to the proximity of the needle to the lung and pleura in
supraclavicular block, an in-plane approach to this ultrasound-guided
block is desirable. After establishing the ideal imaging plane for the
subclavian artery and the nerves lying next to it, the skin either anterior
or posterior to the transducer is anesthetized with a subcutaneous needle.
The author prefers to utilize a 5 cm needle while imaging to establish the
insertion trajectory of the block needle. If desired, peripheral nerve
stimulation may be used to confirm needle contact with appropriate
motor stimulation. Local anesthetic solution is injected upon
confirmation of the appropriate needle tip position. It is essential that the
needle tip be imaged at all times in order to avoid pneumothorax.
Although surrounding the nerve structure in question is always
desirable, it is important to inject some of the local anesthetic solution
between the nerve cluster and the first rib.6 This causes the nerves to rise
up and “float” in the liquid, ensuring anesthesia of the inferior trunk.
For placement of an indwelling catheter, an out-of-plane technique
makes threading the catheter easier. As the needle contacts the nerves of
the plexus, the catheter is inserted with or without stimulation guidance.
As at other levels, the physician may choose to combine ultrasound with
a stimulating needle and stimulating catheter.
Injection of dextrose or local anesthetic solution may then be used to
confirm appropriate catheter placement under ultrasound visualization.
Summary of evidence: To quantify the relationships of the brachial
plexus to the surrounding anatomy at the level of the supraclavicular
block, Apan et al.7 performed sonographic examinations in 30 patients
along with coronal magnetic resonance imaging (MRI) in a subset of
these. The authors found that the depth of the brachial plexus from the
skin at this level, with the transducer in the coronal oblique position,
averaged 1.65 cm in males and 1.45 cm in females.
Several authors have utilized sonography in attempts to simplify the
supraclavicular block and improve its safety profile. Kapral et al.8
evaluated 40 patients for forearm or hand surgery, comparing the
supraclavicular approach with the axillary approach to ultrasound-
guided brachial plexus block. Blocks were conducted as single injections
of 30 mL of 0.5% bupivacaine. In both the axillary and the
supraclavicular block groups, 95% successful surgical anesthesia was
achieved. However, 5 of the 20 patients in the axillary block group did
not achieve block of the musculocutaneous nerve.
Chan et al.9 evaluated ultrasound-guided supraclavicular block in 40
outpatients for arm or hand surgery. The authors localized the plexus at
supraclavicular levels with ultrasound and confirmed the needle position
with nerve stimulation and then injected 40 mL of a mixture of 2%
lidocaine and 0.5% bupivacaine. The needle was repositioned as
necessary to facilitate complete surrounding of the elements of the
brachial plexus with local anesthetic solution. Overall, 95% of blocks
were successful, with one failure due to subcutaneous injection and one
due to a partial intravascular injection. The average time required to
complete the block in these patients was 9.0 minutes, with a mean time
to complete sensory and motor blockade of 16.7 minutes.
In a comparative study, Williams et al.10 performed either ultrasound-
guided supraclavicular block with nerve stimulation (for confirmation)
or landmark-guided supraclavicular block with peripheral nerve
stimulation in 80 patients for hand or forearm surgery. Successful
surgical block, with no requirement for supplementation or conversion to
general anesthesia, was reported in 85% of the ultrasound-guided blocks
and 78% of peripheral nerve stimulation-guided blocks, an insignificant
difference. However, the time required to perform the block was
significantly shorter with use of ultrasound (5.0 versus 9.8 minutes). An
average of 21 seconds was required to obtain an adequate ultrasound
image of the supraclavicular brachial plexus. No significant
complications occurred; two patients (one in each group) developed
respiratory discomfort from diaphragmatic paresis. Five patients
reported altered sensation in the affected extremity after the block and
surgery, and all resolved by 1 week.
References
1. Kulenkampff D, Persky MA. Brachial plexus anesthesia: its
indications, technique, and dangers. Ann Surg. 1928;87:883–891.
2. Partridge BL, Katz J, Benirschke K. Functional anatomy of the
brachial plexus sheath: implications for anesthesia. Anesthesiology.
1987;66:743–747.
3. Cornish PB, Leaper C. The sheath of the brachial plexus.
Anesthesiology. 2006;105:563–565.
4. Winnie AP. Interscalene brachial plexus block. Anesth Analg.
1970;49:455–466.
5. Yang WT, Chui PT, Metreweli C. Anatomy of the normal brachial
plexus revealed by sonography and the role of sonographic
guidance in anesthesia of the brachial plexus. Am J Roentgenol.
1998;171:1631–1636.
6. Soares LG, Brull R, Lai J, et al. Eight ball, corner pocket: the
optimal needle position of ultrasound-guided supraclavicular block.
Reg Anesth Pain Med. 2007;32:94–95.
7. Apan A, Baydar P, Yylmaz S, et al. Surface landmarks of brachial
plexus: ultrasound and magnetic resonance imaging for
supraclavicular approach with anatomical correlation. Eur J
Ultrasound. 2001;13:191–196.
8. Kapral S, Krafft P, Eibenberger K, et al. Ultrasound-guided
supraclavicular approach for regional anesthesia of the brachial
plexus. Anesth Analg. 1994;78:507–513.
9. Chan VW, Perlas A, Rawson R, et al. Ultrasound-guided
supraclavicular brachial plexus block. Anesth Analg.
2003;97:1514–1517.
10. Williams S, Chouinard P, Arcand G, et al. Ultrasound guidance
speeds execution and improves the quality of supraclavicular block.
Anesth Analg. 2003;97:1518–1523.
18
Ultrasound-Guided
Suprascapular and
Axillary Nerve Block
Part 1: Suprascapular Nerve Block

BASSEM ASAAD, JONATHAN M. TAN, AND



PAUL E. BIGELEISEN

Background: The suprascapular nerve provides about 70% of the


shoulder joint sensation, and subsequently, it is an important nerve to
target for the acute and chronic pain management of the shoulder joint.
Smaller contributions come from the axillary nerve, musculocutaneous
nerve, subscapular nerve, and the lateral pectoral nerve. Thus, complete
anesthesia of the shoulder joint requires an interscalene block or
proximal supraclavicular block.
The suprascapular nerve block was first described by Wertheim and
Rovenstine1 in 1941 for the treatment of chronic shoulder pain. Since
then, the block has been used extensively by anesthesiologists, pain
medicine specialists, orthopedists, and rheumatologists.
In chronic pain settings, blocking the nerve provides pain management
for rotator cuff pathology, shoulder osteoarthritis, and adhesive
capsulitis. In acute pain settings, it is used to facilitate postoperative
analgesia following shoulder arthroscopy and manipulation of the
shoulder. Patients receiving a suprascapular nerve block for
postoperative analgesia have demonstrated reduced postoperative opioid
use and decreased hospital length of stay.2
Early presentations of suprascapular nerve block sonographic anatomy
in published studies were limited.3–5 Peng et al.6 describe further
sonographic landmarks with correlations to fluoroscopic and cadaveric
anatomy.

Indications: The suprascapular nerve block has been used successfully in


both the acute and chronic pain settings, including rheumatologic
conditions. In the acute pain setting, suprascapular nerve block has been
used as the sole regional anesthetic and also in combination with other
blocks for pain control during both open and arthroscopic shoulder
surgery.4 Although the interscalene brachial plexus block is most
commonly used for these procedures, block of the suprascapular nerve
can provide a safer alternative for patients who cannot tolerate phrenic
or superior laryngeal dysfunction.9,10 Price11 described blocking both the
suprascapular nerve and the axillary nerve as an alternative to
interscalene brachial plexus block for postoperative pain control after
shoulder surgery.
The use of suprascapular nerve block extends to the diagnosis and
treatment of chronic shoulder pain, including neuropathy from
rheumatoid arthritis and osteoarthritis, frozen shoulder, and rotator cuff
tendonitis.12–14 These procedures employ injection of local anesthetics,
steroids, and neurolysis with phenol. Pulsed radiofrequency ablation of
the nerve is also a choice for chronic painful conditions. Combining the
block with the standard rehabilitative therapy may improve the final
outcome in patients with shoulder rotator cuff tendonitis.15

Anatomy: The suprascapular nerve is derived from the upper trunk of the
brachial plexus (C5 and C6). Following this, the suprascapular nerve
passes inferiorly and laterally deep to the omohyoid and trapezius
muscle (Fig. 18.1). The suprascapular nerve travels in a neurovascular
group along with the suprascapular vein and artery until it reaches the
suprascapular notch. The suprascapular nerve gives off two branches to
the supraspinatus muscle, follows it to the lateral aspect of the scapular
spine, and ends in the infraspinatus fossa to supply the infraspinatus
muscle. The suprascapular nerve carries both motor and sensory fibers.
The sensory component provides innervation to the scapula and the
acromioclavicular joint to cover the majority of the shoulder superiorly
and posteriorly.
Patient position: The suprascapular nerve block can be completed with
ultrasound guidance using several various techniques with the two main
approaches being either superior or posterior. The posterior approach is
conducted with the patient in the sitting or prone position with the
ipsilateral arm at the side and has been well documented in the literature.
The main complication of the posterior approach is the risk for
pneumothorax. The superior approach can be conducted in the supine or
sitting position, carries a lower risk of pneumothorax, and is considered
an easier approach.7 The literature also describes the use of fluoroscopy,
electromyography, and computed tomography to improve accuracy of
the suprascapular nerve block. The patient is in a sitting position and the
ipsilateral shoulder is in a neutral position (Fig. 18.1). The individual
performing the block should be positioned behind the patient with an
unobstructed view of the ultrasound monitor.
Equipment
35-mm linear ultrasound probe (13-6 MHz)
Sterile cover for the ultrasound probe
A pack of sterile towels
2% chlorhexidine/70% isopropyl alcohol skin prep
30G × 5/16″ for skin topicalization
21G, 50-mm bevel needle.
Ropivacaine 0.5% (5 to 7 mL)

Technique: The patient is in a sitting position with the standard monitors


applied. The patient should receive oxygen by nasal cannula or face
mask as medically indicated. A “time-out” should be conducted to
confirm the procedure. Appropriate sedation should be administered as
needed. The site of the procedure should have the skin cleaned and be
prepped and draped in the standard fashion. Furthermore, the ultrasound
probe should be prepared in the standard fashion. Lidocaine 1% can be
used to infiltrate the skin and provide analgesia during the procedure.
A linear ultrasound probe should be first placed along the superolateral
aspect of the shoulder in a coronal plane (Fig. 18.1), between the
suprascapular notch and the spinoglenoid notch. The trapezius muscle
can be visualized as superficial to the supraspinatus muscle. The
suprascapular nerve and artery run together beneath the supraspinatus
muscle fascia. The supraspinatus muscle fascia will contain the injected
local anesthetic. The suprascapular nerve is found as a hyperechoic
structure beneath the transverse scapular ligament. The suprascapular
artery can be identified as a pulsatile structure that can be confirmed
with a color Doppler. The scapular fossa is the deep structure over which
the suprascapular nerve and artery course. An in-plane technique is
preferred (Fig. 18.1). The needle is advanced through the skin,
subcutaneous tissue, trapezius, and supraspinatus muscle. Once the
needle tip is beneath the supraspinatus muscle fascia, at the lateral aspect
of the supraspinous fossa, a negative aspiration is obtained. Ropivacaine
0.5% is then injected and the block is complete.
There is significant variation in the reported literature regarding the local
anesthetic volume (5 to 15 mL) that can be injected. Price8 used 5 mL of
local anesthetic and achieved adequate nerve block effect while
minimizing spread of the local anesthetic to the brachial plexus. The
minimization of spread is of paramount clinical significance when this
block is used for chronic pain management.

Complications: General complications related to suprascapular nerve


block include bleeding, infection, and neuropathy. The specific
complications include pneumothorax, intravascular injection, and
muscle atrophy. Brachial plexus block is considered a side effect due to
the potential spread of the injected local anesthetic.
References
1. Wertheim H, Rovenstine E. Suprascapular nerve block.
Anesthesiology. 1941;2:541–545.
2. Ritchie ED, Tong D, Chung F, et al. Suprascapular nerve block for
postoperative pain relief in arthroscopic shoulder surgery: a new
modality? Anesth Analg. 1997;84:1306–1312.
3. Gofeld M. Ultrasonography in pain medicine: a critical review. Pain
Pract. 2008;8:226–240.
4. Yucesoy C, Akkaya T, Ozel O, et al. Ultrasonographic evaluation
and morphometric measurements of the suprascapular notch. Surg
Radiol Anat. 2009;31:409–414.
5. Harmon D, Hearty C. Ultrasound-guided suprascapular nerve block
technique. Pain Physician. 2007;10:743–746.
6. Peng PW, Wiley MJ, Liang J, et al. Ultrasound-guided
suprascapular nerve block: a correlation with fluoroscopic and
cadaveric findings. Can J Anaesth. 2010;57:143–148.
7. Dangoisse MJ, Wilson DJ, Glynn CJ. MRI and clinical study of an
easy and safe technique of suprascapular nerve blockade. Acta
Anaesthesiol Belg. 1994;45:49–54.
8. Price DJ. What local anesthetic volume should be used for
suprascapular nerve block? Reg Anesth Pain Med. 2008;33:571–
573.
9. Singelyn FJ, Lhotel L, Fabre B. Pain relief after arthroscopic
shoulder surgery: a comparison of intraarticular analgesia,
suprascapular nerve block, and interscalene brachial plexus block.
Anesth Analg. 2004;99:589–592, table of contents.
10. Jerosch J, Saad M, Greig M, et al. Suprascapular nerve block as a
method of preemptive pain control in shoulder surgery. Knee Surg
Sports Traumatol Arthrosc. 2008;16:602–607.
11. Price DJ. The shoulder block: a new alternative to interscalene
brachial plexus blockade for the control of postoperative shoulder
pain. Anaesth Intensive Care. 2007;35:575–581.
12. Gorthi V, Moon YL, Kang JH. The effectiveness of
ultrasonography-guided suprascapular nerve block for perishoulder
pain. Orthopedics. 2010:238–241.
13. Eyigor C, Eyigor S, Korkmaz OK, et al. Intraarticular corticosteroid
injections versus pulsed radiofrequency in painful shoulder: a
prospective, randomized, single-blinded study. Clin J Pain.
2010;26:386–392.
14. Jones DS, Chattopadhyay C. Suprascapular nerve block for the
treatment of frozen shoulder in primary care: a randomized trial. Br
J Gen Pract. 1999;49:39–41.
15. Di Lorenzo L, Pappagallo M, Gimigliano R, et al. Pain relief in
early rehabilitation of rotator cuff tendinitis: any role for indirect
suprascapular nerve block? Eura Medicophys. 2006;42:195–204.
Part 2: Ultrasound-Guided
Selective Axillary Nerve

Block
AARON LANGE AND PAUL E. BIGELEISEN

Background and indications: Proximal brachial plexus block, typically


using the interscalene approach, has long been the standard for
perioperative anesthetic and analgesic coverage of the shoulder.
However, it has been proposed that selective block of the axillary and
suprascapular nerves after they branch off the brachial plexus may be
adequate for postoperative analgesia following shoulder surgery while
avoiding the potential side effects of proximal brachial plexus blockade.
Because the axillary nerve supplies the long head of the triceps in some
patients, some practitioners use the technique as a supplement to axillary
plexus block to prevent tourniquet pain. Physiatrists may use the
technique with botulinum toxin for spasticity of the shoulder girdle, or
they may use it in patients with chronic pain conditions.

Anatomy: Sensory innervation to the shoulder is supplied by multiple


sources, including the axillary, suprascapular, subscapular, lateral
pectoral, and musculocutaneous nerves. However, the axillary and
suprascapular nerves are thought to be the most significant contributors
of sensation to the shoulder. The axillary nerve contains fibers from the
ventral rami of C5 and C6 and becomes a terminal branch arising from
the posterior cord. Its motor fibers supply the deltoid muscle (anterior,
middle, and posterior parts), the teres minor muscle, and the long head
of the triceps brachii. It also provides sensory innervation to the skin
over the inferior portion of the deltoid muscle and to the shoulder joint
itself. It emerges at the lateral border of the subscapularis muscle and
passes through the quadrangular space (also called the quadrilateral
space), running in close relation to the posterior humeral circumflex
artery (PHCA) (Fig. 18.2). The nerve runs posteriorly beneath the
shoulder joint, toward the surgical neck of the humerus. It will
ultimately split into an anterior and posterior branch either in the
quadrangular space or later within the deltoid muscle.
Patient position: The patient is positioned sitting with the ipsilateral
shoulder in a neutral position, the humerus is internally rotated about 45
degrees and the arm is flexed at the elbow to about 90 degrees, such that
the hand may rest on the knee.

Transducer: 25-or 50-mm linear transducer oscillating at 13 MHz.

Transducer position: Some practitioners position the probe in a sagittal


position inferior to the acromion. The authors prefer the transverse
position in the lateroposterior position inferior to the head of the
humerus (Fig. 18.2) so that the surface of the humerus is clearly visible.

Needle: Usually a 22G, 50-mm insulated needle is used. For larger or


more muscular patients, a 90-mm needle may be necessary.

Local anesthetic: 10 to 20 mL of 0.5% ropivacaine or 0.5% bupivacaine


with epinephrine.

Technique: The skin is prepped with sterilizing solution and the


transducer is sterilized or covered with a sterile sheath. With the
transducer aligned as described earlier, the PHCA can be located
traversing through the quadrangular space (color-flow Doppler may be
used to help locate the vessel). The teres minor can be identified superior
to the PHCA, the long and lateral heads of the triceps can be found in
longitudinal section deep to the PHCA, and the posterior portion of the
deltoid muscle can be found more superficially (Fig. 18.2). Finally, the
shaft of the humerus can be found as the probe is moved laterally. The
axillary nerve is located in close relation with, and slightly superior to,
the PHCA but may not always be clearly identifiable. Once these
structures are identified, a skin wheal of local anesthetic is raised over
the needle insertion site. Most clinicians utilize an in-plane approach,
with their needle along the axis of the transducer and the insertion site
posterior to the probe. The nerve block needle is inserted through the
skin wheal and advanced to a position in close proximity to the axillary
nerve, superior to the PHCA, and under the muscle fascia. Electrical
stimulation of the axillary nerve will elicit motor response in the
anterior, middle, and posterior parts of the deltoid muscle, causing
abduction of the arm. The axillary nerve is then blocked by injection of
local anesthetic to create a “donut” or “halo” surrounding the nerve
while aspirating frequently to avoid intravascular injection.

Tips

1. The axillary nerve is very small and may be difficult or impossible to


visualize with ultrasound. Use of a nerve stimulator may be employed
for confirmation, but it is likely that adequate blockade can be
achieved by injection of the local anesthetic just superior to the
PHCA, provided the needle has pierced the muscle fascia superficial
to the neurovascular bundle.
2. It should be noted that the bifurcation of the axillary nerve into its
anterior and posterior branches often occurs within the quadrangular
space, and it is therefore possible to only block one branch, resulting
in a partial block. For a complete axillary nerve block, block of both
branches is necessary. It may be prudent to perform the block more
medially in the quadrangular space in order to minimize the chances
of an incomplete block.
Suggested Readings
Rothe C, Asghar S, Andersen HL, et al. Ultrasound-guided block of the axillary nerve: a volunteer
study of a new method. Acta Anaesthesiol Scand. 2011;55:565–570.
Fredrickson MJ, Krishnan S, Chen CY. Postoperative analgesia for shoulder surgery: a critical
appraisal and review of current techniques. Anaesthesia. 2010;65:608–624.
Pitombo PF, Barros RM, Matos MA, et al. Selective suprascapular and axillary nerve block
provides adequate analgesia and minimal motor block. Comparison with interscalene block.
Rev Bras Anestesiol. 2013;63(1):45–51.
19
Ultrasound-Guided
Infraclavicular Block
STEVEN L. OREBAUGH, GERBRAND J.
GROEN, ELILARY MONTILLA MEDRANO,
AND PAUL E. BIGELEISEN

Background and indications: Infraclavicular brachial plexus block


permits anesthesia of the plexus where most of the major motor and
sensory nerves to the arm are anesthetized. In addition, anesthesia of the
phrenic nerve with resultant diaphragmatic paralysis is unlikely.1
Infraclavicular block is used to provide anesthesia and analgesia for
procedures involving the distal arm and elbow, wrist, forearm, and hand.
The lateral ultrasound-guided technique has gained popularity in recent
years because of its high success rate, low incidence of complications,
and ease of performance. Traditionally, practitioners injected local
anesthetic into or around the individual cords. The deposition of a single
injection of local anesthetic posterior to the axillary artery, however, has
been shown to provide reliable anesthesia and is easier to perform than
the multiple-injection technique. Once the visualization of local
anesthetic spread around the axillary artery is achieved, a 90% success
rate is usually accomplished.

Anatomy: Beneath the clavicle, the cervicoaxillary canal is formed,


bounded by the first rib below and the clavicle above. Through this
passageway, the vessels and brachial plexus enter the apex of the axilla.
Overlying this region, the infraclavicular fossa is formed between the
pectoralis major muscle and the deltoid muscle. Needle insertion for
infraclavicular block at this point will traverse the pectoralis major and
pectoralis minor muscles en route to the plexus (Fig. 19.1). Posterior to
the neurovascular bundle is the rib cage and, posteromedially, the pleura
and lung.

The cords of the brachial plexus are closely aligned with the axillary
artery at the infraclavicular region and derive their names from their
position with respect to the vessel: posterior, lateral, and medial.
Because the plexus spirals around the vessel, this relationship may not
be apparent until cords reach the second or third part of the artery (Fig.
19.2). Utilizing magnetic resonance imaging (MRI), Sauter et al.2
evaluated the relative positions of the cords in volunteers. The authors
found that the cords consistently lay within 2.5 cm of the center of the
artery, in a range from directly inferior to the vessel to cephaloanterior,
arranged circumferentially around the artery. The connective tissue
investment, or “sheath,” that defines the space through which the
neurovascular bundle passes has multiple interdigitations and septations,
which may sequester solutions injected near the nerves (Fig. 19.3).3
The infraclavicular plexus lies deeper than at other sites of the brachial
plexus. Wilson et al.4 evaluated the plexus at the pericoracoid region
with MRI and found a mean depth of the plexus elements of 4.2 cm for
men and 4.0 cm for women, although the relationship to body mass
index was not explored. On ultrasound, this greater depth is readily
appreciated and may require use of a lower frequency transducer setting
to provide adequate imaging.5 The cords of the plexus typically appear
hyperechoic, or bright, in the infraclavicular region (Fig. 19.1).
During ultrasound imaging in the infraclavicular fossa, the structures
that appear superficial to the nerves include the skin and subcutaneous
tissues, the pectoralis major and minor muscles, and the clavipectoral
fascia (Fig. 19.1). Deep to this fascia, the second part of the axillary
artery and the axillary vein are apparent. The artery lies cephalad to the
vein. The vein is usually compressible, even at this depth. The
hyperechoic cords of the plexus lie in close approximation to the artery,
typically reflecting their named positions (Fig. 19.1). Posteromedial and
caudad to the nerves and vessels, the hypoechoic region represents the
lung. The pleura may at times be evident due to its hyperechoic nature
and its motion during respiration.

Patient position: Supine with arm at the side or with the arm abducted
and elbow flexed.

Transducer: 11 mm curved array oscillating at 8 to 10 MHz or 25 mm


linear transducer oscillating at 10 MHz.

Transducer orientation: In or lateral to the deltopectoral groove with a


parasagittal orientation.

Needle: 22G, 5-cm blunt tip needle.

Local anesthetic: 15 to 30 mL of 0.5% ropivacaine or 0.5% bupivacaine


with epinephrine.

Technique: The patient monitors are placed, and sedation is introduced.


Oxygen by face mask is administered. A brief preprocedural “time-out”
is conducted at the bedside, and the operative extremity is marked by the
anesthesiologist. An antiseptic is utilized to prepare the skin of the block
area.
The transducer is placed in the deltopectoral groove between the
pectoralis major and deltoid muscles with a parasagittal orientation (Fig.
19.4). The transducer may be moved laterally or medially or toggled to
provide the optimal images of the nerves and vessels. The arm may be
placed at the side or with the arm abducted and the elbow flexed (Fig.
19.4).

Because of the proximity of the tip of the needle to the lung and pleura
during ultrasound-guided infraclavicular block, an in-plane approach is
preferred. A subcutaneous needle is utilized to anesthetize the skin at the
cephalad end of the transducer and to establish the tract that the block
needle will follow under ultrasound guidance. If desired, a peripheral
nerve stimulator is utilized and attached to an insulated block needle. In
most cases, a 22G, 5-cm blunt tip needle may be used. In larger patients,
a 10-cm, 18G Tuohy needle will be required to reach the target.6 As the
needle is introduced, the transducer is adjusted to obtain a view of the tip
during its progress. It is essential to maintain imaging of the tip of the
needle at all times to avoid vascular or pleural puncture.
The needle is then advanced under continuous observation, toward each
of the cords in turn. It is frequently easiest to guide the needle to the
lateral cord first, using nerve stimulation if desired for confirmation (a
musculocutaneous or median nerve–type motor stimulation would be
expected). Local anesthetic, 5 to 10 mL, is then deposited next to the
nerve. Posterior to the artery, the posterior cord is frequently apparent
but must not be confused with artifact that is uniformly present behind
the vessel. Once again, when the needle tip is adjacent to the cord, 5 to
10 mL of local anesthetic is deposited. It is important to ascertain that
local anesthetic is placed posterior to the vessel, and not just on each
side of it, to ensure adequate anesthesia of the posterior cord. Finally, the
needle is brought into position next to the medial cord, and a similar
volume of the local anesthetic is injected. In each case, the physician
may elect to utilize peripheral nerve stimulation to confirm the target.
In order to place an indwelling catheter in the infraclavicular region, a 5-
to 10-cm, 18G Tuohy needle is introduced at the same site as for the
single-injection technique, utilizing an in-plane technique. Again, nerve
stimulation may be utilized for confirmation of needle position and
catheter position. After anesthetizing the skin and subcutaneous tract, the
needle is introduced, and its tip is brought into position next to one of
the nerve targets. Some authors recommend that the catheter be placed
next to the posterior cord for optimal analgesia.6 Appropriate catheter
position may be confirmed by injecting saline, dextrose, or local
anesthetic solution while observing with ultrasonography. Ideally, the
solution will spread to encompass all three of the cords.

The medial or midinfraclavicular approach Some practitioners


prefer to perform an ultrasound-guided infraclavicular block at a
more proximal position, where the cords are grouped together
cephalad to the artery (Fig. 19.4, Fig. 19.5). This allows for access to
all three cords along one needle path. Using this technique, the onset
time is faster, the axillary nerve is more reliably anesthetized, and
the intercostal brachial nerve is often blocked as well when
compared to a more lateral approach.11 The drawback is that the
cords lie close to the first and second ribs so that pleural puncture is
more likely if the user does not have a thorough understanding of
the anatomy and good technical skills (Fig. 19.4).

The upper extremity is positioned in the same position used for axillary
block. This pulls the plexus cephalad and away from the rib cage and
pleura. At the same time, it rotates the plexus anteriorly so that it is
closer to the skin. The deltopectoral groove is palpated and the
transducer is sited with a sagittal orientation in the groove or 1 to 2 cm
medial to the groove (Fig. 19.5). Occasionally, toggling the transducer
medially may be necessary. The axillary artery is identified, and the
plexus can be seen cephaloanterior to the artery. A small curved
transducer (11 mm, 8 to 10 MHz) is very useful for this block because
there is very little space between the clavicle and transducer. The needle
(5 cm, 22G) is introduced in-line cephalad to the transducer at an angle
that is nearly perpendicular to the skin (Fig. 19.5). The needle is passed
through the skin and pectoral muscles while injecting local anesthetic to
anesthetize the skin and muscles. Once the clavipectoral fascia is
pierced, 5 to 10 mL of local anesthetic is injected into or around each of
the three cords. The plexus may appear as a single cord before injection,
but once injection is begun, the cords begin to separate.
If the practitioner plans to use a catheter, the same technique is used
except that a 5-or 10-cm, 18G, Tuohy needle is used to guide the
catheter to the plexus. In this case, the practitioner must anesthetize the
skin and muscles with local anesthetic using a small-gauge needle before
inserting the Tuohy needle.

Summary of evidence Ootaki et al.7 utilized a 7-MHz transducer for


infraclavicular block (ICB) with ultrasound guidance. The authors
reported a 95% success rate for surgical block of the arm with
sensory block of all nerves in more than 95% of patients and motor
block in over 90%. The technique involved injection of 15 mL of
1.5% lidocaine medial to the artery and 15 mL lateral to it with real-
time ultrasound guidance.
Sandhu and Capan6 reported on their experience of infraclavicular block
with ultrasound guidance in 126 patients undergoing hand surgery. The
authors utilized a 2-to 5-MHz transducer and injected 7 to 11 mL of a
solution of 2% lidocaine with epinephrine and sodium bicarbonate
around each of the three cords at the second part of the axillary artery,
deep to the pectoralis minor muscle. Ninety percent of patients
developed a surgical block, with three requiring conversion to general
anesthesia and seven necessitating local supplementation to complete the
block. One vascular puncture occurred. The mean onset time of the
blocks, which required an average time of 10 minutes to complete, was
6.3 minutes.
Klaasted et al.8 have described a “lateral and sagittal” technique for ICB,
in which the needle is inserted along the medial border of the coracoid
process, aiming inferior and posterior to contact the cords, in the
parasagittal plane. This approach, which is similar to that described in
this section, has been corroborated by Brull et al.9 These authors utilized
real-time sonographic needle guidance with this approach, consistently
contacting the cords at a depth of 4 to 6 cm. By report, this has resulted
in improved block success and reduced morbidity in their practice.
Dingemans et al.10 have characterized the necessary distribution of local
anesthetic solution during ICB with ultrasound guidance by randomizing
72 patients for hand or forearm surgery to one group with placement of
the solution in a U shape, posterior to and on both sides of the axillary
artery, or a second group in which local anesthetic was injected at one
site, guided by ultrasound imaging and detection of a distal motor
stimulation in the wrist or hand. Patients in the group with ultrasound
spread around the artery as the end point for the block required less time
for blockade (3.1 versus 5.2 minutes) and had a higher likelihood of
complete sensory block (86% versus 57%). In addition, successful
attainment of surgical block, with no requirement for supplementation,
was significantly higher in the group in which local anesthetic was
deposited around the artery, without guidance by nerve stimulation (92%
versus 72%).
References
1. Rettig HC, Gielen MJM, Boersma E, et al. Vertical infraclavicular
block of the brachial plexus: effects on hemidiaphragmatic
movement and ventilatory function. Reg Anesth Pain Med.
2005;30:529–535.
2. Sauter AR, Smith H-J, Stubhaug A, et al. Use of magnetic
resonance imaging to define the anatomic location closest to all
three cords of the infraclavicular brachial plexus. Anesth Analg.
2006;103:1574–1576.
3. Partridge BL, Katz J, Benirschke K. Functional anatomy of the
brachial plexus sheath: implications for anesthesia. Anesthesiology
1987;66:743–747.
4. Wilson JL, Brown DL, Wong GY, et al. Infraclavicular brachial
plexus block: parasagittal anatomy important to the coracoid
technique. Anesth Analg. 1998;87:870–873.
5. Perlas A, Chan VW, Simons M. Brachial plexus examination and
localization using ultrasound and electrical stimulation.
Anesthesiology. 2003;99:429–435.
6. Sandhu NS, Capan LM. Ultrasound-guided infraclavicular brachial
plexus block. Br J Anaesth. 2002;89:254–259.
7. Ootaki C, Hayashi H, Amano M. Ultrasound-guided infraclavicular
brachial plexus block: an alternative technique to anatomic
landmark-guided approaches. Reg Anesth Pain Med. 2000;25:600–
604.
8. Klaasted O, Smith HJ, Smedby O, et al. A novel infraclavicular
brachial plexus block: the lateral and sagittal technique, developed
by magnetic resonance imaging studies. Anesth Analg.
2004;98:252–256.
9. Brull R, McCartney CJL, Chan VW. A novel approach to the
infraclavicular brachial plexus block: the ultrasound experience.
Anesth Analg. 2004;99:950–952.
10. Dingemans E, Williams SR, Arcand G, et al. Neurostimulation in
ultrasound-guided infraclavicular block: a prospective, randomized
trial. Anesth Analg. 2007;104:1275–1280.
11. Bigeleisen PE, Wilson M. A comparison of two techniques of
ultrasound guided infraclavicular block. Br J Anaesth. 96:502–507.
20
Ultrasound-Guided
Axillary Block
STEVEN L. OREBAUGH, JASON D. HANKS,

AND PAUL E. BIGELEISEN

Background and indications: Traditionally, axillary block was the most


commonly performed of the brachial plexus blocks because of the ease
of locating the plexus relative to the axillary artery by palpation. It does
have some drawbacks. The arm must be abducted and the elbow flexed
to access the axilla. This can be painful for patients with fractures and
impossible for people with contractures or arthritis. In addition, it was
often difficult to anesthetize the musculocutaneous nerve with a blind or
stimulation technique. The addition of ultrasound allows the user to
identify all four nerves (median, radial, ulnar, musculocutaneous)
necessary for a successful block. That, coupled with the safety of the
block, make it an attractive procedure when ultrasound is used. The
block is used for surgery of the elbow, forearm, and hand. A separate
block of the intercostal brachial nerve is required when the incision is
along the medial aspect of the extremity.

Anatomy: The biceps muscle lies anterosuperior to the neurovascular


bundle, the coracobrachialis muscle is superior to the neurovascular
bundle, and the triceps muscle is inferior to the neurovascular bundle.
The humerus lies deep to the neurovascular bundle. The brachial artery
and one to two brachial veins are evident in the neurovascular bundle.
The radial, median, and ulnar nerves are found within the neurovascular
bundle (Figs. 20.1 and 20.2). Most commonly, the median nerve is
anterior or cephaloanterior to the artery. The radial nerve is most
commonly posterior or posteroinferior to the artery, the ulnar nerve is
most commonly found inferior or anteroinferior to the artery. Proximal
in the axilla, the musculocutaneous nerve may be found cephaloposterior
to the artery. In more distal sites in the axilla, the musculocutaneous
nerve is usually found in the fascia between the biceps and
coracobrachialis muscles 1 to 2 cm cephaloposterior to the artery.
Cutaneous nerves of the arm or forearm may also be visualized (Figs.
20.1 and 20.2).
Patient position: Supine, with ipsilateral arm abducted, externally
rotated, and flexed at the elbow.

Transducer: 25-or 38-mm linear transducer oscillating at 13 MHz.

Transducer position: Transverse across the axilla (sagittal oblique),


placed at the intersection of the pectoralis and biceps muscles.

Needle: 22G, 50-mm insulated needle.

Local anesthetic: 15 to 25 mL of 0.5% ropivacaine or 0.5% bupivacaine


with epinephrine.

Technique: The skin is cleansed with sterile solution and the transducer
is covered with a sterile cover. A wheel of local anesthetic should be
injected beneath the skin along a 5-cm arc from medial to lateral to the
brachial artery pulsation (Fig. 20.3). This allows needle placement from
either side of the artery without repeatedly injecting subcutaneously
local anesthetic as well as providing anesthesia for the intercostobrachial
nerve and the medial brachial cutaneous nerve. The artery should be
localized with the transducer and the hyperechoic nerves sought at its
periphery. Initially, the block needle is inserted in plane, along the long
axis of the transducer, from the superior side of the artery (Fig. 20.3). In
the posterocephalad region, the musculocutaneous nerve is sought. The
peripheral nerve stimulator may be left on throughout the procedure,
with a current level of 0.5 to 1 mA, or it may be switched on as each
nerve is approached, then turned off after confirmation. When elbow
flexion occurs, the nerve is localized. The stimulator can be switched
off, and incremental injections of 2 to 3 mL of local anesthetic are
begun. A “halo” of local anesthetic should be created around the nerve.
A total of 5 mL is injected here (Fig. 20.4).
The needle is then withdrawn and redirected toward the median nerve, if
evident, or to the region anterior and/or superior to the artery. Asking the
patient to flex and extend his or her arm causes the nerve to rotate back
and forth around the artery. The nerve stimulator may be left on
throughout the procedure or turned on at this time. Appropriate contact
of the stimulating needle with the nerve will cause flexion of the wrist
and/or a paresthesia in the third finger. Local anesthetic is then
incrementally injected (5 mL) until a halo appears around the nerve. The
needle is then directed to the ulnar nerve, if evident, or to the inferior
edge of the artery. When motor stimulation of the ulnar nerve occurs, the
fifth digit is flexed and the thumb is adducted. A paresthesia in the fifth
digit may be perceived. Five milliliters of local anesthetic is injected as
described earlier. Finally, the needle is redirected more posterior and
guided to the radial nerve. When the nerve is contacted with the
stimulating needle, extension of the wrist or elbow may occur as well as
a paresthesia in the thumb. Five milliliters of local anesthetic is then
injected incrementally following the procedure outlined earlier. In some
patients, it may be necessary to push the artery out of the way with the
needle in order to anesthetize all four nerves from the same entry point.
Some practitioners prefer to use a perivascular technique. In this
approach, the musculocutaneous nerve is anesthetized as described
earlier. The reminder of the local anesthetic, usually 20 to 25 mL, is
deposited posterior to the artery. Although easy to perform, the
technique requires a larger dose of local anesthetic and usually has a
longer onset time to complete block when compared to identifying and
anesthetizing each nerve individually.

Tips 1. Veins may vary in number, with one, two, or even more
being present. They are easily compressed, and care must be taken
to note their position, as even mild pressure with the transducer can
obliterate the lumen on the ultrasound image. Five percent to 10%
of patients will have an accessory axillary artery located deep or
posterior to the primary axillary artery (Fig. 20.5).
2. It is difficult to contact and anesthetize all four nerve blocks from one
needle insertion site due to the location of the nerves around the
circumference of the artery and the variable location of the
musculocutaneous nerve. Some practitioners prefer to block the
musculocutaneous and median nerves from a cephalad approach and
to block the ulnar and radial nerves by introducing the needle inferior
to the probe.
Suggested Readings
Bigeleisen P. The bifid axillary artery. J Clin Anesth. 2004;16:224–225.
Kovacs P, Gruber H, Bodner G. Interventional techniques. In: Peer S and Bodner G, eds. High-
Resolution Sonography of the Peripheral Nervous System. Berlin, Germany: Springer-Verlag;
2003:94–104.
Retzl G, Kapral S, Greher M, et al. Ultrasonographic findings of the axillary part of the brachial
plexus. Anesth Analg. 2001;92:1271–1275.
Schafhalter-Zoppoth I, Gray AT. The musculocutaneous nerve: ultrasound appearance for
peripheral nerve block. Reg Anesth Pain Med. 2005;30:385–390.
21
Ultrasound-Guided
Blocks at the Elbow
and Forearm
STEVEN L. OREBAUGH, SHRUTHIMA

THANGADA, AND PAUL E. BIGELEISEN

Background and indications: Proximal upper extremity blocks on


occasion fail to provide anesthesia in all of the nerve territories required
for surgery on the forearm or hand. When this occurs, anesthesiologists
may choose to provide a more distal block to supplement the brachial
plexus block, selectively placing local anesthetic around the nerve that
was not adequately anesthetized. Infrequently, respiratory compromise
from a pneumothorax or phrenic nerve block may be a complication of
supraclavicular or infraclavicular block. This complication can manifest
several hours after the procedure is performed. In selected outpatient
surgeries, one may consider a distal elbow block in patients susceptible
to respiratory complications. Selective block of more distal nerves under
ultrasound guidance may also be useful for limited surgical procedures
of the forearm or hand, such as open reduction and fixation of a finger,
carpal tunnel release, tendon repair, or ganglion cyst removal.1 Although
distal extremity blocks are an option in these cases, a recent study
demonstrated no significant difference in patient satisfaction when
comparing distal extremity to proximal brachial plexus blocks where
motor block above the elbow was present.2
Anatomy: With the aid of ultrasound, the ulnar nerve may be followed
from the axilla to the sulcus ulnaris—the bony canal—where it travels
through the ulnar groove between the medial epicondyle of the radius
and the olecranon process of the ulna. Most authors recommend
providing the block 5 cm proximal to the sulcus ulnaris (Fig. 21.1).
Here, the nerve is subcutaneous and amenable to block. Nearer to the
elbow, the nerve lies within the sulcus ulnaris (Fig. 21.2). Performing
the block within the sulcus ulnaris may lead to high pressures deep to the
retinaculum, which confines the nerve to the bony canal in most patients.
Injection at the level of the sulcus ulnaris may also lead to a higher
probability of intraneural injections, seeing as the nerve is less mobile.3
Immediately below the elbow and lying between the flexor digitorum
superficialis and the flexor carpi ulnaris, the nerve is also superficial and
easily blocked (Fig. 21.3). More distally, in the forearm, the ulnar nerve
courses between the flexor tendon layers. Nearby lie the ulnar artery and
median nerve.4,5 At the level of the wrist, the artery and nerve can be
seen passing distally into the hand, medial to the carpal tunnel (Fig.
21.4).
The median nerve remains with the brachial artery in its course through
the medial aspect of the arm to the elbow. At the elbow, it can be located
quite readily with ultrasonography and found to be medial to the brachial
artery (Fig. 21.5). As the median nerve travels distally into the forearm,
it initially remains close to the ulnar artery. In the forearm, the median
nerve is located between the flexor digitorum superficialis and flexor
digitorum profundus. It can be followed with ultrasound scanning into
the wrist, where it is seen in the carpal tunnel among the flexor tendons
(Fig. 21.6). Voluntary motion by the patient of his or her fingers causes
the tendons to “dance,” moving about quite actively while the nerve is
less mobile. This may help distinguish these structures from one another.
Using ultrasound imaging to follow the radial nerve from the axilla to
the elbow can be difficult, seeing as the nerve’s course requires one to
“spiral” the transducer around the humerus. However, in most patients,
the nerve is quite apparent as it emerges from the posterior aspect of the
humerus, piercing the lateral intermuscular septum and arriving at the
elbow on the lateral aspect of the arm (Fig. 21.7). In the antecubital
region, the nerve lies between the biceps tendon and the brachioradialis
muscle, lateral to the brachial artery. The nerve bifurcates at this point,
giving rise to the superficial radial nerve and the posterior interosseous
nerve (deep radial nerve). The former is quite small and difficult to trace
into the forearm (Fig. 21.8), whereas the latter is deep and provides
innervation to all the dorsal (extensor) forearm musculature.

Patient position: Supine with the arm supinated. For ulnar block above
the elbow, the arm is abducted and the elbow is flexed.
Probe: 25-mm linear probe oscillating at 13 MHz.

Probe orientation: Transverse.

Needle: 22G, 5-cm blunt needle.

Local anesthetic: 3 to 7 mL of 0.5% ropivacaine or 0.5% bupivacaine


with epinephrine (do not use epinephrine for wrist or digit blocks).

Technique: The risks and benefits of an elbow nerve block are first
explained to the patient, and then consent is obtained. After placing
appropriate monitors on the patient, supplemental oxygen is given via
nasal cannula. The operative arm or block site is marked by the
anesthesia team. An appropriate, preprocedural “time-out” is then
conducted to the correct surgical site. Sedation is given. An antiseptic is
utilized to prepare the skin of the area to be blocked.
After establishing an appropriate image of the ulnar or median nerve at
the elbow or forearm or the radial nerve just proximal to the elbow, the
physician may choose either an in-plane or out-of-plane approach to
nerve blockade. The in-plane technique allows closer observation of the
tip of the needle and facilitates direct deposit of local anesthetic on the
posterior side of the nerve. Depending on which technique is chosen, the
skin is infiltrated by the block needle at the appropriate site for the
introduction of the local anesthetic.
After an approach is chosen, the block needle is introduced. Peripheral
nerve stimulation may be elected to guide the block jointly with
ultrasound imaging. After the needle approximates the nerve, with or
without nerve stimulation, local anesthetic is injected in small aliquots to
surround the nerve being blocked. Care must be taken when blocking
nerves adjacent to small vessels, such as the ulnar artery, to avoid
intravascular injection. Small arteries are difficult to visualize and may
be compressed when subjected to pressure applied with the probe.

Complications: When compared to proximal upper extremity blocks,


there is no increased incidence of local anesthetic toxicity, vascular
puncture, or peripheral nerve damage in blocks at the elbow and
forearm.
Summary of evidence: Evidence supporting distal nerve block in the
upper extremity with ultrasound guidance is relatively scant and
primarily consists of observational studies or small case series.
Using the technique summarized earlier, Gray and Schafhalter-Zoppoth4
described block of the ulnar nerve in the forearm for surgery on the fifth
finger. Gebhard et al.6 performed a retrospective study on 62 patients
undergoing carpal tunnel release surgery, comparing various anesthetic
techniques in regards to cardiovascular stability and discharge time.
Techniques included Bier block, distal nerve blocks, and general
anesthesia. Patients who received distal nerve blocks had less cardiac
instability and decreased hospital stay than those who received Bier
blocks or general anesthesia. Klezl et al.7 reported an 825-patient series
study in which single distal nerve block failed more frequently than
complete distal nerve block (18% rate of failure versus 2% rate of
failure). They concluded that complete wrist block with the ulnar,
median, and radial nerve is more efficacious as a result of cutaneous
nerves being blocked as well. Ponrouch et al.8 performed a randomized,
double-blinded study comparing the minimum effective anesthetic
volumes (MEAV) for median and ulnar nerve blocks utilizing ultrasound
or nerve stimulation guidance for carpal tunnel release surgery.
Mepivacaine 1.5% was used as the local anesthetic for the blocks, and
sensory block, block onset, and duration were examined. Ponrouch et
al.8 concluded that there was a 50% reduction in the MEAV for
ultrasound-guided median nerve block. They also noted that decreasing
the volume of local anesthetic may decrease sensory block duration but
not onset time.
References
1. Neal JM, Hebl JR, Gerancher JC, et al. Brachial plexus anesthesia:
essentials of our current understanding. Reg Anesth Pain Med.
2002;27:402–428.
2. Fredrickson MJ, Wolstencroft PJ, Chinchanwala S, et al. Does
motor block related to long-acting brachial plexus block cause
patient dissatisfaction after minor wrist and hand surgery? A
randomized observer-blinded trial. Br J Anaesth. 2012;109:809–
815.
3. Chelly JE. Blocks at the elbow. In: Chelly JE, ed. Peripheral Nerve
Blocks—A Color Atlas. 2nd ed. Philadelphia, PA: Lippincott
Williams Wilkins; 2004:58–60.
4. Gray AT, Schafhalter-Zoppoth I. Ultrasound guidance for ulnar
nerve block in the forearm. Reg Anesth Pain Med. 2003;28:335–
339.
5. McCartney CJL, Xu D, Constantinescu C, et al. Ultrasound
examination of peripheral nerves in the forearm. Reg Anesth Pain
Med. 2007;32:434–439.
6. Gebhard RE, Al-Samsam T, Greger J, et al. Distal nerve blocks at
the wrist for outpatient carpal tunnel surgery offer intraoperative
cardiovascular stability and reduce discharge time. Anesth Analg.
2002;95:351–355.
7. Klezl Z, Krejca M, Simcik J. Role of sensory innervations
variations for wrist block anesthesia. Arch Med Res. 2001;32:155–
158.
8. Ponrouch M, Bouic N, Bringuier S. Estimation and
pharmacodynamic consequences of the minimum effective
anesthetic volumes for median and ulnar nerve blocks: a
randomized, double-blind, controlled comparison between
ultrasound and nerve stimulation guidance. Anesth Analg.
2010;111:1059–1064.
Section III
Lower Extremity
22 Ultrasound-Guided Lumbar Plexus Block
(Transverse Approach)

23 Ultrasound-Guided Femoral Nerve Block

24 Ultrasound-Guided Lateral Femoral Cutaneous


Block

25 Ultrasound-Guided Saphenous Nerve Block of the


Thigh, Knee, and Ankle
Part 1: Saphenous Nerve Block of the Thigh
Part 2: Ultrasound-Guided Saphenous Nerve Block
Below the Knee (Infrapatellar) and Ankle

26 Ultrasound-Guided Proximal Obturator Block: A


Proximal Interfascial Technique

27 Ultrasound-Guided Distal Obturator Nerve Block

28 Ultrasound-Guided Proximal Parasacral Block

29 Ultrasound-Guided Distal Parasacral Block

30 Ultrasound-Guided Anterior Sciatic Nerve Block

31 Ultrasound-Guided Subgluteal Sciatic Block

32 Ultrasound-Guided Popliteal and Lateral Sciatic


Block
33 Ultrasound-Guided Ankle Block


22
Ultrasound-Guided
Lumbar Plexus Block
(Transverse
Approach)
SHINICHI SAKURA, KAORU HARA, AND

JEAN-LOUIS HORN

Introduction and indications: The lumbar plexus consists of roots L1


through L4. In some patients, roots T12 and L5 contribute. The roots
join together to form the subcostal, iliohypogastric, ilioinguinal, lateral
femoral cutaneous, genital-femoral and obturator nerves. The lumbar
plexus usually lies within the posterior substance of the psoas muscle.
The lateral femoral cutaneous and femoral nerves usually lie within the
same fascial plane, but the obturator nerve often lies within a separate
fold of the muscle. Lumbar plexus block is used to provide
intraoperative and postoperative analgesia for lower extremity surgery,
sometimes in combination with sciatic nerve block. It may also be used
as a series of injections for patients with chronic lower extremity pain.
Anatomy: The spinous process, the articular process, and the transverse
process are important landmarks. These bony structures are hyperechoic
and create a shadow. When the transducer is moved slightly cephalad or
caudad, a clear image of the psoas and quadratus lumborum muscles is
obtained. The lumbar plexus generally is expected to lie within the
posterior third of the psoas major muscle and 2 to 3 cm posterior to the
anterior surface of the transverse process. The lumbar plexus is not
always visualized under ultrasound but may be observed as a
hyperechoic structure in a young population (Fig. 22.1).
Patient position: Lateral decubitus position with the side to be blocked
uppermost.
Transducer: 40 to 60 mm curved array oscillating at 2 to 5 MHz.
Transducer position: The transducer is positioned transversely in the
midline of the back at the L4 level to capture the spinous process. The
transducer is then moved laterally (approximately 3 cm) while scanning
the paravertebral region until a clear image of the articular and
transverse processes is obtained.

Needle: A 100 to 150 mm, short-bevel, 21G insulated needle.


Local anesthetic: 20 to 35 mL of ropivacaine 0.5% to 0.75%. For pain
therapy, 10 mL of ropivacaine 0.25%.
Approach and technique: The skin is washed with sterile solution, and
the transducer is covered with a sterile sheath. Ultrasound-guided
lumbar plexus block requires advanced skill because of the depth of
needle placement. After the optimal transducer position is found, the
transducer is angled laterally or medially in the transverse plane,
depending on the needle insertion site, while the target is kept in the
middle of the ultrasound image. This allows the practitioner to view the
needle more easily.
Lateral Insertion Approach: After skin infiltration with local anesthetic,
the needle is inserted in plane from the lateral edge of the transducer.
The transducer is perpendicular to the skin (Fig. 22.2). The needle
insertion site should be around 5 cm lateral to the midline, and the
needle direction should be perpendicular to the body plane or directed
slightly toward the midline. When the lumbar plexus is visualized, the
needle should be advanced directly to the plexus. If the lumbar plexus is
not visualized, the location of the target can be estimated from the
distance of the skin to the transverse process. In either case, electrical
stimulation helps identify and confirm the target. When the needle tip is
near the target, a nerve stimulator with pulse duration of 0.1 ms and
stimulating frequency of 2 Hz is turned on to elicit twitches of the
quadriceps muscle. After contractions are obtained by stimulation
between 0.5 and 1.0 mA, local anesthetic is injected in small increments
after negative aspiration. This allows the practitioner to test for an
intravascular or intrathecal position of the needle. During injection, local
anesthetic and tissue expansion is observed within the psoas muscle and
may surround the lumbar plexus (Fig. 22.3).
Medial Approach: The needle can be inserted in plane from the medial
edge of the transducer. In this case, the transducer should be angled
slightly toward the midline (Fig. 22.4). The needle insertion site should
be 3 to 4 cm lateral to the midline with the needle direction
perpendicular to the body. In some cases, it may be necessary to angle
the needle slightly laterally (Fig. 22.5). The remainder of the procedure
is the same as stated previously.
Sagittal Approach: Some practitioners prefer to use a sagittal orientation
of the probe. In this approach, the iliac crest and posterior superior iliac
spine are identified and the probe is sited in a sagittal plane at the level
of L4 approximately 4 to 5 cm lateral to the midline (Fig. 22.6). The
hyperechoic shadows of L3, L4, and L5 are identified (Fig. 22.6). A
stimulating needle set at 1 mA is inserted either in plane or out of plane
in the space between L3 and L4. The needle is advanced between the
spinous processes of L3 and L4 until a contraction is elicited in the
quadriceps muscle. Local anesthetic (15 to 25 mL) is then injected in
aliquots of 3 to 5 mL. Because of the depth of the plexus in obese
patients, some practitioners simply use ultrasound to determine the depth
and location of the transverse processes and then proceed with a blind
technique using electrolocation, a paresthesia, or a loss of resistance
technique.
Tips 1. Because of the depth of the target and the angle of the needle, it is
difficult to observe the entire needle during the whole procedure.
However, the needle tip should be observed or at least estimated tissue
movement when the needle is moved under real-time ultrasound
imaging. The use of an echogenic tip needle is preferable.
2. To obtain a better needle image under ultrasound, the transducer
should be angled medially or laterally so that the angle of the needle
in the ultrasound beam is less steep. This will allow visualization of
the needle tip and shaft more easily. Needle insertion should not be
too lateral to avoid kidney puncture nor should it be too medial to
avoid subarachnoid injection.
Continuous lumbar plexus block is useful for the management of
postoperative pain after major lower extremity surgery. The technique
is similar to a single-shot injection, but the use of a Tuohy needle is
recommended. A catheter is inserted approximately 5 to 10 cm
beyond the needle tip after administration of local anesthetic.
3. In some patients, the iliac artery can be visualized anterior to the
psoas muscle.
Suggested Readings
Awad I, Dugan E. Posterior lumbar plexus block: anatomy, approaches, and techniques. Reg
Anesth Pain Med. 2005;30:143–149.
Kirchmair L, Entner T, Kapral S, et al. Ultrasound guidance for the psoas compartment block: an
imaging study. Anesth Analg. 2002;94:706–710.
Kirchmair L, Enna B, Mitterschiffthaler G, et al. Lumbar plexus in children: a sonographic study
and its relevance to pediatric regional anesthesia. Anesthesiology. 2004;101:445–450.
Robards C, Hadzic A. Lumbar plexus block. In: Hadzic A, ed. Textbook of Regional Anesthesia
and Acute Pain Management. New York, NY: McGraw-Hill Medical; 2007:481–488.
Karmakar MK, Ho AM, Li X, et al. Ultrasound-guided lumbar plexus block through the acoustic
window of the lumbar ultrasound trident. Br J Anaesth. 2008;100:533–537.
23
Ultrasound-Guided
Femoral Nerve Block
STEPHEN M. BRENEMAN

Background and indications: Femoral nerve block has traditionally


been used in conjunction with block of the sciatic nerve for surgery of
the knee, leg, or foot, as well as saphenous vein stripping. In the past,
practitioners used a blind approach (fascia iliaca), which relied on
palpating the iliac crest and the pubic ramus. Other practitioners used
electrolocation or paresthesia, or they simply infiltrated around the
femoral artery after palpating a pulse. Because of the variable
distribution and location of the femoral nerve, ultrasound has made the
procedure more reliable.
Anatomy: The femoral nerve is formed from roots L2, L3, and L4. These
roots combine to form the femoral nerve, which travels in the space
between the psoas and iliacus muscles (Fig. 23.1A). In most cases, the
nerve actually travels within the substance of the psoas muscle (Fig.
23.1B). The nerve emerges below the inguinal ligament, lateral to the
femoral artery and deep to the fascia iliaca. Many textbooks portray the
nerve as a single round or oval structure. In fact, the nerve is often flat,
giving rise to its many branches that subserve the thigh.
Patient position: Supine or recumbent with groin exposed.
Transducer: 8 to 12 MHz linear array or 5 to 8 MHz curved array.
Transducer position: The transducer is placed axially anywhere between
the inguinal crease to the inguinal ligament (Fig. 23.2A).
Needle: 5-cm, 22G nerve block needle.
Local anesthetic: 10 (saphenous), 20 (femoral), or 20 to 40 mL (fascia
iliaca).
Technique: The groin is exposed and examined for skin breakdown or
infection. The skin is washed with sterile solution. The inguinal crease,
anterior superior iliac spine, and lateral superior pubic ramus are
identified. The transducer is placed axially anywhere between the
inguinal crease to the inguinal ligament. The femoral artery and vein are
identified, as are the curved surface of the iliopsoas muscle and the
hyperechoic triangle formed between the muscle and the artery, which
are covered by the fascia iliaca (Fig. 23.2A,B). This triangle can have a
mirror image on the medial aspect of the femoral vein (Fig. 23.2C).
The needle entry is lateral to the probe using an in-line approach A 30-to
45-degree needle trajectory should pierce the surface of the iliopsoas
muscle until a distinct “pop” of the fascia iliaca is seen as the triangle is
entered. (The fascia lata “pop” is often missed visually on ultrasound.)
The triangle is filled, and most of the time, the hyperechoic nerve will
show more clearly once it has been surrounded with local anesthetic.
Tips 1. When using a nerve stimulator, set the current to 0.8 to 1.0 mA,
which is used for verification rather than localization. This lower
amperage is especially important in young patients with significant
muscle development.
2. The most common initial contraction is the sartorius muscle over the
medial thigh. This represents the anterior branch of the femoral nerve
(Fig. 23.3), which is deep to the fascia lata but superficial to the
fascia iliaca. Thus, an injection in this superficial plane will not lead
to adequate anesthesia of the thigh and knee. Usually, moving the
needle more laterally and deeper (and “pop”) will lead to a
contraction of the quadriceps muscle and anterior movement of the
patella (Fig. 23.4). With very small needle movements, the
contraction may move back and forth from the medial thigh to the
anterior and lateral thigh. If the stimulating needle is too deep, there
will be a local twitch of the iliopsoas muscle, which does not extend
to the patella.
3. Unlike nerves of the brachial plexus or the sciatic nerve, the femoral
nerve is not always apparent prior to injection (Fig. 23.5). Also,
hyperechoic areas do not always indicate the nerve (Fig. 23.6).
Because of the difficulty in visualization, often the injection is
intraneural (Fig. 23.7); thus, keep injection pressures low.
4. For obese patients, one may need to tape the pannus up and away
from the groin. Beware of yeast infections in this location. Also, use a
longer needle and start further lateral from the probe so that the
needle angle does not become too steep. A needle angle greater than
30 to 45 degrees can make the block needle even harder to visualize.
5. This block has other names including “3-in-1” and “fascia iliaca”
block. The 3-in-1 described by Winnie included coverage for the
obturator nerve. His results have not been replicated. Thus, the
femoral block could be considered a 2-in-1 block, which includes the
lateral femoral cutaneous nerve when higher volumes of local
anesthetic are administered. The fascia iliaca block was a “block by
feel.” The practitioner placed a finger on the artery. A blunt bevel
needle was sited several centimeters lateral to the pulse place with the
needle oriented perpendicular to the skin. The needle was inserted
through the skin, and then the practitioner advanced the needle until
two additional pops were felt. After this, 20 to 40 mL of local
anesthetic was injected.
6. For ultrasound-guided catheter placement next to the femoral nerve,
use the same lateral approach with an 18G Tuohy needle. Once the tip
of the needle is properly sited, inject local anesthetic through the
needle to dilate the area around the nerve. This also allows
visualization of the Tuohy needle tip adjacent to the nerve. Next,
direct the bevel cranially and pass the catheter. An out-of-plane
method can be used, as well, with the subsequent injection of local
via the catheter. When using a stimulating catheter, D5W can be used
to expand the space to allow easier passage of the stimulating catheter
while maintaining a contraction in the muscle. The catheter can be
threaded cranially to the lumbar plexus region.
Suggested Readings
Marhofer P, Schrogendorfer K, Wallner T, et al. Ultrasonographic guidance reduces the amount of
local anesthetic for 3-in-1 blocks. Reg Anesth Pain Med. 1998;23:584–588.
Tsui BC, Dillane D, Pillay J, et al. Cadaveric ultrasound imaging for training in ultrasound-guided
peripheral nerve blocks: lower extremity. Can J Anaesth. 2007;54:475–480.
Casati A, Baciarello M, Di Cianni S, et al. Effects of ultrasound guidance on the minimum
effective anaesthetic volume required to block the femoral nerve. Br J Anaesth. 2007;98:823–
827.
Sites BD, Beach M, Gallagher JD, et al. A single injection ultrasound-assisted femoral nerve block
provides side effect-sparing analgesia when compared with intrathecal morphine in patients
undergoing total knee arthroplasty. Anesth Analg. 2004;99:1539–1543.
24
Ultrasound-Guided
Lateral Femoral
Cutaneous Block
PAUL E. BIGELEISEN, MILENA MORENO, AND

STEVEN L. OREBAUGH

Background and indications: Lateral femoral cutaneous nerve block is


useful when sensory anesthesia of the lateral thigh is necessary. In most
cases, this block is used in conjunction with femoral nerve block or
when a “3-in-1” block does not provide adequate anesthesia of the
lateral thigh. The block is also useful in the diagnosis and treatment of
meralgia paresthetica.
Anatomy: The lateral femoral cutaneous nerve is a branch of the brachial
plexus. The nerve arises from the L2 and L3 nerve roots. After exiting
the intervertebral foramina at these levels, the roots join the plexus and
the branches combine to form the nerve, which travels in the space
between the quadratus lumborum and psoas major muscles (Fig. 24.1). It
then passes under the inguinal ligament and over the sartorius muscle
into the thigh, where it divides into an anterior and posterior branch. The
anterior branch becomes superficial about 10 cm below the inguinal
ligament and supplies the skin of the anterior and lateral parts of the
thigh as far as the knee. The posterior branch pierces the fascia lata and
passes backward across the lateral and posterior surfaces of the thigh.
After the nerve passes medial to the sartorius muscle and distal to the
inguinal ligament, it can usually be imaged 2 cm medial and 5 cm
inferior to the anterior-superior iliac spine. The nerve appears as a
hyperechoic linear streak deep to the fascia iliaca (Fig. 24.2) at a depth
of 1 to 3 cm. If the nerve is imaged lateral to the femoral artery and
nerve, it can be found deep to the fascia iliaca or in the layer between the
fascia iliaca and the fascia lata (Fig. 24.3).
Patient position: The patient lies in the supine position for this block.
Transducer: A linear transducer, such as 5 to 10 MHz or 6 to 13 MHz.
Transducer orientation: Transverse or axial oblique position (Fig. 24.4).
Needle: 50-mm, 22G, short-bevel needle.
Local anesthetic: 5 mL of ropivacaine (0.2%) or 5 mL of bupivacaine
(0.25%).
Technique: Patient monitors are attached and sedation is administered.
Oxygen by face mask is initiated. Appropriate markings are placed on
the indicated extremity for confirmation and a brief preprocedural “time-
out” is conducted at the bedside. An antiseptic is utilized to prepare the
skin over the block area.
The probe is placed on the skin medial and inferior to the anterior-
superior iliac spine. The sartorius muscle is seen along the lateral aspect
of the probe in this position, and the fascia lata and iliaca are identified
(Figs. 24.2 and 24.3). Using an in-plane approach, the needle can be
placed either medial or lateral to the transducer. After anesthetizing the
skin, the needle is advanced through the fascia iliaca until it is adjacent
to the nerve. Local anesthetic solution (5 to 10 mL) are injected in
incremental doses. Injection of large volumes of local anesthetic in this
location will also anesthetize the femoral nerve.
Some practitioners prefer to place the probe in a transverse orientation in
the inguinal crease over the femoral artery. The fascial iliaca is identified
and then the probe is moved laterally. The lateral femoral cutaneous
nerve is identified medial to the iliac crest. In this location, the nerve or
its anterior and posterior branches are usually found in the layer between
the fascia lata and fascia iliaca. Once the nerve is identified, anesthesia
of the nerve is accomplished by infiltrating deep to the fascia lata.
Summary of evidence Ng et al. have shown that the nerve lies approximately 2
cm medial and 5 cm distal to the anterior-superior iliac spine. In a comparative
study carried out by Shannon et al., use of ultrasound resulted in a 100% success
rate for lateral femoral cutaneous nerve block in volunteers compared to a 40%
success rate using a blind technique.

Suggested Readings Ng I, Vaghadia H, Choi PT, et al. Ultrasound


imaging accurately identifies the lateral femoral cutaneous nerve.
Anesth Analg. 2008;107:1070–1074.
Shannon J, Lang SA, Yip RW, et al. Lateral femoral cutaneous nerve block revisited. A nerve
stimulator technique. Reg Anesth. 1995;20:100–104.
25
Ultrasound-Guided
Saphenous Nerve
Block of the Thigh,
Knee, and Ankle
Part 1: Saphenous Nerve Block of the
Thigh

STEVEN L. OREBAUGH, VIHANG D. SHAH,



MILENA MORENO, AND PAUL E. BIGELEISEN

Background and indications: Block of the saphenous nerve is useful for


anesthesia and postoperative analgesia for foot and ankle procedures
involving the medial dermatomes as well as surgery involving the
medial aspect of the leg. Saphenous nerve block has been described
using different approaches. Techniques for guiding the block have
included sensory nerve stimulation, motor nerve stimulation (of adjacent
femoral nerve branches), and surface landmarks alone. Described
approaches include the perifemoral, subsartorial/adductor canal, medial
femoral condylar, paravenous tibial condylar, and medial
malleolar/cutaneous.1,2 Ultrasonography allows increased precision for
this block.
Anatomy: The saphenous nerve may be approached with ultrasound
guidance at several different levels. The nerve may be difficult to locate
distally because of its small size, and therefore, the block relies on
identification of nearby anatomy.
After the arborization of the femoral nerve at the femoral crease, the
saphenous nerve runs distally with the femoral artery and vein, and,
eventually joins the saphenous vein in the lower leg (Figs. 25.1 to 25.6).
Soon after leaving the femoral triangle, the nerve, artery, and vein dive
deep through the adductor canal. Initially, the saphenous nerve is lateral
to the artery. In the distal part of the adductor canal, just proximal to the
adductor magnus muscle, the saphenous nerve traverses anteriorly over
the femoral artery to then run along the medial side of the artery. The
adductor canal is bounded anteriorly by the sartorius muscle,
anterolaterally by the vastus medialis, and posteromedially by the
adductor longus muscle.3 The nerve may be blocked at any point along
this path by placing local anesthetic perivascularly or next to the nerve if
it is visualized (Fig. 25.1). Distal to the adductor canal, the saphenous
nerve resides below the sartorius and soon passes through the
vastoadductor membrane, the fascial plane just deep to the sartorius
muscle, and adjacent to the vastus medialis (Fig. 25.2). At this point, the
nerve is ensconced in the connective tissue septum and runs with a small
artery and vein, which may be visible on ultrasound with the color
Doppler feature. The nerve, artery, and vein within this plane have been
described as a “string of beads.”4 However, the individual structures are
small and may require a very high-resolution transducer to distinguish.
Local anesthetic may also be injected at this level to block the saphenous
nerve. As the femoral artery and vein descend through the adductor
hiatus to become the popliteal vessels, the saphenous nerve courses
distally beneath the sartorius muscle. It passes medial to the knee joint,3
after which it runs in close juxtaposition to the saphenous vein, which
may be identified on ultrasound to help guide the blockade of the nerve.
In this position it is usually referred to as the infrapatellar nerve (IPN).
Patient position: Supine with the thigh and leg externally rotated.
Transducer: 25-mm linear transducer oscillating at 13 MHz (ankle or
tibia) or 11-mm curved transducer oscillating at 6 to 10 MHz (thigh).
Transducer position: Transverse.

Needle: 22G, 5-or 10-cm blunt needle; 18G, 5-or 10-cm Tuohy needle
(continuous blocks).
Local anesthetic: 5 to 10 mL of 0.2% ropivacaine or 0.25% bupivacaine
with epinephrine.
Technique: Supplemental oxygen and monitors are applied to the patient.
An appropriate marking is placed on the indicated extremity for
confirmation, and a preprocedural “time-out” is performed at bedside.
Mild sedation is administered to the patient. An antiseptic is utilized to
prepare the skin over the block area. Peripheral nerve stimulation is not
typically used for this block because there is no motor component.
For the perivenous approach to the saphenous nerve in the upper leg, one
should identify the saphenous vein and trace it to the level of the tibial
tubercle. Lower approaches may allow branching to occur above the
level of the block, resulting in incomplete anesthesia of the nerve. Once
the vein is identified, the skin is anesthetized with 0.5 to 1 mL of
lidocaine 1%, and the block needle is inserted through this site. Either an
in-plane or out-of-plane technique is possible for any of the saphenous
nerve block techniques. One advantage of the out-of-plane approach is
that only one injection site is used because the needle’s angle of
insertion can be changed using the same entry point. In either case,
ultrasound guidance is used to guide the tip of the needle to a perivenous
position (Fig. 25.1), and 2 to 3 mL of local anesthetic is injected on each
side. Completely surrounding the vein with local anesthetic is desirable
because the nerve is frequently not well distinguished. Two injections
may be required to accomplish this.
In the midthigh approach, the ultrasound probe is placed over the
femoral artery at the midthigh, and the vessels are followed distally
toward the adductor canal. The saphenous nerve is contiguous to the
artery and usually visible on ultrasound (Fig. 25.1). Just distal to the
entry of the femoral vessels into the adductor canal, the vastoadductor
membrane is visible as a fascial plane deep to the sartorius and along the
medial edge of the vastus medialis. More distally, this membrane
contains both the saphenous nerve and an accompanying small vein and
artery, giving the appearance of a string of beads.4 Injecting the
membrane with 6 to 8 mL of local anesthetic at this point results in
anesthesia of the saphenous nerve (Fig. 25.2). It should be noted that the
needle may traverse the sartorius muscle or the vastus medialis via this
approach (in-plane approach), depending on whether the needle is
inserted medial or lateral to the transducer. Some practitioners have used
this site for continuous catheter infusions. Other practitioners prefer to
place a catheter next to the femoral nerve in the inguinal region.
Summary of evidence: In 2003, Gray and Collins5 described the
approach to perivenous block of the saphenous nerve in the calf. The
authors at that time described the utility of the block as “uniformly
successful.” However, in 2007, Krombach and Gray4 noted that the
small size of the nerve at this level frustrates attempts at imaging, as
does the presence of multiple veins in many patients. In addition, they
pointed out that significant branching of the nerve may have occurred
proximal to this point, rendering the block less effective. Instead, the
authors described a block in the distal thigh, usually 5 to 7 cm proximal
to the flexion crease of the knee. The authors noted this approach to be
more reliable because of the more consistent course of the saphenous
nerve in the thigh and its larger size, rendering it more amenable to
imaging, at least with the 14-MHz linear transducer that they used. The
authors had, by this time, evaluated the block’s effect in 20 patients with
success. Manickam et al.6 demonstrated block success for all 20 patients
when approaching the saphenous nerve within the adductor canal 10 to
14 cm proximal to the popliteal crease. Tsui and Ozelsel7 also advocate
approaching the saphenous nerve at 10 to 12 cm proximal to the
popliteal crease. In 2010, Tsai et al.8 reported a 70% block success rate,
but they attributed this lower rate (as compared to Manickam et al.6) to
the fact that inexperienced residents performed the blocks and that this
retrospective review was unable to discern if block failure was truly
saphenous versus popliteal fossa block failure. In 2011, Saranteas et al.9
noticed that the saphenous nerve consistently ran between the sartorius
muscle and femoral artery after exiting the adductor canal in a group of
cadaver dissections, and they also demonstrated that ultrasound-guided
blockade as this site is very effective clinically.
Emerging literature suggests that continuous saphenous nerve blockade
within the adductor canal may be a suitable modality for postoperative
pain control after total knee arthroplasty.10,11 The distal position of
adductor canal blockade makes quadriceps weakness less likely as
compared to femoral nerve blockade12 and may optimize postoperative
rehabilitation in these cases. Further investigation is needed in this area.
Complications: Complications related to saphenous nerve blockade are
unusual. Small doses of local anesthetic make the likelihood of local
anesthetic systemic toxicity lower than in other types of peripheral nerve
blockade.

References 1. Benzon HT, Sharma S, Calimaran A. Comparison of


the different approaches to the saphenous nerve block.
Anesthesiology. 2005;102:633–638.
2. De May JC, Deruyck LJ, Cammu G, et al. A paravenous approach
for the saphenous nerve block. Reg Anesth Pain Med. 2001;26:504–
506.
3. Dunaway DJ, Steensen RN, Wiand W, et al. The sartorial branch of
the saphenous nerve: its anatomy at the joint line of the knee.
Arthroscopy. 2005;21:547–551.
4. Krombach J, Gray AT. Sonography for saphenous nerve block near
the adductor canal (letter). Reg Anesth Pain Med. 2007;32:369–
370.
5. Gray AT, Collins AB. Ultrasound-guided saphenous nerve block.
Reg Anesth Pain Med. 2003;28:148.
6. Manickam B, Perlas A, Duggan E, et al. Feasibility and efficacy of
ultrasound-guided block of the saphenous nerve in the adductor
canal. Reg Anesth Pain Med. 2009;34:578–580.
7. Tsui BC, Ozelsel T. Ultrasound-guided transartorial perifermoral
artery approach for saphenous nerve block. Reg Anesth Pain Med.
2009;34:177–178.
8. Tsai PB, Karnwal A, Kakazu C, et al. Efficacy of an ultrasound-
guided subsartorial approach to saphenous nerve block: a case
series. Can J Anaesth. 2010;57:683–688.
9. Saranteas T, Anagnostis G, Paraskeuopoulos T, et al. Anatomy and
clinical implications of the ultrasound-guided subsartorial
saphenous nerve block. Reg Anesth Pain Med. 2011;36:399–402.
10. Lund J, Jenstrup MT, Jaeger P, et al. Continuous adductor-canal-
blockade for adjuvant postoperative analgesia after major knee
surgery: preliminary results. Acta Anaesthesiol Scand. 2011;55:14–
19.
11. Jenstrup MT, Jaeger P, Lund J, et al. Effects of adductor-canal-
blockade on pain and ambulation after total knee arthroplasty: a
randomized study. Acta Anaesthesiol Scand. 2012;56:357–364.
12. Jaeger P, Nielsen ZJK, Henningsen MH, et al. Adductor-canal-
blockade versus femoral nerve block and quadriceps strength: a
randomized, double-blind, placebo-controlled, crossover study in
healthy volunteers. Anesthesiology. 2013; 118:409–415.
Part 2: Ultrasound-Guided
Saphenous Nerve of the Knee
(Infrapatellar) and Ankle

DIANA L. BESLEAGA, BASSEM ASAAD, ALEX


ROSIOREANU, STEPHEN M. BRENEMAN, AND
JAMES J. NICHOLSON

Background: The IPN is a purely sensory branch of the saphenous


nerve.1 It innervates the anteromedial and anterior inferior part of the
knee joint capsule.1 Damage to this nerve occurs frequently during knee
arthroplasty, arthroscopy, intramedullary nailing, and trauma. According
to Bademkiran et al.,2 injury to the IPN can occur in up to 75% of
arthroscopic or open surgical procedures involving the medial knee. The
incidence of altered sensation due to transection of the infrapatellar
branch of the saphenous nerve can be up to 70% at 1 year following total
knee arthroplasty.3 Commonly, patients with IPN neuropathy walk “stiff
legged” to avoid flexion at the knee.4 The joint may be red, warm, and
swollen, as though infected, with allodynia and pain on movement.4
There may be visible swelling in the medial tibial fossa, and palpation of
this area usually replicates pain that can lead to hypoesthesia,
paresthesia, dysesthesia, neuropathic pain, and complex regional pain
syndrome.4–6
Anatomy: The IPN branch of the saphenous nerve is a purely sensory
nerve that innervates the anterior aspect of the knee and anterior knee
capsule, providing cutaneous sensation to the medial aspect of the knee,
lower leg, and ankle.4,7 The saphenous nerve, composed of fibers from
the L3 and L4 nerve roots, originates from the posterior division of the
femoral nerve and gives rise to infrapatellar branch of saphenous nerve
after passing through the adductor (Hunter’s) canal.2,4 After the
saphenous nerve leaves the adductor canal, it divides into the
infrapatellar branch, which provides a sensory branch to the peripatellar
plexus of the knee and the sartorial branch, which perforates the
superficial fascia between the gracilis and sartorius muscles and emerges
to lie in the subcutaneous tissue below the knee fold8 (Figs. 25.3 to
25.5).
Knowledge of the IPN anatomy with its variable branches and paths is
helpful to decrease the risk of iatrogenic injury during surgery. When
performing knee arthroscopy, a safe zone for blind portal placement has
been outlined via use of a horizontal puncture site parallel to the
articular surface with the knee flexed 90 degrees.9 The safe zone is
within an approximate 30-mm area from the medial margin of the
midpatella and within an approximate 10-mm area from the medial
margin of the patellar ligament at the level of the distal pole of the
patella. Anterior cruciate ligament reconstruction donor harvest sites
have also been highly associated with IPN neuropathy. For bone-
patellar-bone harvesting, Beaufilis et al.10 recommended a dual-incision
technique that may spare the subcutaneous tissue where the nerve
crosses the knee. If a hamstring graft is being used, oblique incisions
have been shown to be superior to vertical-based harvest incisions with
respect to nerve injury.11 Total knee arthroplasty exposure requires a
medial parapatellar incision to the tibial tubercle, thereby leading to
disruption of the crossing IPN. In these cases, the IPN should be
identified, cauterized, and not put under stretch prior to being cut.
While working up a patient for a potential IPN neuroma, it is important
to eliminate a deep cause of his or her pain. It is not uncommon for a
patient with functional pain within a joint to develop neighboring
superficial neuritis. Therefore, the workup and evaluation must rule out
implant loosening, persistent ligamentous instability, and infection.12–15
Both intra-articular and perineural diagnostic lidocaine injections are
helpful in determining the real source of pain.
The majority of the outcome studies on treatment of IPN neuromas
involves open surgical selective partial denervation of the knee.12,16–18
This technique involves large open surgical exposures and has less-than-
ideal outcomes.19 Current less-invasive ultrasound-guided ablation
techniques are a welcome addition to management of this difficult
clinical problem.
Indication for IPN block: Arthroscopic surgery of the knee is associated
with postoperative pain in as many as 60% to 90% of patients.1 The
blockade of this nerve could improve postoperative analgesia. The use of
an ultrasound-guided block of the IPN has shown to improve pain
control and increase the number of sleep hours postoperatively.1
Ultrasonography-guided technique to block the IPN was associated with
an extended duration of cutaneous anesthesia (median block duration 27
hours) and reduced incidence of moderate-severe pain (Visual Analog
Scale >3).1 The IPN block is not only done for postoperative pain
control but also for the management of IPN neuropathy using steroids or
as a diagnostic step before cryoablation or pulsed radiofrequency
ablation.
Technique: The block is performed while the patient is in the supine
position with external rotation of the hip. After aseptic preparation of the
skin, a linear 5-to 10-MHz probe is placed transversely about 10 cm
proximal to the medial joint knee line (Fig. 25.3). The probe is adjusted
until the sartorius and gracilis muscles are visualized (Fig. 25.4). The
saphenous nerve is identified deep to the gracilis fascia. The nerve can
be followed proximally to confirm its relation to the femoral artery.
Scanning of the saphenous nerve is also done in a caudal direction till
the origin of the IPN is identified (Fig. 25.5). After proper skin
topicalization, a 25G, 3.5-inch needle is advanced using an out-of-plane
technique. The volume of the local anesthetic injected depends on the
indication of the block. If the block is done as a diagnostic step before
cryoablation, a small volume (2 mL) should be used to decrease the
chance of spilling over onto the saphenous nerve. Blockade for
postoperative pain control requires a larger volume.
Complications: The common risks include bleeding, infection, and pain.
The other complication is skin necrosis likely secondary to steroid use
during the IPN block.4
Conclusion Injury to the IPN is a known cause of postoperative anterior knee
pain. Pain physicians should become familiar with the pathophysiology,
anatomy, diagnosis, and treatment of IPN pain syndromes, as they are common,
especially after orthopedic procedures. Recognition and aggressive treatment
will decrease suffering and disability. The utilization of ultrasound guidance for
visualization of the nerve, although not very popular at this time, increases the
accuracy of IPN diagnostic blocks. Familiarity with the technique is necessary
for pain physicians as well as orthopedic surgeons. A simple diagnostic block of
the nerve may protect patients from having unnecessary or misdirected
additional surgery.
The saphenous nerve at the ankle can be found next to the saphenous
vein, usually posterior to the vein where it has a hyperechoic appearance
(Fig. 25.6). The saphenous nerve is located inferolateral to the
saphenous vein.
Patient position: Supine or recumbent with foot and ankle exposed.
Transducer: A 25-mm linear array oscillating at 10 to 13 MHz provides
an excellent image. The small transducer size makes it easy to control
the probe.
Transducer position: Saphenous: axially anterosuperior to medial
malleolus (Fig. 25.8).

Needle: 3-cm, 25G nerve block needle.


Local anesthetic: 2 to 5 mL/nerve of 0.5% ropivacaine or 0.5%
bupivacaine.
Technique: The skin is washed, and sterile ultrasound gel is applied to
the skin. A sterilized or covered probe is used for the procedure. The
foot is stabilized, the probe is applied, and the hand is anchored to hold
the probe steady. An anterior or posterior approach may be used for the
saphenous nerve. Very little volume is necessary to create
circumferential spread.
References
1. Lundblad M, Kapral S, Marhofer P, et al. Ultrasound-guided
infrapatellar nerve block in human volunteers: description of a
novel technique. Br J Anaesth. 2006;97:710–714.
2. Bademkiran F, Obay B, Aydogdu I, et al. Sensory conduction study
of the infrapatellar branch of the saphenous nerve. Muscle Nerve.
2007;35:224–227.
3. Mistry D, O’Meeghan C. Fate of the infrapatellar branch of the
saphenous nerve post total knee arthroplasty. ANZ J Surg.
2005;75:822–824.
4. Trescot AM, Brown MN, Karl HW. Infrapatellar saphenous
neuralgia-diagnosis and treatment. Pain Physician. 2013;16:E315–
E24.
5. Tifford CD, Spero L, Luke T, et al. The relationship of the
infrapatellar branches of the saphenous nerve to arthroscopy portals
and incisions for anterior cruciate ligament surgery. An anatomic
study. Am J Sports Med. 2000;28:562–567.
6. Tsenter J, Schwartz I, Katz-Leurer M, et al. A novel technique for
conduction studies of the infrapatellar nerve. PM R. 2012;4:682–
685.
7. Leliveld MS, Verhofstad MH. Injury to the infrapatellar branch of
the saphenous nerve, a possible cause for anterior knee pain after
tibial nailing? Injury. 2012;43:779–783.
8. Horn JL, Pitsch T, Salinas F, et al. Anatomic basis to the ultrasound-
guided approach for saphenous nerve blockade. Reg Anesth Pain
Med. 2009;34:486–489.
9. Mochida H, Kikuchi S. Injury of infrapatellar branch of saphenous
nerve in arthroscopic knee surgery. Clin Orthop Relat Res.
1995:88–94.
10. Beaufils P, Gaudot F, Drain O, et al. Mini-invasive technique for
bone patellar tendon bone harvesting: its superiority in reducing
anterior knee pain following ACL reconstruction. Curr Rev
Musculoskelet Med. 2011;4:45–51.
11. Sabat D, Kumar V. Nerve injury during hamstring graft harvest: a
prospective comparative study of three different incisions. Knee
Surg Sports Traumatol Arthrosc. 2013;21:2089–2095.
12. Dellon AL, Mont MA, Mullick T, et al. Partial denervation for
persistent neuroma pain around the knee. Clin Ortho Relat Res.
1996:216–222.
13. Love C, Marwin SE, Palestro CJ. Nuclear medicine and the infected
joint replacement. Semin Nucl Med. 2009;39:66–78.
14. Ghanem E, Parvizi J, Burnett RS, et al. Cell count and
differentiation of aspirated fluid in the diagnosis of infection at the
site of total knee arthroplasty. J Bone Joint Surg Am.
2008;90:1637–1643.
15. Math KR, Zaidi SF, Petchprapa C, et al. Imaging of total knee
arthroplasty. Semin Musculoskelet Radiol. 2006;10:47–63.
16. Dellon AL, Mont MA, Krackow KA, et al. Partial denervation for
persistent neuroma pain after total knee arthroplasty. Clin Orthop
Relat Res. 1995:145–150.
17. Nahabedian MY, Mont MA, Hungerford DS. Selective denervation
of the knee: experience, case reports, and technical notes. Am J
Knee Surg. 1998;11:175–180.
18. Nahabedian MY, Johnson CA. Operative management of
neuromatous knee pain: patient selection and outcome. Ann Plast
Surg. 2001;46:15–22.
19. Lundblad M, Forssblad M, Eksborg S, et al. Ultrasound-guided
infrapatellar nerve block for anterior cruciate ligament repair: a
prospective, randomised, double-blind, placebo-controlled clinical
trial. Eur J Anaesthesiol. 2011;28:511–518.
26
Ultrasound-Guided
Proximal Obturator
Block: A Proximal
Interfascial
Technique

AHMAD MOUHAMMAD TAHA

Background: The obturator nerve is a branch of the lumbar plexus. It


arises from the anterior divisions of lumbar nerves 2, 3, and 4. The
obturator nerve usually divides into anterior and posterior branches
before it enters the obturator canal, although occasionally it splits from
within the canal. After both obturator branches exit the pelvis, they lie
close together for a short distance, between the pectineus muscle
(superiorly) and the obturator externus muscle (inferiorly) (Fig. 26.1),
where both branches can be blocked using a single injection. Distally,
the two branches diverge. The anterior branch runs lateral, anterolateral,
and then anterior to the adductor brevis muscle, whereas the posterior
branch runs deep. The obturator nerve supplies the hip joint (via its hip
branch), the knee joint (via the posterior branch), the adductor muscles
(via the anterior and posterior branches), and a highly variable skin area
on the medial aspect of the thigh (via the anterior branch).
The main obturator nerve and its two branches are flat nerves and
difficult to identify. Even with combined ultrasound nerve stimulation
guidance, multiple needle passes are usually required. On the contrary,
the facial border of the adductor muscles can be easily identified. On the
ultrasound image, the hyperechoic fascial borders of the pectineus,
adductor longus, and brevis muscles usually form a characteristic
tricompartmental configuration; this visually resembles the letter Y with
its stem directed posteriorly (Fig. 26.2).

Indications
1. Knee and above-knee surgeries
2. Transurethral resection of bladder tumor (TURT)
3. Chronic hip pain
4. When prolonged thigh tourniquet inflation is required
Transducer type: Linear probe (10 to 13 MHz).

Needle: Short-beveled, 5 cm. A longer needle may be required for obese


patients or when an in-plane technique is used.
Local anesthetics: 10 to 15 mL of 2% lidocaine, 0.5% ropivacaine.
Patient position: Supine. The thigh is abducted and externally rotated.
Transducer position: On the medial aspect of the inguinal crease and
aimed posteriorly.
Block technique: After skin preparation, draping, and probe sheathing:
A. Identify the pectineus muscle. It lies between the femoral vessels
(medially) and the letter Y (laterally) (Fig. 26.3).

B. Tilt the ultrasound probe 40 to 50 degrees cranially while following


the pectineus muscle to identify the superior pubic ramus. It appears
as a hyperechoic structure deep to the lateral fibers of the pectineus
muscle. In this plane, the obturator externus could be seen deep to
the pectineus muscle (Fig. 26.4).
C. Advance the needle and infiltrate the local anesthetic (LA) between
the pectineus and the obturator externus muscles (Fig. 26.5).

Assessment: Motor examination: Marked weakness of adduction power


could be usually achieved within 5 minutes after a successful block;
however, complete loss of adduction could not be achieved except when
the sciatic and femoral nerves are also blocked as well. Sensory
examination is of little value.
Tips
1. At the level of inguinal crease, the anterior obturator nerve can be
identified at the center of the letter Y (Fig. 26.2); however, it is rarely
blocked alone.
2. The pectineus muscle is thin near its origin; therefore, its deeper
facial plane (between it and the obturator externus muscle) can
usually be visualized at 4 cm depth. Before needle advancement, it is
better to decrease the depth of the ultrasound beams as possible, just
enough to visualize this facial plane. This improves the resultant
sonographic image and the accuracy of LA placement.
3. If the needle is advanced from lateral to medial (via an in-plane
technique), it may injure the femoral vein (Fig. 26.6), and if
advanced in the opposite direction (from medial to lateral), the
presence of genitalia will leave no room to manipulate the needle.
Therefore, the author prefers to advance the needle via the out-of-
plane technique (inferior to the ultrasound probe).

4. The obturator nerve is accompanied by numerous vessels; therefore,


before LA injection, careful aspiration is important (Fig. 26.7).
5. The selection of LA depends on the duration of surgery. When the
patient is at risk of LA toxicity (old or hepatic), lower LA volume
(just enough to separate the pectineus and obturator externus
muscles) and concentration can be used successfully.
6. During the TURT, the bladder is distended and its wall become close
to the obturator nerve. Therefore, electrocauterization may induce the
obturator reflex. This reflex can be prohibited by bilateral obturator
nerve block that can be added to the general (if muscle relaxant is not
desirable) or spinal anesthesia.
7. This proximal block technique can block the hip branch of the
obturator nerve as well; therefore, it could be used as a treatment for
hip pain.

Suggested Readings
Akkaya T, Ozturk E, Comert A, et al. Ultrasound-guided obturator nerve block: a sonoanatomic
study of a new methodologic approach. Anesth Analg. 2009;108:1037–1041.
Anagnostopoulou S, Kostopanagiotou G, Paraskeuopoulos T, et al. Anatomic variations of the
obturator nerve in the inguinal region: implications in conventional and ultrasound regional
anesthesia techniques. Reg Anesth Pain Med. 2009;34:33–39.
Bouaziz H, Vial F, Jochum D, et al. An evaluation of the cutaneous distribution after obturator
nerve block. Anesth Analg. 2002;94:445–449.
Fujiwara Y, Sato Y, Kitayama M, et al. Obturator nerve block: from anatomy to ultrasound
guidance. Anesth Analg. 2008;106:350–352.
Sinha SK, Abrams JH, Houle TT, et al. Ultrasound-guided obturator nerve block: an interfascial
injection approach without nerve stimulation. Reg Anesth Pain Med. 2009;34:261–264.
Soong J, Schafhalter-Zoppoth I, Gray AT. Sonographic imaging of the obturator nerve for regional
block. Reg Anesth Pain Med. 2007;32:146–151.
Taha AM. Ultrasound-guided obturator nerve block: a proximal interfascial technique. Anesth
Analg. 2012;114:236–239.
27
Ultrasound-Guided
Distal Obturator
Nerve Block
YOSHIHIRO FUJIWARA AND
TORU KOMATSU

Background and indications: The obturator nerve emerges from the


medial border of the iliopsoas muscle, posterior to the psoas muscle, and
anterior to the obturator internus muscle. The nerve lies approximately 2
cm lateral and 2 cm distal to the pubic tubercle (Fig. 27.1). It carries a
mixed population of sensory and motor fibers from roots L2, L3, and L4.
As it enters the obturator canal, it divides into an anterior and posterior
branch. The anterior branch supplies the adductor longus, adductor
brevis, and gracilis muscles as well as the skin over the medial thigh or
posterior knee and a branch to the hip joint. The posterior branch
supplies the obturator externus, the quadratus femoris, and the adductor
magnus muscles as well as the knee joint. Ten percent of patients have
an accessory branch from the lumbar plexus.
Obturator nerve block is used to prevent adductor contraction evoked by
electrocautery near the bladder wall and to supplement analgesia for
major knee surgery. Obturator nerve block can also be used to treat
spasticity of the adductor thigh muscles, obturator neuralgia, or pain in
the hip joint.
Anatomy: The femoral neurovascular bundle and pectineus muscle
medial to the femoral vein serve as major landmarks. More medially, the
three muscle layers of the adductor muscles (from superficial to deep:
adductor longus muscle, adductor brevis muscle, and adductor magnus
muscle) must also be identified. The anterior branches of the obturator
nerve at this level can be found between the adductor longus muscle and
the adductor brevis muscle. The posterior branch is found between the
adductor brevis muscle and the adductor magnus muscle. The nerves
appear as oval hypoechoic structures within the hyperechoic fascia.

Patient position: Supine. The leg is externally rotated with the hip and
knee flexed.

Transducer type: 25 mm linear probe oscillating at 10 to 13 MHz. 11


mm curved array oscillating at 6 to 10 MHz (obese patients).

Transducer position: The probe is positioned parallel to the inguinal


crease and medial to the femoral artery (Fig. 27.2).

Needle: Short-beveled, insulated, 70-mm 22G needle. Whenever


possible, a nerve stimulator should be used in combination.

Local anesthetic: 10 to 15 mL of 0.2% ropivacaine or 0.25% bupivacaine


with epinephrine, 5 mL of 6% phenol in glycerin (neuroablation for pain
patients).

Technique: After sterile skin preparation, a short-beveled stimulating


needle is advanced via an out-of-plane or an in-plane technique (Fig.
27.2). The needle is placed at the lateral border of the probe and under
ultrasound guidance is advanced toward the fascia between the adductor
longus and adductor brevis muscles (Fig. 27.3). Then, the nerve
stimulator is turned on. The intensity of the stimulating current is
initially set to deliver 1 mA and gradually decreased. If a stimulating
current less than 0.6 mA evokes a motor response, 3 to 5 mL of local
anesthetic is injected after negative aspiration (Fig. 27.4). This
procedure is repeated along the fascia until the motor response is
eliminated. The needle is then advanced until it is in contact with the
posterior branch between the adductor brevis and adductor magnus
muscles. After the nerve is identified by stimulation, 3 to 5 mL of local
anesthetic is injected. The same approach and technique are used as for
surgical block. After a prognostic block with 1% lidocaine, neurolytic
agents (alcohol or phenol) are injected.
Tips 1. The authors usually block the anterior branch first. In some
patients, a motor response to the posterior branch cannot be obtained.
This is presumably because of posterior or proximal spread of local
anesthetic injected around the anterior branch.
2. Even complete block of both branches may not guarantee the
complete inactivation of adductor muscle of the lower limb. In some
patients, an accessory obturator nerve, which may not be blocked
with this technique, innervates the adductor muscles.
3. When this block is provided for patients without general or spinal
anesthesia, subcutaneous infiltration of local anesthetic is needed. If
infiltrated too deep, the motor response may be lost to nerve
stimulation as a result of unintentional block of the anterior branch.
4. At this site of injection, both branches of the obturator nerve may
consist of several small fibers. Thus, the authors search for an
adductor response along the fascia and inject small amounts of local
anesthetic to each adductor response, rather than give a large single
injection.
5. Puncture of the obturator artery has led intraperitoneal and
retroperitoneal hemorrhage.

Bibliography
Fujiwara Y, Sato Y, Kitayama M, et al. Obturator nerve block: from anatomy to ultrasound
guidance. Anesth Analg. 2008;106:350-a–351-a.
Fujiwara Y, Sato Y, Kitayama M, et al. Obturator nerve block using ultrasound guidance. Anesth
Analg. 2007;105:888–889.
Helayel PE, da Conceição DB, Pavei P, et al. Ultrasound-guided obturator nerve block: a
preliminary report of a case series. Reg Anesth Pain Med. 2007;32:221–226.
Saranteas T, Paraskeuopoulos T, Alevizou A, et al. Identification of the obturator nerve divisions
and subdivisions in the inguinal region: a study with ultrasound. Acta Anaesthesiol Scand.
2007;51:1404–1406.
Soong J, Schafhalter-Zoppoth I, Gray AT. Sonographic imaging of the obturator nerve for regional
block. Reg Anesth Pain Med. 2007;32:146–151.
28
Ultrasound-Guided
Proximal Parasacral
Block
KEVIN KING AND JACQUES E. CHELLY

Background and indications: The parasacral block was originally


described using a landmark approach and neurostimulation. Early
reported success rates range from 60% to 90%. The use of ultrasound
may enable the practitioner to improve on this success rate as well as to
avoid some of the described complications and side effects associated
with the performance of this block. The sacral plexus originates from
roots L4, L5, S1, S2, and S3. The roots emerge from the deep surface of
the sacrum, where they join to form the sciatic nerve and its tributaries.
It leaves the pelvis at the level of the greater sciatic foramen. At this
level, the nerves are covered by the gluteus maximus and the piriformis
muscles. After the sciatic nerve passes underneath the piriformis muscle,
it extends caudally and passes over the superior and inferior gemelli
muscles and the obturator internus muscles (Fig. 28.1). The parasacral
plexus block allows the anesthesiologist to block the sciatic nerve, the
posterior femoral cutaneous nerve, and the superior gluteal nerve, which
contributes articular sensory branches to the posterior capsule of the hip.
The performance of the parasacral block is associated with a risk of
bowel or rectum perforation, as the pelvic aponeurosis lies immediately
deep to the parasacral plexus. This aponeurosis or fascia is often poorly
visualized on ultrasound except in athletes and very thin individuals. It is
impossible to appreciate in a blind technique. Parasacral block may also
anesthetize the pudendal and other sacral nerves that lie nearby.
Parasacral block can be used for surgical anesthesia of the hip or knee in
conjunction with lumbar plexus block.

Anatomy: Due to the depth of the sacral plexus, practitioners may wish
to combine stimulation with ultrasound guidance. The probe is placed at
6 to 8 cm caudal to the posterior superior iliac spine (PSIS) along the
line joining ischium and the PSIS (Fig. 28.2). Utilizing ultrasound, the
landmarks include the PSIS and ischial tuberosity (IT) and the gluteus
maximus and the piriformis muscle and gluteal artery (Figs. 28.2 to
28.4A to C). The sacral plexus is a flat hyperechoic structure at the same
depth as the ischium and sacrum.
Patient position: Lateral decubitus with the operative side up. The hip
and knee are flexed.

Transducer: 60-mm curved array oscillating at 2 to 5 MHz.

Transducer position: Transverse in horizontal plane over superior


medial aspect of buttock.

Needle: 15-cm either 21G (single-block) or 18G Tuohy (catheter


insertion).

Local anesthetic: For surgical anesthesia administered as a one-shot


block, use 0.5% ropivacaine or bupivacaine. In the case of a continuous
block, use 0.2% or 0.1% ropivacaine or 0.06% or 0.03% bupivacaine
depending on the indication and the need for preserving motor function.

Volume of local anesthetic: 7 to 15 mL for a single block and a rate of 3


to 6 mL/hr for an infusion.

Technique: The patient is placed in the lateral decubitus position with the
operative side up. The hip and knee are flexed whenever possible.
Standard monitors are applied, sedation administered, and the skin over
the lower back and buttock is cleaned/disinfected. The PSIS and IT are
identified, and a line connecting these points is drawn. The curvilinear
transducer is placed 6 to 8 cm from the PSIS oriented perpendicular and
lateral to this line at the level of the coccyx (Figs. 28.2 and 28.4A). The
level of the coccyx is located approximately at the superior aspect of the
intergluteal or natal cleft. The piriformis muscle extends from the
anterior-medial surface of the sacrum to the femur. Although in most
cases it is difficult to identify the piriformis muscle, its identification is
greatly facilitated by rotating the thigh. This movement allows for a
dynamic identification of the muscle as the muscle contracts. The
sacrum and the ischium are usually easy to identify. The sacral plexus is
identified on ultrasound as a flat, hyperechoic structure at the same
depth as the ischium and sacrum (Fig. 28.3). The needle is inserted
medial or lateral to the probe, using an in-plane approach.
The stimulating needle is also set up to deliver a current of 0.8 to 1.5
mA. Once the parasacral plexus has been identified by ultrasound
visualization and a proper motor response is observed at a current of 0.5
mA or above, the local anesthetic solution is slowly injected after
negative aspiration for blood. Stimulation currents at 0.5 mA or below
0.5 mA are virtually guaranteed to result in intraneural placement of the
needle.

Alternate techniques: With the patient in the previously described


position and with sterile prep performed, mark the location of the PSIS
and the lateral border of the sacrum. Then mark the posterior-superior-
medial aspect of the trochanter and draw a line connecting this point to
the PSIS. Orient the transducer/probe along this line. With the medial
part of the transducer/probe over the presumed PSIS, note the shallow
depth of a hyperechoic line or arc with dark shadow below which
verifies the location of the PSIS. A cortical bone line descending
laterally and continuously across the viewing screen is the extrapelvic
border of the ilium. Then slowly move the probe in a inferomedial
direction until the ilium is no longer visible (about 1 to 2 cm). At this
point, visualization to deeper structures indicates that the viewing field is
the greater sciatic foramen. The lateral border of the sacrum may also be
visualized on the medial side of screen. A deep cortical line (4 to 9 cm
depth in adults) on the lateral side of the screen indicates the ischium or
ischial spine. The gluteus maximus muscle should also be evident. The
study that described this viewing orientation referred to the technique as
the parasacral parallel shift and reported reliable
visualization/identification of the piriformis muscle and the underlying
parasacral plexus. More importantly, it reported successfully
anesthetizing seven ASA class III–IV patients for total hip arthroplasty
with only mild awake sedation added to lumbar plexus blockade and
parasacral plexus blockade for images of the greater sciatic foramen and
parasacral plexus obtained using the parasacral parallel shift approach.
The piriformis muscle can be seen thickened/contracted and in a relaxed
state (Fig. 28.5A to C).
A variant of this approach is represented by simply moving the
horizontal viewing plane about 2 cm cephalad of the intergluteal cleft, so
that the probe lies transversely over the lateral border of the mid/upper
sacrum and immediately caudad to the PSIS. This position should yield
a view that shows the sacral border and the plexus with associated
gluteal vessels approximately 1 to 2 cm deep and 2 to 3 cm lateral to the
sacrum (Fig. 28.6A,B). The piriformis may be noted at this level. The
piriformis muscle can be identified by asking the patient to gently and
slowly turn his or her foot outward, thus externally rotating the femur
through piriformis contraction. This allows for dynamic identification of
this adductor muscle/external rotator muscle. Seeing the piriformis
shorten and thicken just above the presumed parasacral plexus is another
method of confirming the location of the nerves. It is also an indicator
that lends safety to the procedure, as one should not advance the needle
deeper than 1 cm past the anterior surface of the piriformis.
Although the classic approach is very effective, some have advocated
the more cephalad techniques. In support of the cephalad approaches to
the parasacral plexus is the claim that the lower classic approach may
not adequately block nerves that carry sensory information from the hip
joint. Anatomic studies verify posterior hip capsule innervation by
branches of L5 and S1 components of the parasacral plexus. A German
anatomical study found that blockade of the superior gluteal nerve is
necessary for effective surgical block. Since superior and inferior gluteal
nerves leave the parasacral plexus immediately before or after exiting
the greater sciatic foramen, theoretically block of these nerves is
achieved more reliably by using a more cephalad approach. More
proximal blocks should be regarded as paravertebral blocks with
increased risk of central spread and damage to the nerves. More
caudal/distal block often yields the undesirable effect of genital/perineal
numbness or pudendal nerve block. When low stimulation thresholds are
used (less than 0.5 mA), either approach provides good surgical
anesthesia because the local anesthetic spreads within the mesoneurium
of the nerve plexus itself to both proximal and distal points along the
nerve plexus.

Suggested Readings Bendtsen TF, Lonnqvist PA, Jepsen KV, et al.


Preliminary results of a new ultrasound-guided approach to block
the sacral plexus: the parasacral parallel shift. Br J Anaesth.
2011;107:278–280.
Birnbaum K, Prescher A, Hessler S, et al. The sensory innervation of the hip joint—an anatomical
study. Surg Radiol Anat. 1997;19:371–375.
Latzke D, Marhofer P, Zeitlinger M, et al. Minimal local anaesthetic volumes for sciatic nerve
block: evaluation of ED99 in volunteers. Br J Anaesth. 2010;104:239–244.
Ben-Ari AY, Joshi R, Uskova A, et al. Ultrasound localization of the sacral plexus using a
parasacral approach. Anesth Analg. 2009;108:1977–1980.
Mansour NY. Reevaluating the sciatic nerve block: another landmark for consideration. Reg
Anesth. 1993;18:322–323.
29
Ultrasound-Guided
Distal Parasacral
Block

AHMAD MOUHAMMAD TAHA

Background: The sciatic nerve is the terminal branch of the sacral


plexus. It arises in the parasacral area at the back of the pelvis between
the piriformis (posteriorly) and the pelvic fascia (anteriorly). It exits the
pelvis through the greater sciatic foramen, just medial to the posterior
border of ischium (PBI). The PBI forms the lower lateral bony edge of
the greater sciatic foramen (Fig. 29.1). The sciatic nerve supplies the hip
and knee joints, the hamstring muscle, and nearly the entire leg and foot.
In its proximal course, the sciatic nerve is deep and difficult to identify;
therefore, bones, which are easily identified sonographic landmarks, are
used to localize the sciatic nerve. In the parasacral technique, the PBI is
used to identify the sciatic nerve. Unlike other sciatic block techniques,
the parasacral technique can block the entire sacral plexus as well. The
technique described here is slightly distal to the technique described by
Ben-Ari1 and more proximal than the traditional transgluteal technique
described by Labat.2
Indication: All lower limb surgeries (knee, below-knee, above-knee, and
hip surgeries).

Transducer type: A curved probe (2 to 5 MHz).

Needle: Short-beveled, 12-cm insulated needle connected to a nerve


stimulator unit. Longer needle may be required for morbidly obese
patients.

Local anesthetics: 20 mL of 2% lidocaine, 0.5% ropivacaine. The


selection of local anesthetic (LA) depends on the duration of surgery.

Patient position: Sims position.

Transducer position: On an axial plane, 8 cm lateral to the uppermost


point of the gluteal cleft.

Block technique: After skin preparation, draping, and probe sheathing:

1. Identify the PBI, a characteristic curved hyperechoic bony structure


(Fig. 29.2).
2. The sciatic nerve appears as a hyperechoic structure, just medial to
the PBI. Adjust the ultrasound probe (tilt, rotation and/or rocking to
improve the sciatic image).
3. Advance the needle toward the nerve. Adjust the needle tip, under the
ultrasound guidance, to elicit hamstring leg or foot twitches with 0.3
to 0.5 mA current. Usually, there is no need to reposition the needle.
Assessment: Successful block will result in:

1. Common peroneal nerve block: loss of dorsal flexion and sensation at


the dorsal surface of the foot.
2. Tibial nerve block: loss of plantar flexion and sensation at the plantar
surface of the foot.
-The parasacral block usually has a rapid onset but sometimes delays for
30 minutes. Before surgery starts, make sure that both tibial and
common peroneal nerves are successfully blocked.

Tips
• Usually, the PBI could be identified at the level of uppermost point of
the gluteal cleft. Otherwise, slide the probe cranially to identify the ala
of ilium and then slide it caudally until the cranial aspect of greater
sciatic foramen appears. At this point, tilt the probe caudally to
identify the PBI (Fig. 29.3).
• The sciatic nerve runs from deep (the pelvis) to superficial (the gluteal
region). Therefore, a caudal tilt of the ultrasound probe improves its
image resolution.
• Other parasacral contents can confirm the nerve identification. The
sciatic nerve lies deep to the piriformis muscle, lateral to the inferior
gluteal artery, and if followed caudally, it comes to rest on the back of
ischium (Fig. 29.4). In thin patients, peristaltic movements can be
seen deep to the nerve.

• The author prefers to advance the needle via the out-of-plane


technique. The in-plane technique could be used; however, because
the sciatic nerve is deep at the parasacral area, the needle
advancement angle is steep, impairing its visualization, and because
the needle passage may also be obstructed with the sacrum (when the
needle is inserted medial to the ultrasound probe).
• The sciatic nerve is related to numerous vessels: the inferior gluteal
and the internal pudendal (Fig. 29.4). Therefore, careful aspiration and
slow LA injection are important.
• In morbidly obese patients, the nerve may not be identified. In this
case, you can advance the needle, with electrical stimulation, just
medial to PBI (Fig. 29.5).

References
1. Ben-Ari AY, Joshi R, Uskova A, Chelly JE. Ultrasound localization
of the sacral plexus using a parasacral approach. Anesth Analg.
2009;108:1977–1980.
2. Labat G. Regional Anesthesia: its technique and clinical
applications. Philadelphia: W.B. Saunders, 1922;286–291.

Suggested Readings
Ben-Ari AY, Joshi R, Uskova A, et al. Ultrasound localization of the sacral plexus using a
parasacral approach. Anesth Analg. 2009;108:1977–1980.
Bendtsen TF, Lonnqvist PA, Jepsen KV, et al. Preliminary results of a new ultrasound-guided
approach to block the sacral plexus: the parasacral parallel shift. Br J Anaesth. 2011;107:278–
280.
Mansour NY. Reevaluating the sciatic nerve block: another landmark for consideration. Reg
Anesth. 1993;18:322–323.
Taha AM. A simple and successful sonographic technique to identify the sciatic nerve in the
parasacral area. Can J Anaesth. 2012;59:263–267.
30
Ultrasound-Guided
Anterior Sciatic
Nerve Block
YASUYUKI SHIBATA, TORU KOMATSU, AND

LAURENT DELAUNAY

Background and indications: Ultrasound-guided block of the sciatic


nerve can be conducted from the anterior aspect of the patient as well as
from the lateral or posterior positions. When the block is conducted from
the anterior approach (at the level of the lesser trochanter), it is too distal
to be of use in hip surgery. However, the anterior approach allows the
patient to remain supine. This is especially advantageous in patients who
have fractures of the lower extremity. Anterior sciatic block is used for
surgery of the distal posterior thigh, posterior knee, leg, or foot.

Anatomy: From the anterior approach, the sciatic nerve lies deep and
medial to the femur (Fig. 30.1). The nerve is bounded laterally by the
gluteus maximus muscle and medially by the biceps femoris and
semimembranosus/semitendinosus muscles (frequently referred to as the
hamstring muscles). Just anterior to the nerve lies the adductor magnus
muscle. Medial to the nerve, and quite superficial in the thigh at these
levels, the femoral vessels and nerve can be seen (Fig. 30.1). On
ultrasound, the sciatic nerve will appear as a hyperechoic oval or round
structure (Fig. 30.2). In some patients, the obturator artery and nerves
can also be seen lying deep to the adductor longus and between the
adductor brevis and adductor magnus muscles (Fig. 30.3).
Patient position: The patient remains supine. The thigh is externally
rotated, and the knee is flexed. The leg should be externally rotated so
that the sciatic nerve is rotated to a position medial to the femur.

Transducer: 60 mm curved array oscillating at 2 to 5 MHz.

Transducer orientation: The transducer is placed transversely on the


anterior thigh (Fig. 30.2).

Needle: 10 to 15 cm, 21G (single injection) or 18G Tuohy (continuous


block) insulated needle.

Local anesthetic: 10 to 20 mL of 0.5% ropivacaine or bupivacaine 0.5%


with epinephrine.

Technique: Patient monitors are attached and sedation is administered.


Oxygen by face mask is initiated. Appropriate markings are placed on
the indicated extremity for confirmation, and a brief preprocedural
“time-out” is conducted at the bedside. An antiseptic is utilized to
prepare the skin over the block area.
The lesser trochanter is imaged, and the probe is moved proximally or
distally along the thigh until the nerve is imaged medial and posterior
(deep) to the femur (Fig. 30.2). Toggling of the probe is usually
necessary to create the optimum image of the nerve, usually at a depth of
6 to 10 cm. Some practitioners also use the femoral artery as a landmark.
In many patients, the sciatic nerve lies along a line drawn perpendicular
through the femoral artery. The needle is placed in-line, medial to the
probe and advanced in a posterolateral direction until the nerve is
contacted or penetrated. Most practitioners prefer to utilize nerve
stimulation in conjunction with ultrasound imaging with this approach
because of the depth of the nerve. A stimulation threshold of 1 mA or
more is utilized to identify the nerve. Even at this setting, puncturing the
nerve is commonly required to elicit stimulation. Once the nerve is
identified, local anesthetic is injected in 3 to 5 mL aliquots. If the needle
is within the nerve sheath, or mesoneurium, 5 to 10 mL of local
anesthetic solution is sufficient for a profound block. If the local
anesthetic solution is deposited outside the nerve, a “halo” should be
created around the nerve, utilizing 10 to 20 mL of solution. Because of
the size of the nerve and the dual covering (epineurium and
mesoneurium), long latencies are common when local anesthetic is
deposited outside the nerve.
Some practitioners prefer to image the nerve in short axis and use an
out-of-plane approach for the needle (Fig. 30.4). Other practitioners
image the nerve in the short axis and then rotate the probe into a long-
axis view.1 In this case, the nerve appears as a hyperechoic streak
between the adductor magnus muscle and the gluteus maximus muscle.
When this approach is used, the needle is inserted in line with the probe.

Summary of evidence: Chan et al.2 evaluated 15 volunteers with


ultrasound at different approaches to establish optimal visualization of
the sciatic nerve. For the anterior approach, the authors utilized a 2-to 5-
MHz curved probe. Overall, the nerve was easily visualized in 87% of
patients, and the authors were able to stimulate the nerve within two
attempts in all patients. Chantzi et al.3 described the anterior approach to
sciatic block in 18 obese subjects utilizing ultrasound guidance. The
authors were able to identify the nerve in 17 out of 18 patients and
provided successful nerve stimulation in all of these within four
attempts. They noted that due to the depth of the nerve and the angle of
needle insertion, it is quite difficult to simultaneously image both needle
and nerve during this technique.
References
1. Tsui BCH, Ozelsel TJP. Ultrasound-guided anterior sciatic nerve
block using a longitudinal approach: Expanding the view [Letter].
Reg Anesth Pain Med. 2008;33:275–276.
2. Chan VWS, Nova H, Abbas S, et al. Ultrasound examination and
localization of the sciatic nerve. Anesthesiology. 2006;104:309–
314.
3. Chantzi C, Saranteas T, Zogogiannis J, et al. Ultrasound
examination of the sciatic nerve at the anterior thigh in obese
patients. Acta Anaesthesiol Scand. 2007;51:132.
31
Ultrasound-Guided
Subgluteal Sciatic
Block
STEVEN L. OREBAUGH AND PAUL E.

BIGELEISEN

Background and indications: Sciatic nerve block has been in use by


anesthesiologists for over 90 years. The traditional approach has been to
create landmarks over the gluteus and traverse this large muscle en route
to the nerve. However, di Benedetto1 described a somewhat more distal
approach in which the nerve is blocked in the proximal thigh just as it
leaves the pelvis. In di Benedetto’s1 original description, the procedure
was referred to as the subgluteus block, and will hereafter be referred to
as the infragluteal approach to avoid confusion with the subgluteal
space. When compared to the transgluteal approach, the infragluteal
approach resulted in less patient discomfort and required less time with
fewer needle punctures than the time-honored approach through the
gluteus. Ultrasound has the potential to make the infragluteal sciatic
block even more efficient with less discomfort for the patient. Sciatic
block is provided for surgical procedures involving the distal femur,
knee joint, leg, ankle, and foot.

Anatomy: Infragluteal sciatic block guided by landmarks and nerve


stimulation relies on palpation of the greater trochanter and the ischial
tuberosity, as the nerve lies approximately midway between them. The
“groove” that lies between the hamstring muscles, which originate from
the ischial tuberosity, and the lateral edge of the vastus lateralis muscle
are then palpated. The nerve typically lies deep to this groove and is
sought 4 cm below the line connecting the two bony landmark (Figs.
31.1 and 31.2).
With ultrasound, the same landmarks may be located. The ischial
tuberosity and greater trochanter are readily imaged just proximal to the
gluteal fold (Fig. 31.1); below this fold, the overlying gluteus maximus
muscle is seen to thin out considerably. Deep to the gluteus, and
superficial to the quadratus femoris muscle, lies the sciatic nerve at a
depth of approximately 3 to 12 cm, depending on the patient’s habitus.
The nerve itself has a fusiform or wedge-shaped appearance and is
hyperechoic (Fig. 31.2). As one proceeds distally, the quadratus femoris
disappears and the nerve becomes more round or oval and lies on the
belly of the adductor magnus. Deeper, and lying somewhat more lateral
than the nerve, the hyperechoic arc of the femur is evident. There are
usually no significant vessels evident adjacent to the nerve at this level.

Patient position: Lateral decubitus position, with the affected side up or


prone.

Transducer: 40 to 60 mm curved array oscillating at 2 to 5 MHz. 11 mm


curved array oscillating at 6 to 10 MHz (smaller patients and children).

Transducer orientation: Transverse below the gluteal cleft. Significant


pressure on the transducer may be necessary in larger subjects.

Needle: 10-cm, 21G blunt needle (single injection) or 18G Tuohy needle
(continuous infusion).

Local anesthetic: 10 to 20 mL of 0.5% ropivacaine or 0.5% bupivacaine


with epinephrine.

Technique: The patient monitors are placed and judicious sedation is


introduced. Oxygen by face mask is initiated. Appropriate markings are
placed on the indicated extremity for confirmation, and a brief
preprocedural “time-out” is conducted at the bedside. An antiseptic is
utilized to prepare the skin of the block area.
Infragluteal block with real-time ultrasound guidance may be conducted
with or without nerve stimulator confirmation. In addition, the needle
may be inserted using an out-of-plane (superior or inferior to the axis of
the transducer) or an in-plane approach at the medial or lateral end of the
transducer.
After establishing an appropriate image of the nerve, the skin over the
area of needle insertion is anesthetized, along with the subcutaneous
tissues. The block needle is then introduced along the predicted pathway,
with subtle transducer adjustments to maintain an image of both the
nerve and needle. However, it is frequently difficult to maintain an
image of the needle due to the steep insertion angle, and careful
observation of tissue distortion caused by the advancing needle is
necessary. If doubt exists as to the position of the tip, a small amount of
dextrose may be injected to help localize the needle.
The nerve stimulator, if utilized, is turned on before needle introduction,
or later, as the needle approaches the nerves. The threshold for
stimulation has not been found to be of importance in this technique.
Local anesthetic solution, typically 20 to 25 mL, is then injected under
ultrasound guidance, ensuring that the nerve is surrounded by the
solution. This frequently requires two or more needle positions to
accomplish successfully. It is common to find that the needle, even
though placed against the suspected nerve, does not elicit motor
stimulation at typical currents (0.8 to 1.0 mA).2,3 As the needle is
advanced further, stimulation frequently ensues and may be vigorous.
This may be due to intraneural placement of the needle. In order to avoid
intraneural injection, the needle should then be pulled back, and a 1-mL
volume of local anesthetic solution injected to help localize the needle
tip before the full dose is injected.
For indwelling catheter placement, either an in-plane or an out-of plane
technique is possible. A sterile sheath should be utilized for placement of
an indwelling device, along with sterile gloves, cap, mask, and drapes.
After establishing an optimal view of the nerve, similar to that for the
single-injection block, the skin is anesthetized with a subcutaneous
wheal of local anesthetic. The needle is then guided to the nerve, and
nerve stimulation is utilized for confirmation if desired. For
nonstimulating catheters, a bolus of local anesthetic, saline, or dextrose
solution is used to confirm accurate needle tip placement and the
catheter is then inserted. If a stimulating catheter is desired, then the
stimulator is attached to the catheter after needle localization, and
placement is confirmed by continued stimulation as the catheter is
inserted (ultrasound will not be particularly useful during this step). By
either technique, after catheter placement, ultrasound imaging can be
used to confirm that solution injected through the catheter is deposited in
juxtaposition to the nerve. Color Doppler may improve the ability to
image in this situation, as the injection occurs.
Summary of clinical evidence: Chan et al.4 described the ultrasound-
guided sciatic nerve block via several different routes, including the
transgluteal, infragluteal, and anterior approach. The authors reviewed
the relevant anatomy in 15 volunteer subjects and were able to easily
visualize the nerve in 87% of them. Although blocks were not
performed, nerve stimulation was conducted successfully in all patients
within two attempts.
In an evaluation of the subgluteal space with ultrasound, Karmakar et
al.5 emphasized the importance of the ischial tuberosity, the greater
trochanter, and the space between the fascia lining the deep aspect of the
gluteus muscle and that lining the superficial aspect of the quadratus
femoris muscle, in which the sciatic nerve lies. The subgluteal space,
while somewhat proximal to the true infragluteal block, is distal to the
site of traditional transgluteal blocks. The authors evaluated five patients
undergoing ultrasound-guided blocks with deposition of local anesthetic
in the subgluteal space and noted successful anesthesia in all of them
within 20 minutes.
Van Geffen et al.6 described placement of subgluteal catheters in 10
children receiving lower limb surgery, utilizing the ultrasound guidance
as well as nerve stimulation for the placement of the stimulating
catheters. Spread of injectate on ultrasound predicted successful catheter
placement, and all 10 children received excellent postoperative
analgesia. In a randomized trial of ultrasound guidance versus nerve
stimulation for placement of sciatic blocks, Oberndorfer et al.7 found
that the success rate in 46 children was similar but that sensory block
was prolonged in the ultrasound group.
References
1. di Benedetto P, Bertini L, Casati A, et al. A new posterior approach
to the sciatic nerve block: a prospec-tive, randomized comparison
with the classic posterior approach. Anesth Analg. 2001;93:1040–
1044.
2. Perlas A, Niazi A, McCartney C, et al. The sensitivity of motor
response to nerve stimulation and paresthesia for nerve localization
as evaluated by ultrasound. Reg Anesth Pain Med. 2006;31:445–
450.
3. Sinha SK, Abrams JH, Weller RS. Ultrasound-guided interscalene
needle placement produces successful anesthesia regardless of
motor stimulation above or below 0.5 mA. Anesth Analg.
2007;105:848–852.
4. Chan VWS, Nova H, Abbas S, et al. Ultrasound examination and
localization of the sciatic nerve: a volunteer study. Anesthesiology.
2006;104:309–314.
5. Karmakar MK, Kwok WH, Ho AM, et al. Ultrasound guided sciatic
nerve block: description of a new approach at the subgluteal space.
Br J Anaesth. 2007;98:390–395.
6. van Geffen GJ, Gielen M. Ultrasound-guided subgluteal sciatic
nerve blocks with stimulating catheters in children: a descriptive
study. Ped Anesth. 2006;103:328–333.
7. Oberndorfer U, Marhofer P, Bosenberg A, et al. Ultrasonographic
guidance for sciatic and femoral nerve block in children. Br J
Anaesth. 2007;98:797–801.
32
Ultrasound-Guided
Popliteal and Lateral
Sciatic Block
STEVEN L. OREBAUGH, VIHANG D. SHAH,

AND PAUL E. BIGELEISEN

Background and indications: Block of the sciatic nerve at the popliteal


fossa is the most distal of the many approaches to the sciatic nerve.
When ultrasound is used for imaging, local anesthetic is readily
deposited around the sciatic nerve or around the tibial and common
peroneal nerves when the block is done distal to the bifurcation of the
sciatic nerve. Historically, the block was done with the patient prone
(posterior approach). With the advent of ultrasound, the patient is
usually placed in the lateral decubitus position with the operative side up
or in the supine position.
Block of the sciatic nerve at the popliteal fossa is indicated for surgical
procedures involving the toes, foot, or subtalar joint. It may not be
suitable as a primary anesthetic for major ankle procedures because
surgeons often choose a thigh (rather than calf) tourniquet for control of
bleeding during the procedure.

Anatomy: It is useful to identify the arteries and veins in the area


because of their position relative to the nerve. In the popliteal region, the
popliteal artery and vein can be located at the flexor crease (Fig. 32.1).
The tibial nerve is usually visible at this level, just lateral and/or
superficial to the vein. Lateral to the nerve lies the biceps femoris
muscle. Medial to the nerve lie the semimembranosus and
semitendinosis muscles (Fig. 32.1). Once the tibial nerve is identified at
the flexor crease, the probe can be moved slowly proximally, keeping
the nerve in the center of the image. The tibial nerve is joined by the
common peroneal nerve as the hamstring muscle bellies in the
surrounding area enlarge and more subcutaneous adipose tissue becomes
evident (Fig. 32.2). Between 5 and 10 cm above the flexor crease, the
smaller peroneal nerve will be seen lateral to the tibial nerve; above this
level, a single rounded nerve structure is apparent (Fig. 32.3), which
represents the conjoined tibial and peroneal nerves. The peroneal nerve
is frequently more difficult to locate than the tibial because it is smaller
and often appears to be made up of one or more large fascicles with little
surrounding connective tissue.
When imaging the level at which the sciatic nerve bifurcates, the
popliteal vessels are usually considerably deeper than the nerve (Fig.
32.2). Because significant pressure is often required to optimize the
nerve image, it is important to periodically decrease the pressure applied
with the transducer; this will help reveal the popliteal vein, which
otherwise may remain compressed and invisible. Deep to the vessels, the
hyperechoic, arcuate form of the cortex of the femur is evident.
At this level, the sciatic nerve is encased by a thick mesoneurial sheath.
Within the nerve, both the peroneal and tibial nerves are encased by their
proper epineurium. Within the epineurium of each nerve lie the fascicles,
each of which is bounded by perineurium. Currently, the term used to
describe the fascial sheath of the sciatic nerve, which surrounds both
components (tibial and common peroneal nerve), is a topic of
contention. Authors have referred to this anatomic structure as a
paraneural sheath1 or complex fascial layer,2,3 whereas others termed it
a common epineural sheath4 (Fig. 32.1E). The authors’ preference is to
call this the mesoneurium. Deposition of local anesthetic within the
paraneural sheath at the level of the nerve bifurcation, but outside the
epineurium of the tibial and peroneal nerves, may afford some
advantages in the block.5

Patient position/transducer: Patients should be positioned in the lateral


decubitus position with the operative side up. A linear 5-to 10-MHz
transducer is ideal for this application. Only for the heaviest patients
would a lower frequency, curvilinear transducer be necessary. For small
or thin subjects, a higher frequency linear transducer, such as 10 to 13
MHz, will provide excellent images. The transducer is held in the
transverse position, directly over the flexor crease, and is moved
proximally (Fig. 32.3). Very commonly, the image is optimized with a
significant degree of caudad tilt of the transducer. For the supine
approach, a 40-to 60-mm curved array oscillating at 3 to 5 MHz is
required. For thin patients, a small curved array oscillating at 6 to 10
MHz will suffice.

Transducer orientation: Transverse.

Needle: 10-cm, 21G (single injection) or 18G Tuohy needle (continuous


infusion).

Local anesthetic: 10 to 20 mL of 0.5% ropivacaine or 0.5% bupivacaine


with epinephrine.
Technique: Supplemental oxygen and monitors are applied to the patient.
An appropriate marking is placed on the indicated extremity for
confirmation, and a preprocedural time-out is performed at bedside.
Mild sedation is administered to the patient. An antiseptic is utilized to
prepare the skin over the block area.

Posterior popliteal approach Sciatic block at the popliteal level with


ultrasound guidance may be conducted with or without nerve
stimulator confirmation. The needle is inserted lateral to the
transducer using an in-plane approach (Fig. 32.3). One may also
utilize an out-of-plane approach, in which the transducer position is
identical, but the needle is introduced from a position superior or
inferior to the midpoint of the transducer.
The tibial nerve is located with scanning (Fig. 32.1) and traced to the
level of its union with the peroneal nerve or just proximal to this level.
The skin is anesthetized along with the subcutaneous tissues at the site
of block needle entry. The block needle is then introduced along the
predicted path to the nerve, with subtle adjustments made to the needle
and/or transducer to maintain the image. The peripheral nerve stimulator
(PNS), if utilized, is switched on as the needle is introduced or as the
needle tip approaches the nerve. After the needle is positioned next to
the nerve and the desired motor response is elicited, local anesthetic
solution in a volume sufficient to surround the nerve is injected
(typically 20 to 25 mL) (Fig. 32.3). If necessary, the needle is
repositioned during intermittent injection to completely surround the
nerve. When PNS is used to assist with ultrasound-guided sciatic block,
intraneural placement of the needle is common,6–8 if standard (less than
0.5mA) electrical threshold are utilized. Injection within the nerve will
be readily apparent if this occurs (Fig. 32.3). Some practitioners prefer
to perform the block at the midthigh. In this case, the nerve is still
imaged at the popliteal crease and then traced more proximally (Fig.
32.4).
For indwelling catheter placement, either an in-plane or out-of-plane
technique may be utilized. A sterile sheath is placed on the transducer,
and sterile gloves, gown, and drapes are utilized, along with a face mask.
After establishing an optimal image of the tibial and peroneal nerves as
they come together, or of the sciatic nerve just above this level, the skin
is anesthetized. When utilizing an in-plane approach, the technique is
similar to that described earlier. For an out-of-plane technique, the
needle should be inserted distal to the midpoint of the transducer at a
distance equal to the depth of the nerve on ultrasound imaging. The
needle is inserted at a 45-degree angle. This should bring the needle tip
to the nerve in the imaging plane of the transducer. The needle is then
guided to the nerve, and PNS is used for confirmation if desired. For
nonstimulating catheters, a bolus of local anesthetic, saline, or dextrose
solution may be utilized to confirm accurate needle tip placement as well
as catheter tip placement. When the needle has been accurately placed,
the catheter is inserted. If a stimulating catheter is utilized, the nerve
stimulator is attached to the catheter after needle placement, and the
catheter is placed using stimulation as a guide. Once catheter insertion is
begun, the transducer may be put aside or held by an assistant to image
the catheter as it is placed. When the catheter is in its desired position, it
may be imaged by injecting local anesthetic, saline, or dextrose through
it. Color flow Doppler may improve the ability to image in this situation.

Supine/lateral approach The lateral approach to sciatic block is


convenient for patients who cannot turn over, who have fractures, or
who have immobilization devices in place. Some practitioners use
this approach exclusively. Ultrasound-guided lateral block of the
sciatic nerve can be performed at a level 5 to 10 cm above the patella
or at the midthigh (Figs. 32.5 and 32.6). At the midthigh, the sciatic
nerve lies anterior to the semimembranosus and semitendinosis
muscles. It is bounded laterally by the long and short heads of the
biceps femoris muscles and medially by the adductor brevis muscle
(Fig. 32.6).
With the patient supine, the thigh and leg are internally rotated 30
degrees if this is not painful to the patient. In slim patients, the sciatic
nerve can be easily found at the lateral suprapatellar level or at the
midthigh. In heavier patients, the nerve is difficult to see at the midthigh
level. For this reason, some practitioners prefer to start the scan just
proximal to the head of the fibula and trace the common peroneal and
tibial nerves (Fig. 32.5) proximally where they join to form the sciatic
nerve (Fig. 32.6). The common peroneal and tibial nerves are imaged
several centimeters posterior to the femur. Here, the nerves usually
appear as round or oval hyperechoic structures. Toggling the transducer
is often necessary to improve visualization. Once the proper level of
block is imaged, the needle is inserted in line, superior to the probe. The
needle is directed posteriorly and medially at a 30-to 45-degree angle
until the nerve is contacted. Some practitioners prefer to use PNS to help
localize the nerve. When thresholds less than 1 mA are used, the needle
invariably penetrates the mesoneurium or paraneural sheath. Injection of
local anesthetic should proceed slowly in aliquots of 3 to 5 mL. If the
needle appears within the sheath, 10 mL of local anesthetic is sufficient.1
The needle should be withdrawn if pain or resistance occurs during
injection. If the needle lies outside the paraneural sheath, the entire
structure must be surrounded with at least 20 mL of local anesthetic.
Because of the size of the nerve and its dual covering, a long latency (30
to 60 minutes) is expected when the local anesthetic is injected around
the nerve, rather than within the paraneural sheath.

Summary of clinical evidence Perlas et al.9 and Dufuour et al.10


reported that compared to localization with landmark PNS,
utilization of ultrasound results in a higher success rate and
decreased latency of popliteal sciatic nerve blocks.
With the implementation of ultrasonography for peripheral never blocks,
several studies have demonstrated that intraneural injection of local
anesthetic is not a rare occurrence with nerve stimulation–guided
popliteal sciatic nerve blocks.6–8 In these investigations, PNS technique
was used to perform the block; ultrasound imaging of the sciatic nerve
was performed preblock and postblock to assess for nerve swelling (one
study utilized postblock computed tomography). These studies have
shown that low-current nerve stimulation–only blocks often result in
intraneural injections. Sala-Blanch et al.8 performed neurologic exams
24 hours, 1 week, and 4 weeks postblock as well as nerve conduction
studies and late response studies 1 and 4 weeks after the block. In this
small series of 17 patients, none of the patients had clinical or
electrophysiologic evidence of neurologic injury.
Multiple investigations have sought to identify techniques that would
decrease the latency of popliteal sciatic nerve blocks. In three studies,
ultrasound-guided injections were compared proximal to the bifurcation
of the sciatic nerve to injections around the individual tibial and
common peroneal nerves distal to the bifurcation. All three studies found
that blocks performed distal to the bifurcation, with separate injections
around the tibial and common peroneal division, resulted in quicker
sensory onset.11–13 Morau et al.14 and Brull et al.15 demonstrated that
circumferential spread around the sciatic nerve in the popliteal fossa can
shorten the time to complete sensory block as compared to deposition of
local anesthetic along only a portion of the circumference of the nerve.
Tran et al.5 compared postbifurcation separate injections around the
tibial and common peroneal nerves with a single injection placed within
the paraneural sheath of the sciatic nerve at the level of bifurcation.
There was no nerve swelling of either the tibial or common peroneal
component when injecting within the sheath at the bifurcation, and thus,
no evidence suggesting subepineural injection. This study exhibited that
injection within the confines of the paraneural sheath at the level of the
bifurcation resulted in a higher success rate and decreased onset time at
30 minutes compared to individual injections postbifurcation. Missair et
al.3 compared a single injection within the paraneural sheath to a
circumferential deposition, around the sheath, of the local anesthetic at
the level of the bifurcation. The investigators found that placement of
the needle tip within the paraneural sheath that encases the sciatic nerve
is associated with a greater success rate at 30 minutes, greater perineural
local anesthetic volume in contact with the nerve, and greater spread of
local anesthetic along the length of the nerve (Fig. 32.1E).

Complications: Complications related to peripheral nerve block of the


sciatic nerve at the popliteal fossa are similar to those experienced with
other peripheral blocks.16 Nerve injury and local anesthetic toxicity are
most concerning but are rare complications. Minor complications, such
as soreness at the site and localized bruising or hematoma, occur as well.
References
1. Anderson HL, Anderson SL, Tranum-Jensen J. Injection inside the
paraneural sheath of the sciatic nerve: direct comparison among
ultrasound imaging, macroscopic anatomy, and histologic analysis.
Reg Anesth Pain Med. 2012;37:410–414.
2. Sites BD, Neal JM, Chan VM. Ultrasound in regional anesthesia:
where should the “focus” be set? Reg Anesth Pain Med.
2009;34:531–533.
3. Missair A, Weisman R, Suarez MR, et al. A 3-dimensional
ultrasound study of local anesthetic spread during lateral popliteal
nerve block: what is the ideal end point for needle tip position? Reg
Anesth Pain Med. 2012;37:627–631.
4. Vloka JD, Hazdic A, Lesser JB, et al. A common epineural sheath
for the nerves in the popliteal fossa and its possible implications for
sciatic nerve block. Anesth Analg. 1997;84:387–390.
5. Tran DQH, Dugani S, Pham K, et al. A randomized comparison
between subepineural and conventional ultrasound-guided popliteal
sciatic nerve block. Reg Anesth Pain Med. 2011;36:548–552.
6. Robards C, Hadzic A, Somasundaram L, et al. Intraneural injection
with low-current stimulation during popliteal sciatic nerve block.
Anesth Analg. 2009;109:673–677.
7. Sala-Blanch X, Lopez AM, Carazo J, et al. Intraneural injection
during nerve stimulator-guided sciatic nerve block at the popliteal
fossa. Br J Anaesth. 2009;102:855–861.
8. Sala-Blanch X, Lopez AM, Pomes J, et al. No clinical evidence or
electrophysiologic evidence of nerve injury after intraneural
injection during sciatic popliteal block. Anesthesiology.
2011;115:589–595.
9. Perlas A, Brull R, Chan VW, et al. Ultrasound guidance improves
the success of sciatic nerve block at the popliteal fossa. Reg Anesth
Pain Med. 2008;33:259–265.
10. Dufour E, Quennesson P, Van Robais AL, et al. Combined
ultrasound and neurostimulation guidance for popliteal sciatic nerve
block: a prospective, randomized comparison with neurostimulation
alone. Anesth Analg. 2008;106:1553–1558.
11. Buys MJ, Arndt CD, Vagh F, et al. Ultrasound-guided sciatic nerve
block in the popliteal fossa using a lateral approach: onset time
comparing separate tibial and common peroneal nerve injections
versus injecting proximal to the bifurcation. Anesth Analg.
2010;110:635–637.
12. Prasad A, Perlas A, Ramlogan R, et al. Ultrasound-guided popliteal
block distal to sciatic nerve bifurcation shortens onset time. Reg
Anesth Pain Med. 2010;35:267–271.
13. Germain G, Levesque S, Dion N, et al. A comparison of an
injection cephalad or caudad to the division of the sciatic nerve for
ultrasound-guided popliteal block: a prospective randomized study.
Anesth Analg. 2012;114:233–235.
14. Morau D, Levy F, Bringuier S, et al. Ultrasound-guided evaluation
of the local anesthetic spread parameters required for a rapid
surgical popliteal sciatic nerve block. Reg Anesth Pain Med.
2010;35:559–564.
15. Brull R, Macfarlane AJ, Parrington SJ, et al. Is circumferential
injection advantageous for ultrasound-guided popliteal sciatic nerve
block? A proof of concept study. Reg Anesth Pain Med.
2011;36:266–270.
16. Auroy Y, Benhamou D, Bargues L, et al. Major complications of
regional anesthesia in France: the SOS regional anesthesia hotline
service. Anesthesiology. 2002;97:1274–1280.
33
Ultrasound-Guided
Ankle Block
STEPHEN M. BRENEMAN

Background and indications: The sciatic nerve gives off four branches
below the knee that must be blocked for surgery of the ankle or foot.
These are the sural nerve, the superficial and deep peroneal nerves, and
the posterior tibial nerve. In addition, the saphenous nerve, a branch of
the femoral nerve, must also be anesthetized when surgery on the ankle
or foot is planned. The superficial peroneal and sural nerves are very
small and difficult to image on ultrasound. Because they are superficial,
most practitioners prefer to block these with skin infiltration. The deep
peroneal nerve runs along the anterior surface of the tibia, deep to the
extensor retinaculum (Fig. 33.1). The posterior tibial nerve can be found
posterior to the medial malleolus near the posterior tibial artery (Fig.
33.2). The saphenous nerve can be found next to the saphenous vein,
usually posterior to the vein (Fig. 33.3). Ankle block is indicated for
surgery on the foot below the ankle.
Anatomy: The deep peroneal nerve can often be found lateral to the
anterior tibial artery. The nerve has a round or oval hyperechoic
appearance (Fig. 33.1). The posterior tibial nerve can be found posterior
to the posterior tibial artery, where it has a round hyperechoic
appearance (Fig. 33.2). The saphenous nerve is located inferolateral to
the saphenous vein. The nerve has a hyperechoic appearance (Fig. 33.3).

Patient position: Supine or recumbent with foot and ankle exposed.

Transducer: A 25-mm linear array oscillating at 10 to 13 MHz provides


an excellent image. The small transducer size makes it easy to control
the probe.
Transducer position: Deep peroneal: axially superior to anterior ankle
(Fig. 33.1A). Tibial: axially posterosuperior to the medial malleolus
(Fig. 33.2A). Saphenous: axially anterosuperior to medial malleolus
(Fig. 33.3 A).

Needle: 3-cm, 25G nerve block needle.

Local anesthetic: 2 to 5 mL per nerve of 0.5% ropivacaine or 0.5%


bupivacaine.

Technique: The skin is washed, and sterile ultrasound gel is applied to


the skin. A sterilized or covered probe is used for the procedure. The
foot is stabilized, the probe is applied, and the hand is anchored to hold
the probe steady. Use a lateral approach for the deep peroneal nerve to
place the needle by the nerve. For the tibial nerve, use a posterior
approach and infiltrate around the nerve and artery. An anterior or
posterior approach may be used for the saphenous nerve. Very little
volume is necessary to create circumferential spread. To cover the
remaining superficial peroneal and sural nerve, perform an infiltration
field block from the lateral aspect of the Achilles tendon to the medial
malleolus superior to the malleoli.

Tips 1. Use a control-top syringe to place block when working alone.


2. This is a very shallow block on a curved surface, so a smaller
ultrasound probe makes the block easier to perform.
3. Typically, only the two deep nerves associated with the arteries (deep
peroneal and tibial) are targeted directly with ultrasound, whereas the
infiltration field block covers the highly arborized superficial nerves
(i.e., saphenous, superficial peroneal, and sural).

Suggested Reading Sites B, Spence B, Beach M, et al. Ultrasound


Guidance in Regional Anesthesia: Techniques for Lower-Extremity
Nerve Blocks. New York, NY: McMahon Publishing Group; 2006.
Section IV
Truncal Blocks
34 Ultrasound-Guided Ilioinguinal and Iliohypogastric
Nerve Blocks

35 Ultrasound-Guided Transversus Abdominus Plane


Block and Quadratus Lumborum Block
Part 1: Transversus Abdominis Plane Block
Part 2: Quadratus Lumborum Block

36 Anatomy of the Paravertebral Space

37 Ultrasound-Guided Medial Paravertebral Block,


Lateral Paravertebral Block, Pectoral Block, and
Serratus Anterior Block
Part 1: Medial and Lateral Paravertebral Block
Part 2: Lateral Thoracic Paravertebral Block
Part 3: Pectoral and Serratus Anterior Block

38 Ultrasound-Guided Caudal, Lumbar, and Epidural


Injection
Part 1: Caudal Epidural Steroid Injection: A Primer
on the Combined Use of Ultrasound and
Fluoroscopy for Pain Management
Part 2: Ultrasound for Labor Epidural Placement

34
Ultrasound-Guided
Ilioinguinal and
Iliohypogastric Nerve
Blocks
MICHAEL GOFELD AND URS
EICHENBERGER

Background and indication: Local anesthetic block of the ilioinguinal


and iliohypogastric nerves may be used as the primary intraoperative
anesthetic for patients undergoing inguinal herniorrhaphy or
perioperative pain management or for patients with chronic pain in the
inguinal region. Cadaveric studies have suggested that improved
accuracy may be achieved using ultrasound techniques for blockade of
these nerves.

Anatomy: The ilioinguinal and iliohypogastric are mixed sensory and


motor nerves, formed from the ventral ramus of L1 with a lesser
contribution from T12. As such, they typically arise from the upper
portion of the lumbar plexus. Both nerves supply branches to the
abdominal musculature. After passing through these muscles, they
provide sensory innervation to the structures at their terminal locations.
Thereby, the iliohypogastric nerve provides sensory innervation to the
skin over the lower quadrant of the abdominal wall. Terminal branches
of the ilioinguinal nerve also provide sensory innervation to the lower
abdominal wall as well as the inguinal canal, the skin of the scrotum or
labium majoris, the mons pubis, and the adjacent medial aspect of the
thigh. The L1 ventral ramus enters the abdomen posterior to the medial
arcuate ligaments and courses inferolaterally over the quadratus
lumborum and posterior to the renal fascia. Near the iliac crest, it pierces
the aponeurosis of the deepest of the three muscle layers of the
abdominal wall, the transverse abdominal muscle. In this region, the
single nerve splits into the ilioinguinal and the iliohypogastric nerves.
Near the anterior superior iliac spine, both nerves pierce the middle
muscular layer of the abdominal wall, the internal abdominal oblique
muscle, as it fuses with transverse abdominis muscle. Hereon, the
ilioinguinal and the iliohypogastric nerves continue their course between
the internal and external abdominal oblique muscles, although the exact
path may be fairly variable. The iliohypogastric nerve enters this muscle
plane medially to the ilioinguinal nerve. It pierces the external oblique
aponeurosis and terminates an average 4 cm lateral of midline. The
ilioinguinal nerve begins its travel between the internal oblique and
external oblique muscles just medially to the anterior superior iliac
spine. It also courses medially, with its contributions to the lower
abdominal wall ending 3 cm lateral to the midline. A continuation of the
ilioinguinal nerve also courses inferiorly, passing through the external
inguinal ring. The mean diameter of the ilioinguinal nerve among adults
is 2.2 mm. High variation of the emergence and distribution of these
nerves has been noted. The mean diameter of the iliohypogastric nerve is
2 mm. Using ultrasound imaging, they appear as oval hypoechoic areas
with hyperechoic spots and are often in close proximity to a branch of
the deep circumflex artery. The most consistent place to perform the
block is at the level of the iliac crest where both nerves are lying
between the internal oblique and transverse muscles.

Patient position: Supine. Lateral position is a better option for obese


patients (abdominal wall will be displaced medially, decreasing needle-
to-target depth).
Probe: A broadband linear array transducer. A curvilinear transducer may
be necessary for obese individuals.

Probe position: The transducer is placed at the iliac crest and directed to
the contralateral shoulder. This will ensure insonation of the nerves in
short axis (Fig. 34.1).

Needle: 22G Quincke-type spinal needle or stimulating needle.

Local anesthetic: 2 to 5 mL of 0.5% of bupivacaine or ropivacaine.

Technique: To begin, identify surface landmarks including the iliac crest


and abdominal muscles. The skin is then prepped with sterile solution
and the transducer is covered with a sterile sheath. The transducer is
placed at the iliac crest and directed to the contralateral shoulder (Fig.
34.1). The external oblique, internal oblique, and transversus abdominis
muscles are imaged. The ilioinguinal nerve is typically visualized
between the internal oblique and transverse abdominis within 1 to 3 cm
from the iliac crest. The iliohypogastric nerve is typically identified
immediately adjacent or slightly medial to the ilioinguinal nerve (Fig.
34.2). Frequently, a branch of the deep circumflex artery is seen between
those two nerves (Fig. 34.3). The probe is positioned so the target nerve
is in the middle of the ultrasound image. The needle is inserted caudally
to the transducer and advanced out-of-plane with the ultrasound beam.
When the needle tip approximates the nerve, local anesthetic is injected
in small increments while observing distention of the tissues with each
aliquot and collection of the hypoechoic fluid encircling the target nerve.
Separation of the internal and external oblique fasciae should be clearly
seen (Fig. 34.4). When the nerves cannot be clearly identified, fascial
splitting is targeted and the nerve stimulation with the highest frequency
is performed to elicit a sensory stimulation.
Tips 1. If the nerve cannot be clearly identified, the local anesthetic can
be injected adjacent to the deep circumflex iliac artery (locate using
color Doppler imaging) which lies between the internal and transversus
abdominis muscles.
2. If both the nerve and the deep circumflex iliac artery cannot be
identified, the block can be performed by injecting local anesthetic
between the internal oblique and transversus abdominis muscles.
Observe for separation of the fascial layer.
3. If the block provides inadequate relief for patients with chronic pain
in the inguinal region, consider the possible contribution of the
genitofemoral nerve.
4. If repeated local anesthetic blockade resulted in incomplete pain
relief, consider cryoablation, surgical neurectomy, or peripheral nerve
stimulation.

Suggested Readings Eichenberger U, Greher M, Kirchmair L, et al.


Ultrasound-guided blocks of the ilioinguinal and iliohypogastric
nerve: accuracy of a selective new technique confirmed by
anatomical dissection. Br J Anaeth. 2006;97:238–243.
Gofeld M, Christakis M. Sonographically guided ilioinguinal nerve block. J Ultras Med.
2006;25:1571–1575.
Klaassen A, Marshall E, Tubbs RS, et al. Anatomy of the ilioinguinal and iliohypogastric nerves
with observations of their spinal nerve contributions. Clin Anat. 2011;24:454–461.
Moore K, Dalley A, Agur A. Clinically Oriented Anatomy. 6th ed. Philadelphia, PA: Lippincott
Williams & Wilkins; 2009.
Ndiave A, Diop M, Nydoye J, et al. Emergence and distribution of the ilioinguinal nerve in the
inguinal region: applications to the ilioinguinal anaesthetic block (about 100 dissections). Surg
Radiol Anat. 2010;32:55–62.
35
Ultrasound-Guided
Transversus
Abdominis Plane
Block and Quadratus
Lumborum Block
Part 1: Transversus Abdominis Plane
Block

KIM RUSSON AND RAFAEL


BLANCO

Background and indications: The transversus abdominis plane block,


more commonly referred to as the TAP block, places local anesthetic in
the lateral abdominal wall in a plane between the internal oblique and
the transversus abdominis muscles. Here, the local anesthetic block can
block many of the abdominal nerves as they pass to the abdominal
structures. The TAP block was first described by O’Donnell et al.1 in
2006. Their original technique was a blind technique using a blunt
regional anesthesia needle and relied on feeling a double pop as the
needle passed through the layers in an area known as the triangle of
Petit. Since then, the technique has been modified to utilize ultrasound to
confirm placement of the local anesthetic. Studies (although numbers
remain small) have shown bilateral TAP blocks to provide effective
postoperative analgesia after prostatectomy,1 large and small bowel
resection,2 and cesarean section3 compared to morphine. Intuition
suggests that if the nerves to the abdominal wall and peritoneum can be
numbed with this block, then it is likely to be useful in providing pain
relief following many other abdominal surgical, gynecological, and
urological procedures.

Anatomy: The abdominal wall between the iliac crest and the subcostal
margin consists of three layers of muscle (external oblique, internal
oblique, transversus abdominis) covered by connective tissue and skin
(Fig. 35.1). The transversus abdominis is the deepest layer, and below it
is the peritoneum. The skin, muscles, and peritoneum of the anterior
abdominal wall are innervated by the lower six thoracic nerves and the
first intercostal nerve. At the costal margin, thoracic nerves 7 to 11 leave
their intercostal spaces and enter the neurovascular plane of the
abdominal wall between transversus abdominis and internal oblique
(Fig. 35.1). Running across the surface of the transversus abdominis
muscle and aponeurosis are the lower intercostal, subcostal and
iliohypogastric nerves.
Patient position: Supine.

Probe: 40-or 60-mm curved array oscillating at 3 to 8 MHz.

Probe position: Transverse or transverse oblique between the margin of


the 12th rib and superior iliac spine.

Needle: 22G blunt needle (5 to 10 cm) for single injection; 18G Tuohy
needle (5 to 10 cm) for continuous infusions.

Local anesthetic: Bupivacaine 0.25% (20 to 30 mL), ropivacaine 0.2%


(20 to 30 mL).
Technique: This block may be performed before or following induction
of general or spinal anesthesia. The patient needs to be lying supine. An
aseptic technique is advocated using a no-touch technique, an
appropriate cleaning solution, and a sterile cover for the ultrasound
probe. The ultrasound probe is positioned horizontally on the skin just
above the iliac crest in the midaxillary line (Fig. 35.2). The muscle
layers are identified (Fig. 35.3). The authors prefer to use a peripheral
nerve block needle because it allows “distant” injection by your assistant
while you remain in control of the ultrasound probe and needle. A 50-
mm needle is usually sufficient. The authors use an in-plane approach,
inserting the needle posteriorly and directing anteriorly. The needle is
followed under direct vision as it passes through the muscle layers until
the tip lies between the internal oblique and transversus abdominis
muscles (Fig. 35.4).
Injection of the local anesthetic must be seen to ensure correct
placement. It is very characteristic to see the layer expanding in an
ellipsoid way (Fig. 35.5). The authors use 20 mL of 0.375%
levobupivacaine in each side, as do McDonnell et al.2 Use of 20 mL of
0.5% ropivacaine4 has also been described, resulting in a block from T8
to the symphysis pubis.
References
1. O’Donnell BD, McDonnell JG, McShane AJ. The transversus
abdominis plane (TAP) block in open retropubic prostatectomy.
Reg Anesth Pain Med. 2006;31:91.
2. McDonnell JG, O’Donnell B, Curley G, et al. The analgesic
efficacy of transversus abdominis plane block after abdominal
surgery: a prospective randomized controlled trial. Anesth Analg.
2007;104:193–197.
3. McDonnell JG, Curley G, Carney J, et al. The analgesic efficacy of
transversus abdominis plane block after cesarean delivery: a
randomized controlled trial. Anesth Analg. 2008;106:186–191.
4. Hebbard P, Fujiwara Y, Shibata Y, et al. Ultrasound-guided
transversus abdominis plane (TAP) block. Anaesth Intensive Care.
2007;35:616–617.
Part 2: Quadratus
Lumborum Block
ROGER F. SHERE-WOLFE, VIKRAM BANSAL,
KAREN BORETSKY, MIHAELA VISOIU, AND
PAUL E. BIGELEISEN

Background and indications: Ultrasound-guided quadratus lumborum


block was first described by Blanco in 2007.1 This technique is a
retroperitoneal fascial plane block that uses a more posterior version of
the original landmark TAP block. The exact mechanism of paravertebral
anesthesia is not clear. Other approaches have used an injection between:
1. The quadratus lumborum muscle and the posterior renal fascia (Figs.
35.6 and 35.7): Carney has shown that this results in spread along the
quadratus lumborum muscle superiorly, with extension cephalad above
the diaphragm to the thoracic paravertebral space.2
2. Into the plane between the quadratus lumborum and psoas muscles
(Fig. 35.7) with similar spread.
Thus, injection of an adequate volume of local anesthetic results in
broad retrograde thoracolumbar spread with cephalad extension to T5–
T6. Quadratus lumborum block, then, offers an alternative to TAP block,
paravertebral block, or neuraxial anesthesia. The block can be performed
in the lateral position or in the modified supine position. It is not known
whether a more cephalad injection point or use of a greater volume of
local anesthetic results in more cephalad spread, higher spread, or
hypotension from sympathetic block.
Anatomy: The quadratus lumborum muscle lies lateral to the transverse
processes of L5–T12, medial to the aponeurosis of the muscles of the
anterior abdominal wall, anterior to the erector spinae muscles, and
posterior to the psoas muscle on its medial aspect and to the posterior
renal fascia and peritoneum on its lateral aspect (Fig. 35.8). It originates
from the iliolumbar ligament and from the tips of the transverse
processes of the lower lumbar vertebrae and inserts onto the 12th rib and
the tips of the transverse processes of the upper lumber vertebrae (Fig.
35.8). The fascia of the quadratus lumborum includes the anterior and
medial lamina of the thoracolumbar fascia. These unite at the lateral tip
of the muscle with the posterior lamina from the erector spinae muscles
and become continuous with the aponeurotic origin of the internal
oblique and transversus abdominis muscles of the anterior abdominal
wall (Figs. 35.6 and 35.7). In addition, the anterior fascia of the
quadratus lumborum and the adjacent psoas muscle includes a lamina
that is continuous with the transversalis fascia, which is the inner
epimysium of the transversus abdominis muscle and which forms the
innermost investiture of the abdominal wall. The transversalis fascia and
the anterior lamina of the thoracolumbar fascia are closely apposed at
this juncture.
The lateral arcuate ligament arises from the thoracolumbar fascial
covering of the quadratus lumborum and psoas muscles at the level of
the 12th rib and forms part of the posterior origin of the diaphragm along
with the medial arcuate ligament, which arises similarly from the psoas
muscle (Fig. 35.9). The quadratus lumborum thus forms a conduit for
thoracolumbar spread both cephalad and caudad. Karmarkar et al.3 and
Saito et al.4 have independently shown that spread from the thoracic
paravertebral space caused spread via the splanchnic nerves to the
diaphragm and then beneath the arcuate ligaments to the psoas and
quadratus lumborum muscles.
The quadratus lumborum muscle runs posterior to the perinephric fat
and posterior renal fascia (Fig. 35.6). The plane between the posterior
renal fascia and the transversalis fascia is classically referred to in
anatomic and radiologic literature as the posterior pararenal space. This
space, along with the anterior pararenal and perirenal spaces, forms the
basis of classic tricompartmental theory of the retroperitoneal space and
spread of pathologic fluid collections (Fig. 35.10). Thus, extension
above the diaphragm to the thoracic paravertebral space could also be
via fascial spread between the transversalis fascia (which is continuous
with the endothoracic fascia) and the thoracolumbar fascia.
Patient position: Supine with a bump under the hip or lateral decubitus
with the operative side up.

Transducer: Both linear high-frequency (13 to 18 MHz) and curvilinear


low-frequency (2 to 5 MHz) probes may be used, depending on patient
size and the depth of the structures to be imaged.

Transducer position: Transverse superior to the iliac crest, then moved


posteriorly to provide visualization of the quadratus lumborum.

Needle: 22G, 90-to 150-mm insulated needle for single injections or an


8-to 14-cm 17 to 18G Tuohy needle for catheter insertion.

Local anesthetic: 25 to 30 mL of 0.25% bupivacaine in adults.

Technique: The transducer is placed between the iliac crest and the costal
margin in the midaxillary line. When the TAP is identified the transducer
is then moved posteriorly (Fig. 35.11) until the aponeurosis of the
internal oblique and transversus abdominis muscles comes into view,
close to the posterior lumbar triangle (Fig. 35.7). The quadratus
lumborum muscle should be posterior and medial. Ideally, the anterior
border of the quadratus lumborum muscle should be identified along
with the perinephric fat overlying the posterior renal fascia. The
injection point will be along the anterolateral edge of the quadratus
lumborum muscle, between the fascia of the muscle (which consists of
the transversalis fascia and the anterior lamina of the thoracolumbar
fascia posterior to it) and the posterior renal fascia (Fig. 35.12). Once the
appropriate anatomy has been identified, the skin is cleansed with sterile
solution, and a wheal of local anesthetic should be raised just posterior
to the transducer, with deeper infiltration as needed. The block needle is
inserted in-plane in a lateral-medial direction. Ideally, the needle should
be visualized advancing into the TAP close to the aponeurosis of the
internal oblique muscle and then advanced posteriorly through the
transversus abdominis muscle and under the lateral edge of the
quadratus lumborum (Fig. 35.12B). The needle should be flattened and
advanced slightly medially along the anterior border of the muscle,
taking care not to advance too deeply to avoid renal or peritoneal
puncture. Following negative aspiration, a few milliliters of local
anesthetic may be injected and should produce spread along an
interfascial plane between the quadratus lumborum and the posterior
renal fascia. Following appropriate identification of the proper spread,
the remainder of the local anesthetic can be carefully injected, with
intermittent negative aspiration.
Tips 1. The relevant structures are deeper more medially, posterior to the
lumbar triangle of Petit.
2. Needle localization lateral to the quadratus lumborum muscle will
produce spread along the transversalis fascial plane. There will be
some spread anterior to the quadratus lumborum muscle, but this will
produce limited cephalad spread compared to with the ideal injection
point for the quadratus lumborum block.
3. Onset of the block may be slow because of the time required for the
local anesthetic to diffuse cephalad to the lumbar nerve segments at
the level of the arcuate ligaments and to the thoracic level.

References
1. Blanco R. TAP block under ultrasound guidance: The description of
a ‘nonpopstechnique’. Reg Anaesth Pain Med. 2007;32(Suppl
1):130.
2. Carney J, Finnerty O, Rauf J, et al. Studies on the spread of local
anesthetic solution in transversus abdominus plane blocks.
Anaesthesia. 2011;66:1023–1030.
3. Karmarkar MK, Gin T, Ho AM-H. Ipsilateral thoracolumbar
anesthesia and paravertebral spread after low thoracic paravertebral
injection. Br J Anaesth. 2001;87:312–316.
4. Saito T, Den S, Tanuma K, et al. Anatomical basis for paravertebral
anesthetic block: fluid communication between the thoracic and
lumbar paravertebral regions. Surg Radiol Anat. 1999;21:359–363.

Suggested Readings Anderson JK, Cadeddu JA. Surgical anatomy


of the retroperitoneum, adrenals, kidneys and ureters. In: Wein AJ,
Kavoussi LR, Novick AC, et al., eds. Campbell-Walsh Urology. 9th ed.
Philadelphia, PA: WB Saunders; 2007:3–37.
Børglum J, Jensen K. Advances in the use of ultrasound-guided truncal blocks for perioperative
pain management. In: Derbel F, ed. Abdominal Surgery. Rijeka, Croatia: InTech; 2012:69–94.
Finnerty O, Carney J, McDonnell JG. Trunk blocks for abdominal surgery. Anaesthesia.
2010;65:76–83.
Finnerty O, McDonnell JG. Transversus abdominis plane block. Curr Opin Anaesthesiol.
2012;25:610–614.
Hebbard PD. Transversalis fascia plane block, a novel ultrasound-guided abdominal wall field
block. Can J Anaesth. 2009;56:618–620.
Miralas P, Skandalakis JE. Surgical anatomy of the retroperitoneal spaces—part I: embryogenesis
and anatomy. Am Surg. 2009;75:1091–1097.
The extraperitoneal spaces. Normal and pathologic anatomy. In: Meyers MA, Camsangavej C,
Oliphant M, eds. Meyers’ Dynamic Radiology of the Abdomen: Normal and Pathologic
Anatomy. 6th ed. New York, NY: Springer Verlag; 2011:109–202.
36
Anatomy of the
Paravertebral Space
ANNELOT KREDIET, NIZAR MAOYERI,
GERBRAND J. GROEN, AND PAUL E.
BIGELEISEN

Thoracic paravertebral (TPV) blockade produces somatic and


sympathetic blockade in adjacent unilateral dermatomes. It can be used
to provide anesthesia and/or analgesia for a variety of surgical
procedures in the thorax and abdomen and for treatment of painful
conditions such as rib fractures.1 In recent years, a growing number of
reports on ultrasound-guided TPV blocks have been published.2–10
Considering that ultrasound guidance has become a standard of care for
peripheral nerve blockade, its application for TPV blockade is an
attractive prospect. Whereas a traditional landmark-based technique is
mainly dependent on the operator’s sense of touch, ultrasound-guided
techniques offer additional information by visualizing the patient’s
anatomy, the block needle trajectory, and local anesthetic spread.
Previewing the intended insertion site can detect individual anatomical
variation (e.g., scoliosis or osteophyte formation) and define the
maximum safe depth for needle insertion. It is also valuable as a
teaching tool for demonstrating the anatomy of the TPV space.
However, these benefits are only tangible if anatomical structures can be
recognized quickly and reliably.
In this chapter, anatomical aspects relevant for ultrasonography of the
TPV space are highlighted. Correlations between high-frequency
ultrasound scans with high-resolution digitized anatomy are discussed in
depth. This enables practitioners of regional anesthesia to appraise the
spatial orientation of landmarks for TPV blockade on cross-sectional
images that correspond to ultrasound images in clinical practice.
The anatomical images were obtained by slicing blocks of tissue from a
human cadaver in transversal sections (interval: 78 µm) using a heavy-
duty sledge cryomicrotome. The surface of each section was
photographed at high resolution. Thousands of digitized photographs
were thus obtained and then digitally stacked on top of each other using
self-developed software to reconstruct the three orthogonal planes
(sagittal, coronal, and transversal). The technique is described in detail
elsewhere.11
Review of thoracic paravertebral (TPV) space anatomy The TPV space is
wedge-shaped.12 Figure 36.1 shows the TPV space in transverse, sagittal, and
coronal cross sections at the level of T7. The base of the wedge is formed by the
vertebral body, the intervertebral disk, and the intervertebral foramen medially.
Laterally, the space tapers and proceeds as the intercostal space. The anterior
boundary is formed by the parietal pleura (T2–T9/T10), the diaphragm (T9/T10–
T12), and the mediastinum. Posteriorly, the transverse process, the superior
costotransverse ligament (SCTL), and the paraspinal musculature confine the
TPV space. The SCTL runs from the superior crest of the neck of the rib toward
the anterior-inferior surface of the transverse process superior to it. The TPV
space is filled with adipose and loose connective tissue and furthermore contains
the spinal nerve, the sympathetic trunk, and the intercostal vessels. It
communicates with the epidural space medially, the intercostal space laterally,
adjacent TPV spaces cranially and caudally, the prevertebral space anteriorly,
and with adipose tissue posterior to the SCTL (Figs. 36.2 to 36.7).
Anatomical Cross Sections: Transverse paravertebral space Review of a
series of consecutive transversal cross sections from cephalad to caudal
illustrates the metameric segmentation of the transverse paravertebral
spaces. In Figure 36.2A and B, the vertebral body, the transverse process,
and the rib are situated in the same transversal plane and confine the
TPV space to a strip of adipose tissue anterior to the costovertebral joint.
Inferior to the rib (Fig. 36.3A, right side of Fig. 36.3B and C), the
transverse paravertebral space expands to a triangle between vertebral
body, transverse process, and parietal pleura. Inferior to the transverse
process, the transverse paravertebral space can be seen lateral and deep
to the lamina (Fig. 36.4A, left side of Fig. 36.4B and C). The SCTL has
an oblique orientation of 50 to 60 degrees to the skin in the transversal
plane. The pleura runs almost parallel to the skin until it curves acutely
to the mediastinum anteromedially (Fig. 36.3C).
The spinal nerve leaves the intervertebral foramen via the dorsal root
ganglion, which lies directly anterior to the facet joint. The posterior
division of the spinal nerve passes through the SCTL, and the anterior
division continues as intercostal nerve through the transverse
paravertebral space. It is joined by the intercostal artery and vein. The
sympathetic trunk lies anterolateral to the costovertebral joint and is
connected to the spinal nerve via the gray ramus communicans. The
three layers of intercostal musculature can be differentiated at
approximately 6 cm from the midline (Fig. 36.3C). More medially, the
deep muscles of the back (i.e., levator costarum, multifidi) fill the space
between transverse process, spinous process, and vertebral body. The
erector spinae muscles lie superficially.
Ultrasound images For a transverse scan of the paravertebral space, the
transducer is placed slightly lateral to the spinous process in a horizontal
position. If the ultrasound beam is directed onto the transverse process and the
rib, it produces a large acoustic shadow that masks soft-tissue structures deep to
them (Fig. 36.2). The transducer can then be moved slightly inferior and rotated
to an oblique position parallel to the rib. Directly inferior to the rib, the
ultrasound image is dominated by the “thumb”-like reflection of the transverse
process (right side of Fig. 36.3A to C). Lateral to the transverse process, the
hyperechoic curved reflection of the pleura can be identified (Fig. 36.3C).
Movement of the pleura with respiration is typically a “lung sliding” motion.
Deep to the pleura, “comet tail” reverberation artifacts may be visible that are
synchronous with respiration.13 The transverse paravertebral space lies medial
and anterior of the transverse process, and injections can be targeted at the space
between the pleura and the paraspinal/intercostal musculature.
Sliding the transducer further inferior results in a free view of the
transverse paravertebral space (left side of Fig. 36.4A to C). The
reflection of the pleura now stretches further medially and fades as the
angle of insonation becomes increasingly steep. The transverse
paravertebral space is situated just posterior to the pleura and has a
triangular shape. The posterior demarcation of the transverse
paravertebral space is formed by the paraspinal muscles and the SCTL.
Anatomical Cross Sections: Sagittal paravertebral space On a series of
consecutive sagittal cross sections (Figs. 36.5 to 36.7), the transverse
process and its corresponding rib and soft-tissue structures can be
identified. Figure 36.5A shows the intercostal space, which is a lateral
extension of the sagittal paravertebral space. The external intercostal
muscle arises from the tubercle of the rib (i.e., lateral to the tip of the
transverse process) and runs in lateral–caudad direction form rib to rib.
The internal intercostal muscle is initially a thin membrane lying directly
anterior to the external intercostal muscle. Laterally, the muscle fibers on
the intercostal muscle start to extend in a lateral–cephalad direction.
Likewise, the innermost intercostal muscle commences as a membrane
anterior to the neurovascular bundle and forms a proper muscle
approximately 5 cm lateral of the midline.
More medially, the rib and transverse processes form the costotransverse
joint (Fig. 36.6). The anterior and lateral costotransverse ligaments
connect the transverse process with the neck and tubercle of the rib of
the same thoracic level. In contrast, the SCTL stretches from a rib to the
transverse process of the thoracic level superior to it. It can be seen at
approximately 2.5 to 3.5 cm lateral of the midline. It may have more
than one layer, and gaps are occasionally seen. The intercostal nerve lies
directly anterior to the SCTL, and the intercostal vessels and the
sympathetic trunk are situated further anterior in the sagittal
paravertebral space. The anterior boundary of the sagittal paravertebral
space is formed by the parietal pleura. Posterior to the SCTL, the levator
costarum and multifidus muscles can be identified. More laterally, the
rib and transverse process begin to separate, and the ultrasound image
may show both the transverse process and rib (Fig. 36.6A,B). The
external intercostal muscle arises from the tubercle of the rib (i.e., lateral
of the tip of the transverse process) and runs in a lateral–caudal direction
from rib to rib (Fig. 36.7A,B). The internal intercostal muscle is initially
a thin membrane lying directly anterior to the external intercostal
muscle. Laterally, the muscle fibers of the intercostal muscle start to
extend in a lateral–cranial direction. Likewise, the innermost intercostal
muscle commences as a membrane anterior to the neurovascular bundle
and forms a proper muscle at approximately 5 cm lateral of the midline
(Fig. 36.7A,B).
Ultrasound images Positioning the ultrasound transducer in a paramedian
(sagittal) orientation about 4 cm lateral to the midline of the spine facilitates
visualizing the intercostal space between adjacent ribs (Fig. 36.5). With color
Doppler imaging, the identification of the intercostal artery and vein may be
seen. The contour of the reflection changes when rib and transverse process
connect at approximately 3 cm from the midline. The costotransverse joint is
characterized by a compound reflection of the transverse process (caudad and
posterior) and the rib (cephalad and anterior) as seen in Figure 36.6B. Further
medially, the rib can be completely masked by the transverse process (Fig. 36.7).
The acoustic windows between bony structures are narrow. Reflections of the
musculature, ligaments, and the pleura can be intermittently seen. Reliable
identification of the SCTL among them is challenging because its appearance is
highly dependent on the angle of insonation (Fig. 36.5D). Due to its oblique
orientation, it is often oblique (nonperpendicular) to the ultrasound beams.
Visualization may be improved by tilting the transducer laterally14 and/or
rotating it into a slightly oblique axis.15 The characteristic echogenic reflection
of the pleura may be absent when the transducer is positioned close (less than 3
cm) to the midline, particularly on the left side of the body.
Summary
Reliable interpretation of sonoanatomy can contribute to patient safety
and block success by improving the accuracy of injection.

References 1. Abrahams M, Horn J, Noles LM, et al. Evidence-


based medicine: ultrasound guidance for truncal blocks. Reg Anesth
Pain Med. 2010;35:S36–S42.
2. Hara K, Sakura S, Nomura T, et al. Ultrasound guided thoracic
paravertebral block in breast surgery. Anaesthesia. 2009;64:223–
225.
3. Shibata Y, Nishiwaki K. Ultrasound-guided intercostal approach to
thoracic paravertebral block. Anesth Analg. 2009;109:996–997.
4. Renes SH, Bruhn J, Gielen MJ, et al. In-plane ultrasound-guided
thoracic paravertebral block: a preliminary report of 36 cases with
radiologic confirmation of catheter position. Reg Anesth Pain Med.
2010;35:212–216.
5. Ben-Ari A, Moreno M, Chelly JE, et al. Ultrasound-guided
paravertebral block using an intercostal approach. Anesth Analg.
2009;109:1691–1694.
6. O Riain SC, Donnell BO, Cuffe T, et al. Thoracic paravertebral
block using real-time ultrasound guidance. Anesth Analg.
2010;110:248–251.
7. Marhofer P, Kettner SC, Hajbok L, et al. Lateral ultrasound-guided
paravertebral blockade: an anatomical-based description of a new
technique. Br J Anaesth. 2010;105:526–532.
8. Bouzinac A, Delbos A, Mazieres M, et al. Interest of ultrasound in
the realization of thoracic paravertebral block in breast surgery. Ann
Fr Anesth Reanim. 2011;30:453–455.
9. Paraskeuopoulos T, Saranteas T, Kouladouros K, et al. Thoracic
paravertebral spread using two different ultrasound-guided
intercostal injection techniques in human cadavers. Clin Anat
2010;23:840–847.
10. Bondar A, Szucs S, Iohom G. Thoracic paravertebral blockade. Med
Ultrason. 2010;12:223–227.
11. Moayeri N, Bigeleisen PE, Groen GJ. Quantitative architecture of
the brachial plexus and surrounding compartments, and their
possible significance for plexus blocks. Anesthesiology.
2008;108:299–304.
12. Eason MJ, Wyatt R. Paravertebral thoracic block—a reappraisal.
Anaesthesia. 1979;34:638–642.
13. Lichtenstein DA, Menu Y. A bedside ultrasound sign ruling out
pneumothorax in the critically ill. Lung sliding. Chest.
1995;108:1345–1348.
14. Karmakar MK. Ultrasound-guided thoracic paravertebral block.
Tech Reg Anesth Pain Manag. 2009;13:142–149.
15. Luyet C, Eichenberger U, Greif R, et al. Ultrasound-guided
paravertebral puncture and placement of catheters in human
cadavers: an imaging study. Br J Anaesth. 2009;102:534–539.
37
Ultrasound-Guided
Medial Paravertebral
Block, Lateral
Paravertebral Block,
Pectoral Block, and
Serratus Anterior
Block
Part 1: Medial and Lateral
Paravertebral Block

PAUL E. BIGELEISEN, ALON BEN-ARI,


ANDREA FANELLI, AND DANIELA ELENA
FRANCESCA GHISI

Background and indications: Thoracic paravertebral blocks have been


shown to provide unilateral or bilateral anesthesia and postoperative
analgesia comparable to thoracic epidural anesthesia and analgesia.1,2 In
addition, paravertebral blocks have decreased the recurrence of cancer in
a retrospective study.3
Thoracic paravertebral blocks are performed by injecting the local
anesthetic solution into the paravertebral space, which contains the
thoracic nerves, their branches, and the sympathetic trunk. Thoracic
paravertebral nerve blocks can be used for patients undergoing many
types of surgery and also for those with trauma pain and chronic pain.
The indications include anesthesia and postoperative analgesia for breast
surgery, video-assisted thoracoscopies, ventral hernia repairs, and
postoperative analgesia after major and minor thoracic and abdominal
procedures.4 Compared to neuraxial blocks, paravertebral blocks can be
managed in the presence of a moderate degree of hemostatic deficiency:
a careful balance of risks and benefits rather than a dogmatic and rigid
approach is suggested.5 Urinary retention and pruritus are not concerns.
There is a small risk of pneumothorax. Occasionally, injections may
enter the epidural space. The rate of absorption of local anesthetics by
the pleura is high; thus, toxicity has been described as a possible side
effect.
Several methods to reach the paravertebral space have been described.
The classic technique for the paravertebral block uses a blind approach:
The needle is inserted 2.5 to 4 cm lateral to the posterior spinous process
in search of the transverse process. When the transverse process is
contacted, the needle is withdrawn and redirected caudal to the
transverse process, approximately 1 cm deeper than the transverse
process.6–8 Once the costotransverse ligament is pierced, local anesthetic
is injected.
A loss-of-resistance technique was described by Eason and Wyatt.9 More
recently, Richardson et al.10 introduced a pressure-measurement
technique in which the correct needle tip position within the
paravertebral space is detected by different pressure values. An
autodetect syringe was used also by Mundey et al.11 for loss-of-
resistance technique. Recent advances in ultrasonography have made it
possible to place the needle in the paravertebral space under direct
ultrasound guidance.12,13
Anatomy: The thoracic paravertebral space, in sagittal cross section, is
triangular. Its boundaries include posteriorly the superior costotransverse
ligament, the transverse processes, and the ribs above and below the
intercostal nerve and vessels; medially, the boundary includes the rib and
vertebral foramen. The endothoracic fascia forms the deep border of the
space and separates the nerve roots from the sympathetic ganglia (Fig.
37.1A,B).

In most cases, the endothoracic fascia cannot be visualized on


ultrasound, and the pleura, which lies immediately deep to the
endothoracic fascia, is used as a landmark to prevent the needle from
being inserted too far. In thin patients, the epidural space, the transverse
process, the costotransverse ligament, and the pleura can be imaged in
sagittal view by placing the transducer lateral to the posterior spinous
process and toggling from medial to lateral.
Transducer: A 4-to 6-cm linear probe oscillating at 8 to 12 MHz may be
used in small patients. If the pleura is located deeper than 4 cm, a 4-to 6-
cm curved probe oscillating at 2 to 5 MHz is recommended.

Needle Single-Shot Thoracic Paravertebral Block: 21G, 9-cm, Tuohy


needle.

Continuous Thoracic Paravertebral Block: 18G, 9-cm, Tuohy needle.


Local anesthetic: Single-injection thoracic paravertebral blocks.
Anesthesia: Mepivacaine 1.5%, 3 to 5 mL for each level.
If postoperative analgesia is required: ropivacaine 0.5% or
levobupivacaine 0.375% to 0.5%, 3 to 5 mL for each level or 15 mL
injected at a single level.
Continuous thoracic paravertebral blocks Continuous infusion:
ropivacaine 0.2% or levobupivacaine 0.125%, 5 to 10 mL/h.
Programmed intermittent anesthetic bolus: ropivacaine 0.2% or
levobupivacaine 0.125%, 8 to 10 mL every 60 minutes.
Techniques Sagittal Approach: The patient is placed in a sitting position.
The relevant thoracic spinous processes are palpated starting from C7
and marked on the skin (Fig. 37.2). A point 2.5 to 3 cm lateral to the tip
of the spinous process is marked (Fig. 37.2). The region is scanned using
a 2-to 5-MHz curved array transducer placed parallel to the spinous
processes (sagittal) in search of the transverse processes, the
costotransverse ligament, and the pleura. If the pleura is visualized at a
depth of 4 cm or less, a linear probe oscillating at 8 to 12 MHz may be
used (Fig. 37.3).
Once the anatomical structures are clearly identified, the skin is
disinfected and anesthetized with 1% lidocaine using a 25G, 3.75-cm
needle at the site of introduction of the block needle. The probe is
covered by a sterile cover and repositioned in a sagittal plane in search
of the initial anatomic landmarks (Fig. 37.4). The midpoint of the
transducer is aligned midway between the two adjacent transverse
processes. A Tuohy needle, connected to an extension tube, is inserted at
the cephalad or caudal extremity of the transducer using an in-plane
technique (Fig. 37.5). The needle is directed deep to the costotransverse
ligament and superficial to the pleura (Fig. 37.6). The paravertebral
space is entered midway between the two transverse processes, avoiding
bony contact. After negative aspiration for blood, 3 to 5 mL of local
anesthetic is injected slowly into the paravertebral space. This results in
a slight anterior displacement of the pleura (Fig. 37.7). Multiple levels
may need to be injected if the local anesthetic volume is limited to 3 to 5
mL and if surgical anesthesia is required. For postoperative analgesia,
some practitioners prefer to inject a high volume of local anesthetic (10
to 15 mL) at a single site, with the assumption that it will spread to both
caudal and cephalad levels. If a continuous technique is required, once
the initial injection is completed a 20G catheter is introduced through the
18G Tuohy needle and placed 3 to 5 cm beyond the needle tip. The
needle is removed and the catheter is fixed in place using adhesive strips
and then covered with a transparent dressing. An additional 5 to 10 mL
of a long-acting local anesthetic is slowly injected through the catheter
after negative aspiration for blood.
Transversal Approach: An alternative to the classical approach uses a
transverse approach: The patient is placed in a sitting position. The
relevant thoracic spinous processes are palpated starting from C7 and
marked on the skin. A 2-to 5-MHz curved probe is positioned in a
transverse and partial oblique position to the vertebral column, parallel
to the rib, to identify the spinous process, the transverse process of the
thoracic vertebra, and the rib (Figs. 37.8 and 37.9A to C). Afterward, the
transducer is moved cranially until the parietal pleura and the external
and internal intercostal muscles are in view (Fig. 37.10A to C). A Tuohy
needle is inserted via an in-plane technique from lateral to medial until
the tip of the needle is positioned posterior from the parietal pleura in the
region where the innermost intercostal muscle is absent. After negative
aspiration for blood, 3 to 5 mL of local anesthetic is injected slowly into
the paravertebral space. Once the injection is completed, a catheter may
be inserted into the space after the space has been dilated with local
anesthetic. This lateromedial approach exposes the patient to a higher
risk of epidural spread.14
Tips 1. The paravertebral space is bisected by the endothoracic fascia,
which creates two compartments. The anterior compartment contains the
sympathetic chain, and the posterior compartment contains the
intercostal nerve, dorsal ramus, intercostal blood vessels, and rami
communicantes.
2. The paravertebral catheter can also be placed using an ultrasound-
guided intercostal approach.
3. To decrease the risk of epidural spread, the needle should not be
introduced in a lateromedial direction.
4. The use of a Tuohy needle is recommended to minimize the risk of
nerve or pleural puncture.
References
1. Joshi GP, Bonnet F, Shah R, et al. A systematic review of
randomized trials evaluating regional techniques for
postthoracotomy analgesia. Anesth Analg. 2008;107:1026–1040.
2. Davies RG, Myles PS, Graham JM. A comparison of the analgesic
efficacy and side effects of paravertebral vs. epidural blockade for
thoracotomy—a systematic review and meta-analysis of
randomized trials. Br J Anaesth. 2006;96:418–426.
3. Exadaktylos AK, Buggy DJ, Moriarty DC, et al. Can anesthetic
technique for primary breast cancer surgery affect recurrence or
metastasis? Anesthesiology. 2006;105:660–664.
4. Karmakar MK. Thoracic paravertebral block. Anesthesiology.
2001;95:771–780.
5. Chelley JE, Szczodry DM, Neuman KJ. International normalized
ratio and prothrombin time values before removal of lumbar plexus
catheter in patients receiving warfarin after total hip replacement.
Br J Anaesth. 2008;101:250–254.
6. Chelly JE. Peripheral Nerve Blocks: A Color Atlas. Philadelphia,
PA: Lippincott Williams & Wilkins; 1999.
7. Chelly JE, Uskova A, Merman R, et al. A multifactorial approach to
the factors influencing determination of paravertebral depth. Can J
Aneasth. 2008;55:587–594.
8. Greengrass R, O’Brien F, Lyerly K, et al. Paravertebral block for
breast cancer surgery. Can J Anaesth. 1996;43:858–861.
9. Eason MJ, Wyatt R. Paravertebral thoracic block—a reappraisal.
Anaesthesia. 1979;34:638–642.
10. Richardson J, Cheema SP, Hawkins J, et al. Thoracic paravertebral
space location. A new method using pressure measurement.
Anaesthesia. 1996;51:137–139.
11. Mundey DA, Buckenmaier CC III, Plunkett AR. Loss of resistance
technique for paravertebral nerve blockade using the Episure
Autodetect Syringe—a case report. Pain Med. 2009;10:854–857.
12. Cowie B, McGlade D, Ivanusic J, et al. Ultrasound-guided thoracic
paravertebral blockade: a cadaveric study. Anesth Analg.
2010;110:1735–1739.
13. Renes SH, Bruhn J, Gielen MJ, et al. In-plane ultrasound-guided
thoracic paravertebral block: a preliminary report of 36 cases with
radiologic confirmation of catheter position. Reg Anesth Pain Med.
2010;35:212–216.
14. Luyet C, Eichenberger U, Greif R, et al. Ultrasound-guided
paravertebral puncture and placement of catheters in human
cadavers: an imaging study. Br J Anaesth. 2009;102:534–539.
Part 2: Lateral Thoracic
Paravertebral Block

PAUL E. BIGELEISEN, ALON BEN-ARI,


ANDREA FANELLI, AND DANIELA ELENA
FRANCESCA GHISI

Background and indications: Ultrasound-guided thoracic paravertebral


block is indicated for procedures on the thorax that are lateral to the
paravertebral muscles, such as breast surgery and thoracotomy. When
the pleura is entered surgically, intubation is required. Breast surgery can
be done without intubation. Superficial surgery of the upper abdominal
wall can be done with bilateral thoracic paravertebral block.
The thoracic spinal nerve roots and sympathetic chain emerge from the
lateral vertebral foramina and course anterior to the transverse processes
in close proximity to the parietal pleura (Fig. 37.11). Medially, the nerve
root is bounded by the vertebral body. As the nerve root travels laterally,
it is bounded anteriorly by the endothoracic fascia and superiorly by the
inferior margin of the rib. The nerve becomes the intercostal nerve as it
enters the plane between the innermost and inner intercostal muscles
(Fig. 37.11). Posteriorly, the nerve is bounded the transverse process and
the costotransverse ligament.
Historically, thoracic paravertebral block was performed 2 to 3 cm
lateral to the midline using a blind approach. In this space, there is a
fascial plane, so local anesthetic can spread from two to six dermatomes
with a single injection. When this technique is used, there is occasional
spread to the epidural space.
Anatomy: Ultrasound has some benefits when performing thoracic
paravertebral block close to the midline. The transverse spinous process
and rib can often be visualized, and measuring their approximate depth
from the skin can be helpful. This is particularly true in obese patients,
in whom the posterior spinous process may be difficult to palpate.
Unfortunately, it is difficult to visualize the costotransverse ligament in
all but the thinnest of patients. For this reason, the authors prefer to use a
lateral subcostal approach at the angle of the rib. Here, the space
between the external and internal intercostal muscles and the space
between the internal and inner intercostal muscles can be identified.
Local anesthesia injected into either of these spaces can track back to the
paravertebral space, where it can ascend or descend several dermatomes
depending on the volume of local anesthetic injected. When this
techniques is used, the thick paraspinous muscles are avoided, so the
block is usually less painful to perform.
Patient position: Supine.
Transducer: 11-cm curved array oscillating at 5 to 8 MHz or 25-mm
linear array oscillating at 10 to 13 MHz.
Transducer orientation: Initially sagittal to identify the rib and pleura.
The transducer is then rotated 90 degrees to an oblique axial position.
Here, the rib, intercostal muscles, and pleura are identified.

Needle: 5-to 10-cm, 22G blunt needle or 18G Tuohy needle.


Local anesthetic: 10 to 15 mL of 0.5% ropivacaine.
Technique: The patient is placed prone with the arms at the side or under
the chest. The desired thoracic level is identified, and the rib is palpated
at its angle lateral to the paraspinous muscles (Fig. 37.12). The
transducer is originally oriented sagittally, and the rib and pleura are
identified. The transducer is then rotated so that it is along the long axis
of the rib. In this orientation, toggling the transducer allows the
practitioner to identify the rib and pleura. Asking the patient to breathe
deeply helps facilitate identification of the visceral and parietal pleura as
they slide past each other (Fig. 37.12). In thin patients, the fascia
between the intercostal muscles can also be identified. The needle is
introduced at the lateral border of the transducer using an in-plane
approach (Fig. 37.13). As the needle passes below the depth of the rib, 2
to 3 mL of local anesthetic is injected. If a plane appears between
muscles, this likely is the plane between the external and internal
intercostal muscles. The needle is then advanced until it lies
approximately 2 mm superficial to the pleura, at which point 3 more mL
of local anesthetic is injected. If a second tissue plane appears, this is the
plane between the internal and inner intercostal muscles (Fig. 37.13).
Once this plane has been identified, 10 to 15 more mL of local anesthetic
can be injected in aliquots of 5 mL. At the conclusion of this injection, a
catheter may be inserted 5 to 10 cm beyond the end of the needle if a
continuous block is desired (Fig. 37.14). If radiopaque dye is injected
through the catheter and a chest x-ray is obtained, then the dye usually
spreads in a spindle-like pattern in the paravertebral space (Fig. 37.15).
Tips 1. The block usually takes 20 to 30 minutes to set up.
2. Sympathetic block is common. When bilateral block is performed,
hypotension is possible.
3. Pneumothorax can occur.
4. If the space between the pleura is entered by mistake, a catheter may
be inserted into the intrapleural space and local anesthetic can be
infused to provide analgesia provided that a chest tube is not present.
Part 3: Pectoral and Serratus
Anterior Block

HANNI MONROE, RON SAMET, AND PAUL



BIGELEISEN

Introduction and indications: Block of the medial and lateral pectoral


nerves has been described in both the surgical and anesthesia literature.
The medial and lateral pectoral nerves are responsible for the motor
innervation of the pectoralis major and minor muscles. The nerves are
also thought to carry sensory fibers, although the details of the sensory
innervation are less well defined. Block of these nerves is indicated
primarily for mastectomy, breast reconstruction or reduction, and breast
augmentation.
When surgery of the chest wall is anticipated, deposition of local
anesthetic between the pectoralis minor and serratus anterior muscle, as
well as deep to the serratus anterior muscle, is indicated. Some studies
indicate that both the thoracic wall and skin over the thorax can be
anesthetized by deposition of local anesthetic via this approach.
Anatomy: The anatomic course of the medial and lateral pectoral nerves
has been the subject of multiple studies. The origins and course of these
nerves exhibit significant anatomic variability. The lateral pectoral nerve
(LPN) usually arises from either the anterior divisions of the upper and
middle trunks (33.8%) or as a single root from the lateral cord (23.4%).1
It has contributions from C5 to C7. The medial pectoral nerve (MPN)
arises most commonly from the medial cord (49.3%) or the anterior
division of the lower trunk (43.8%).1 It has contributions from C8 to T1.
The two nerves are usually connected by the ansa pectoralis just distal to
the thoracoacromial artery (Fig. 37.16).

The LPN generally supplies the upper portion of the pectoralis major
muscle.2 It crosses the axillary artery anteriorly and pierces the
coracoclavicular fascia. It then courses with the pectoral branch of the
thoracoacromial artery on the deep surface of the pectoralis major
between the muscle and its posterior fascia, eventually penetrating the
pectoralis major medial to the pectoralis minor. The course of the MPN
is more variable. The MPN innervates the pectoralis minor and the distal
portions of the pectoralis major. It arises behind the axillary artery and
passes beneath the pectoralis minor (Fig. 37.16). A number of branches
penetrate the pectoralis minor (usually at the third intercostal space close
to the midclavicular line), and often a branch wraps around the lateral
border of the pectoralis minor.3 Branches then cross the fascial plane
between the two pectoral muscles in order to penetrate and innervate the
distal portion of the pectoralis major.
The second part of the axillary artery, the axillary vein, and the cords of
the brachial plexus lie deep to the pectoralis minor muscle. The pleura
and lung are also deep to the muscles and relatively medial (Fig. 37.16).
The lateral and medial chest wall is innervated by the intercostal nerves
of T1–T12 and its branches as well as by the autonomic fibers to the
visceral and parietal pleura (Fig. 37.17). The branches to the skin over
the lateral and medial chest wall leave the intercostal nerves at several
locations proximal and distal to the midaxillary line. This deposition of
local anesthetic deep and superficial to the serratus anterior muscle at the
midaxillary line and in the deltopectoral groove usually provides
anesthesis to the lateral and anterior chest wall as well as the skin.
Patient position: Supine with the arm at the side or with the arm
abducted and elbow flexed.
Transducer: Linear 50 mm oscillating at 8 to 12 MHz.
Transducer orientation Blanco2: Parasagittal, below the clavicle, over
the pectoral muscles in the deltopectoral groove.

Needle: 5-or 10-cm, 22G, short-bevel needle.


Local anesthetic: 10 to 15 mL of 0.2% to 0.5% ropivacaine or 0.25% to
0.5% bupivacaine with epinephrine.
Technique: Patient monitors are placed, oxygen is administered, and
sedation is given as appropriate. A brief procedural “time-out” is
performed.
The patient is positioned in the supine position with the arm abducted
and elbow flexed. The anesthesiologist stands alongside the patient on
the side to be blocked, with the ultrasound on the opposite side (Fig.
37.18). A linear 50-mm, 8-to 12-MHz probe is placed in the parasagittal
plane over the pectoral muscles, below the coracoid process. The
pectoralis major and pectoralis minor muscles are identified (Fig.
37.19). This is a similar view to that obtained for the midinfraclavicular
block. The target of this approach is the plane between the pectoralis
major and pectoralis minor muscles. The local anesthetic should be
deposited in the fascial plane between the two muscles. The nerves
themselves are not usually identifiable with ultrasound. Once this is
accomplished, local anesthetic is deposited in the plane between the
pectoralis minor and the serratus anterior muscles and then deep to the
serratus anterior muscle (Fig. 37.20).
The skin at the needle insertion site is washed with a sterile solution.
The needle insertion site is just cephalad to the ultrasound probe and
caudad to the clavicle (Fig. 37.18). A 5-or 10-cm, 22G block needle is
then advanced under ultrasound guidance using an in-plane technique.
The needle should traverse the subcutaneous tissue and pectoralis major
muscle and enter the facial plane between the pectoralis major and
pectoralis minor (Fig. 37.20). With injection of local anesthetic, the
facial plane between the two muscles should split. Expansion of the
plane should be clear as further local anesthetic is injected. A total of 10
mL of local anesthetic should be injected in incremental doses, with
negative aspiration between each injection.
When surgery of the chest wall is anticipated, the needle should be
advanced deep to the pectoralis minor muscle and an additional 10 mL
of local anesthetic should be injected into the space between the
pectoralis minor and serratus anterior muscle. The needle should then be
advanced deep to the serratus muscle, where it will lie between the
serratus muscle and the external intercostal muscle (Fig. 37.20). An
additional 10 mL of local anesthetic should be injected between the
serratus muscle and chest wall/external intercostal muscle. The probe is
then moved to the intercostal space at the midaxillary line in a transverse
orientation (Fig. 37.21). The serratus muscle and rib are identified (Fig.
37.22). In some patients, the trapezius muscle may also be seen lateral
and superficial to the serratus muscle. Adipose tissue from the breast is
often seen superficial to the serratus muscle. The needle may be inserted
at the medial or lateral edge of the transducer using an inline approach.
The needle is advanced until it is just superficial to the serratus muscle,
and 10 mL of local anesthetic is injected. The needle is then advanced
deep to the serratus muscle, and an additional 10 mL of local anesthetic
is injected (Figs. 37.23 and 37.24).
Tips 1. Although the nerves are not typically visualized with ultrasound,
the LPN almost always travels with the pectoral branch of the
thoracoacromial artery, which may be seen with color Doppler.
2. This block should be performed in plane with constant needle tip
visualization because of the risk of pneumothorax. Special care
should be taken when placing local anesthetic on the posterior side of
the pectoralis minor and deep to the serratus as this is the site closest
to the pleura.
References
1. Porzionato A, Macci V, Stecco C, et al. Surgical anatomy of the
pectoral nerves and the pectoral musculature. Clin Anat.
2012;25:559–575.
2. David S, Balaguer T, Baque P, et al. The anatomy of the pectoral
nerves and its significance in breast augmentation, axillary
dissection and pectoral muscle flaps. J Plast Reconstr Aesthet Surg.
2012; 65:1193–1198.
3. Macci V, Tiengo C, Porzionato A, et al. Medial and lateral pectoral
nerves: course and branches. Clin Anat. 2007;20:157–162.
Suggested Readings
Sefa Ozel M, Ozel L, Zer Toros S, et al. Denervation point for neuromuscular blockade on lateral
pectoral nerves: a cadaver study. Surg Radiol Anat. 2011;33:105–108.
Blanco R. The ‘pecs block’: a novel technique for providing analgesia after breast surgery.
Anaesthesia. 2011;66:847–848.
Perez MF, Miguel JG, Alfaro de la Torre P. A new approach to pectoralis block. Anaesthesia.
2013; 68:430.
38
Ultrasound-Guided
Caudal, Lumbar, and
Epidural Injection
Part 1: Caudal Epidural Steroid
Injection: A Primer on the Combined
Use of Ultrasound and Fluoroscopy
for Pain Management

IMANUEL R. LERMAN, JOSEPH C. HUNG, AND



DMITRI SOUZDALNITSKI

The use of ultrasonography in pain medicine procedures allows the pain


physician to eliminate or significantly reduce exposure to the ionizing
radiation associated with fluoroscopy and computed tomography (CT)–
guided procedures.1–5 Ultrasound guidance in interventional pain
procedures provides a unique advantage over the use of fluoroscopy
alone by visualizing the advancing needle tip in real time in addition to
the surrounding nerves, vasculature, bones, tendons, and musculature.
This ability to visualize soft-tissue structures and especially the
underlying vasculature in real time is touted as one of the major
advantages of ultrasound guidance. Recent studies have shown that
inadvertent vascular transgression can be avoided when ultrasound
guidance is used in interventional pain procedures.6–10 In contrast,
fluoroscopic guidance does detect intravascular transgression but only
after the needle has entered the vessel.
However, there are clear limitations to the use of ultrasound as a sole
imaging modality. This is especially true when performing interventional
procedures near bony structures, as bony artifacts can obscure the
visualization of the advancing needle tip and shaft. Bony artifacts or
obscuration of the needle is termed bone shielding. Besides bone
shielding, there can be significant decrement in the axial resolution of an
ultrasound image when attempting deeper ultrasound-guided
interventions. Accordingly, obtaining a clear ultrasound image in deeper
tissues of an overweight or obese patients can be difficult. Without
proper visualization, accuracy of needle placement also suffers and can
result in inaccurate needle placement.11–13 The use of fluoroscopic
guidance may be more beneficial in this situation, as x-rays are able to
penetrate bone and deeper tissues without a significant loss of image
resolution, targeted structures, or the needle tip.

Sonofluoroscopic-guided caudal epidural steroid injection There are


established advantages when ultrasound guidance is used for caudal
epidural steroid injection, including decreased exposure to ionizing
radiation, visualization of the advancing needle tip in real time, and
delineation of soft-tissue anatomy. Clear visualization of the
underlying anatomy of the sacral hiatus can be especially helpful, as
this portion of the sacrum can have significant anatomical
variation.14–18
The disadvantages of using ultrasound as the sole image guidance
modality in the caudal epidural steroid injection is largely associated
with inadequate visualization of the needle tip as it advances under bony
structures, including the sacral corneae and apex of the sacral hiatus. At
this point in the procedure, fluoroscopic guidance is necessary for clear
visualization of the advancing needle tip. Likewise, it is necessary to
confirm that the needle is extravascular and extradural by the injection
of nonionic radiographic contrast under live fluoroscopy or digital
subtraction angiography. Of note, intravascular uptake has been shown
to occur at rates ranging from 9% to 40% even with the use of
fluoroscopic guidance.19–20 Moreover, needle aspiration has been shown
to be insensitive, as well as nonspecific, for detecting intravascular
needle placement when controlled with fluoroscopic digital subtraction
angiography imaging.13,20,21

Background: Caudal epidural steroid injection (via the sacral hiatus)


offers an alternative and safe entry point for the epidural space. The
caudal epidural steroid injection is considered to be of particular use in
patients who have undergone prior lumbar surgery or instrumentation, as
the prior surgery will result in a high likelihood of scarring at the site,
which can obliterate the potential epidural space.
When the landmark palpation technique is used for caudal epidural
access, failure rates range from 14% to 56%.22–25 This wide range in
success rates has been attributed to sacral anatomical variation as well as
differences in body habitus.13–18 Although ultrasound-guided caudal
epidural steroid injection is a relatively new technique (first described by
Klocke26 in 2003), recent studies have confirmed that this approach is
accurate, with success rates ranging from 96% to 100% when confirmed
with contrast-enhanced fluoroscopic imaging.13,27

Anatomy: The sacral hiatus is defined as a natural defect secondary to


incomplete fusion of the posterior midline section of the S4 and S5
vertebrae. This incomplete fusion results in a posterior opening between
sacrum and coccyx, a defect that has been termed the sacral hiatus.
The sacral hiatus is bounded posterior superficially by the
sacrococcygeal ligament, named for its location spanning between the
tip of the sacrum to the coccyx. The sacrococcygeal ligament is covered
by subcutaneous fat and skin. The rostral-to-caudad length of the
sacrococcygeal membrane has been measured in multiple studies and
averages between 22 and 26 mm.14–17 The floor of the sacral hiatus
comprises the posterior aspect of the sacrum, which can also be referred
to as the dorsal surface of the anterior sacral plate. The medial cephalad
boundary of the sacral hiatus, termed the sacral hiatus apex, is formed
by the medial sacral crest at the caudalmost tip of the posterior sacral
plate. The lateral boundaries of the sacral hiatus are formed by two
sacral corneae. The axial distance between the two sacral cornua is
approximately 10 to 17 mm.17,18 It should be noted that large anatomical
studies have consistently demonstrated the absence of the sacral hiatus
or one or both sacral cornua in 3% to 10% and 50% of patients
examined, respectively.15,17,18,28 Anatomic structures found within the
sacral hiatus include the caudal epidural canal and epidural venous
plexuses. The anteroposterior diameter of the sacral hiatus just caudal to
the apex ranges from 4.11 to 4.66 mm.15–17 Of note, the maximum
anteroposterior diameter of the sacral hiatus has been shown to be
consistently in the upper third of the sacral hiatus.16
The epidural space is contiguous from above the dens to the level of the
sacral hiatus. The sacral epidural space proximal to the sacral hiatus
contains nerve roots, spinal vessels, and the filum terminale. The dural
sac usually terminates between levels S1 and S3.16,17,22 The distance
from the apex of the sacral hiatus to the caudal edge of the S2 vertebral
level has been estimated at approximately 30 to 60 mm.16,18,29 Because
the dural sac terminates cephalad to the sacral hiatus, dural puncture
during caudal epidural steroid injection can be avoided as long as the
needle is not advanced further than 2 cm past the apex of the sacral
hiatus.

Patient position: The patient is placed in the prone position.

Transducer: A linear array 5-cm transducer oscillating between 5 and 12


MHz or other compatible transducer. A curved low-frequency transducer
can be used in obese patients.

Transducer position: The transducer positioning should be carried out in


two steps. The first position should be carried out in short-axis view, in a
transverse midline position just proximal to the sacral hiatus (Fig.
38.1A). The second transducer position is in the longitudinal or long axis
view to prepare for an in-plane approach of the needle through the
sacrococcygeal ligament (Fig. 38.1B).

Needle: 1.5-inch, 25G needle for local anesthetic; 3.5-inch, 22G or 25G
spinal needle.

Local anesthetic: 1% lidocaine.


Nonionic radiographic contrast: 1 to 1.5 mL of iohexol 180 mg/mL.
Steroid solution: 80 mg of methylprednisolone acetate diluted in 10 mL
or normal saline.
Procedure
The patient should be placed in the prone position, and the targeted area
should be cleaned with chlorhexidine and draped in a sterile fashion.
The linear array high-frequency transducer oscillating at 5 to 12 MHz is
then covered with a sterile plastic sleeve. Alternately, a low-frequency
curvilinear transducer can be used for overweight or obese patients. The
approximate position of the sacral hiatus can be localized by palpating
the sacral corneae on either side of the hiatus, usually found just superior
and slightly lateral to the gluteal cleft. The transducer is placed midline
in the short-axis, transverse position to view the sacral hiatus (Fig.
38.1A). The superficial sacral cornua appear as two hyperechoic
reversed U-shaped structures, which have been described as resembling
frogs eyes (Fig. 38.2). In between the corneae are two linear
hyperechoic structures, one superficial and one deep. The most
superficial structure is the sacrococcygeal ligament, and the deeper
structure is the dorsal bony surface of the anterior plate of the sacrum.
Between these two hyperechoic structures lies the hypoechoic area that
corresponds to the sacral hiatus (Fig. 38.2). The transducer is then
aligned so that the sacral corneae are placed just lateral to the
sacrococcygeal ligament, at which time the position of the transducers is
marked with a sterile marker (Fig. 38.1A). At this point, the transducer
is rotated 90 degrees longitudinal to the spine to obtain the long-axis
view (Fig. 38.1B). In the long-axis, longitudinal view the transducer is
positioned between the sacral corneae on either side, which can be
visualized by sliding the transducer laterally. In this view, the posterior
sacral plate cephalad to the sacral hiatus is visualized as a superficial
hyperechoic linear structure that is contiguous with the sacrococcygeal
ligament (Fig. 38.3). The two anatomic structures deep to the
sacrococcygeal ligament include the hypoechoic sacral hiatus and a
deeper hyperechoic linear line representing the dorsal surface of the
anterior plate of the sacrum (Fig. 38.3). The hyperechoic coccyx caudad
to the dorsal surface of the anterior plate is joined to the insertion of the
sacrococcygeal ligament is said to resemble a bird’s beak when
examined under ultrasound (Fig. 38.3). With the transducer held in the
longitudinal position, the entry site is anesthetized with 1 to 2 mL of 1%
lidocaine with a 25G needle. A 3.5-inch 22G or 25G needle is then
advanced in plane under real-time visualization as it pierces the
sacrococcygeal ligament (Fig. 38.4). In some patients, either the left or
right sacral cornu can cast a bony shadow, which can obscure the medial
caudal sacrococcygeal ligament as well as the advancing needle tip
(Figs. 38.5 and 38.6). However, the needle tip re-emerges as it passes
under the sacral cornu (Fig. 38.6). During this process, the angle of the
needle should be directed so that the tip is pointed at the midpoint
between the dorsal sacrococcygeal ligament and the underlying anterior
plate of the sacrum. Fluoroscopic imaging should be obtained once the
needle tip passes under the apex of the sacral hiatus (Fig. 38.7). As
mentioned previously, fluoroscopic guidance is necessary at this point
because the apex of the sacral hiatus casts a bony shadow that shields the
view of the advancing needle tip (Fig. 38.6). As the needle tip is
advanced 1 to 2 cm past the sacral hiatus apex, fluoroscopic
anteroposterior and lateral view images confirm the needle’s position in
the caudal epidural space. Iohexol 180 mg/mL (1.5 mL) is injected under
live fluoroscopy to ensure that the needle tip is extravascular and the
caudal epidural space is outlined (Fig. 38.8). Typically, a solution
containing 10 mL of 80 mg of methyl prednisolone acetate diluted in 10
mL of normal saline is injected after obtaining confirmatory fluoroscopic
contrast imaging.
Summary
The caudal epidural steroid injection is best performed with enhanced
dual image (sonofluoroscopic) guidance. There are established
advantages when ultrasound guidance is used for caudal epidural steroid
injection, including decreased exposure to ionizing radiation,
visualization of the advancing needle tip in real time, and delineation of
soft-tissue anatomy. Clear visualization of the anatomic structures in real
time can be especially helpful for this procedure because this portion of
the sacrum is known to have significant anatomical variation.15–18
However, the use of ultrasound is not without disadvantages.
Fluoroscopic guidance is needed to visualize the needle once it passes
under the apex of the sacral hiatus. In addition, contrast-enhanced
fluoroscopic imaging is needed to confirm that the needle tip is indeed
extravascular. Future studies are needed to validate the sonofluoroscopic
approach to the caudal epidural steroid injection as well as other
interventional pain procedures. Mastering the combined
sonofluoroscopic approach will allow the pain physician a unique
opportunity to combine and coordinate the respective advantages of both
ultrasound and fluoroscopic guidance for interventional pain procedures.

References 1. The 2007 Recommendations of the International


Commission on Radiological Protection. ICRP publication 103. Ann
ICRP. 2007;37:1–332.
2. National Council on Radiation Protection and Measurements.
Limitation of Exposure to Ionizing Radiation. Bethesda, MD:
National Council on Radiation Protection and Measurements; 1993.
Tech. Rep. 116.
3. Ainsbury E, Bouffler S, Dörr W, et al. Radiation cataractogenesis: a
review of recent studies. Radiat Res. 2009;172:1–9.
4. Fishman SM, Smith H, Meleger A, et al. Radiation safety in pain
medicine. Reg Anesth Pain Med. 2002;27:296–305.
5. Fish D, Kim A, Ornelas C, et al. The risk of radiation exposure to
the eyes of the interventional pain physician. Radiol Res Pract.
2011;609:537–609.
6. Narouze S, Vydyanathan A. Ultrasound-guided cervical
transforaminal injection and selective nerve root block. Tech Reg
Anesth Pain Manag. 2009;13:137–141.
7. Narouze S, Vydyanathan A, Kapural L, et al. Ultrasound-guided
cervical selective nerve root block: a fluoroscopy-controlled
feasibility study. Reg Anesth Pain Med. 2009;34:343–348.
8. Jee H, Lee J, Kim J, et al. Ultrasound-guided selective nerve root
block versus fluoroscopy-guided transforaminal block for the
treatment of radicular pain in the lower cervical spine: a
randomized, blinded, controlled study. Skeletal Radiol.
2013;42:69–78.
9. Finlayson RJ, Gupta G, Alhujairi M, et al. Cervical medial branch
block: a novel technique using ultrasound guidance. Reg Anesth
Pain Med. 2012;37:219–223.
10. Narouze S. Beware of the “Serpentine” inferior thyroid artery while
performing stellate ganglion block. Anesth Analg. 2009;109:289–
290.
11. Rauch S, Kasuya Y, Turan A, et al. Ultrasound-guided lumbar
medial branch block in obese patients: a fluoroscopically confirmed
clinical feasibility study. Reg Anesth Pain Med. 2009;34:340–342.
12. Galiano K, Obwegeser AA, Walch C, et al. Ultrasound-guided
versus computed tomography-controlled facet joint injections in the
lumbar spine: a prospective randomized clinical trial. Reg Anesth
Pain Med. 2007;32:317–322.
13. Blanchais A, Le Goff B, Guillot P, et al. Feasibility and safety of
ultrasound-guided caudal epidural glucocorticoid injections. Joint
Bone Spine. 2010;77:440–444.
14. Trotter M. Variations of the sacral canal: their significance in the
administration of caudal analgesia. Anesth Analg. 1947;26:192–
202.
15. Nagar SK. A study of sacral hiatus in dry human sacra. J Anat Soc
India. 2004;53:18–21.
16. Crighton IM, Barry BP, Hobbs GJ. A study of the anatomy of the
caudal space using magnetic resonance imaging. Br J Anaesth.
1997;78:391–395.
17. Sekiguchi M, Yabuki S, Satoh K, et al. An anatomic study of the
sacral hiatus: a basis for successful caudal epidural block. Clin J
Pain. 2004;20:51–54.
18. Senoglu N, Senoglu M, Oksuz H, et al. Landmarks of the sacral
hiatus for caudal epidural block: an anatomical study. Br J Anaesth.
2005;95:692–695.
19. Ergin A, Yanarates O, Sizlan A, et al. Accuracy of caudal epidural
injection: the importance of real-time imaging. Pain Practice.
2005;5:251–254.
20. Renfrew DL, Moore TE, Kathol MH, et al. Correct placement of
epidural steroid injections: fluoroscopic guidance and contrast
administration. AJNR Am J Neuroradiol. 1991;12:1003–1007.
21. Eastwood D, Williams C, Buchan I. Caudal epidurals: the whoosh
test. Anaesthesia. 1998;53:296–307.
22. Price CM, Rogers PD, Prosser AS, et al. Comparison of the caudal
and lumbar approaches to the epidural space. Ann Rheum Dis.
2000;59:879–882.
23. Stitz MY, Sommer HM. Accuracy of blind versus fluoroscopically
guided caudal epidural injection. Spine (Phila Pa 1976).
1999;24:1371–1376.
24. White AH, Derby R, Wynne G. Epidural injections for the diagnosis
and treatment of low-back pain. Spine (Phila Pa 1976). 1980;5:78–
86.
25. El-Khoury GY, Ehara S, Weinstein JN, et al. Epidural steroid
injection: a procedure ideally performed with fluoroscopic control.
Radiology. 1988;168:554–557.
26. Klocke R, Jenkinson T, Glew D. Sonographically guided caudal
epidural steroid injections. J Ultrasound Med. 2003;22:1229–1232.
27. Chen C, Tang S, Hsu T, et al. Ultrasound guidance in caudal
epidural needle placement. Anesthesiology. 2004;101:181–184.
28. Black MG. Anatomic reasons for caudal anesthesia failure. Anesth
Analg (Cleve.). 1949;28:33–39.
29. Aggarwal A, Harjeet D, Sahni. Morphometry of sacral hiatus and its
clinical relevance in caudal epidural block. Surg Radiol Anat.
2009;31:793–800.
Part 2: Ultrasound for Labor
Epidural Placement

MANUEL C. VALLEJO AND MANASI BADVE

Accidental dural puncture (ADP) is a common complication of epidural


insertion for labor analgesia, with a reported incidence of 1% to 5%.
Resultant postdural puncture headache following ADP can be very
distressing and extremely disabling. A failed labor epidural is also
challenging because the parturient has inadequate analgesia and requires
either another neuraxial block for analgesia or a general anesthetic to
provide anesthesia, with its inherent complications consisting of difficult
ventilation and intubation, apnea, and aspiration, all of which can
increase maternal morbidity and mortality.
The first report of ultrasound-guided lumbar puncture dates back to 1971
in a Russian publication.1 Cork et al. described neuraxial anatomy as
well as advantages and disadvantages of ultrasound-guided localization
of epidural space in 1980.2 With the improvement in quality of
ultrasound images and availability of low-frequency probes, a number of
anesthesia-related studies have been conducted and published. The
National Institute for Health and Clinical Excellence published
guidelines for ultrasound-guided catheterization of epidural space in
2008.3
Ultrasound assistance for epidural placement can calculate the distance
from the skin to the epidural space; determine midline; provide an
estimation of the proper insertion angle and the best epidural insertion
point, especially in patients with difficult or altered anatomy (i.e.,
scoliosis, obesity, decompressive neuraxial surgery and instrumentation);
and ultimately decrease the failed epidural placement rate.

Applied anatomy: The first stage of labor begins with maternal


perception of regular, painful uterine contractions and ends with
complete cervical dilatation. Analgesia for the first stage of labor can be
achieved by blocking segments from the tenth thoracic dermatome to the
fifth lumbar dermatome (T10–L5). The second stage of labor begins
with complete cervical dilatation and ends with delivery of the neonate.
Analgesia for the second stage of labor can be achieved by blocking the
pudendal nerve distribution; sacral nerves 1 through 5 (S1–S5).
Labor epidural insertion landmarks include traversing the skin followed
by the supraspinous ligament and then the interspinous ligament
followed by the ligamentum flavum. Entry into the epidural space is
confirmed with loss of resistance to air or saline, and the epidural
catheter is inserted 3 to 5 cm into the epidural space and then secured
with adhesive dressing and tape.

Patient position: Sitting or lateral position, flexed forward, and curled


inward.

Indications: Labor analgesia/anesthesia for the first and second stage of


labor.

Needle size: 17G epidural Tuohy needle.

Transducer position: Both the longitudinal (long-axis or sagittal) view


and the transverse (perpendicular) view can be used to identify the
epidural space (ligamentum flavum/posterior dura) at the L3–L4, or L4–
L5 intervertebral space.

Ultrasonographic landmarks: Identifiable anatomy using the


longitudinal approach includes the sacrum, articular process,
ligamentum flavum and posterior dural mater, anterior dura mater,
posterior longitudinal ligament, and vertebral body (Fig. 38.9).
Identifiable anatomy using the transverse approach include the spinous
process, articular process, transverse process, ligamentum flavum and
posterior dural mater, anterior dura mater, posterior longitudinal
ligament, and vertebral body (Fig. 38.10).
Block technique: For the longitudinal approach, the ultrasound scan is
performed by positioning the ultrasound probe vertically, perpendicular
to the long axis of the spine. A “saw” sign will be visualized (Fig.
38.11). The saw represents the spinous processes (saw teeth), and the
interspinous spaces (groves of the saw) represent the ligamentum
flavum/posterior longitudinal ligament and vertebral body (Fig. 38.11).
For the transverse approach, the probe is placed horizontally,
perpendicular to the long axis of the spine in the intervertebral space. In
the midline, the spinous process is identified as a small hyperechoic
signal immediately underneath the skin, which continues as a long
triangular hypoechoic (dark) acoustic shadow (Fig. 38.12). The probe is
then moved slightly cephalad or caudad to capture a view of an acoustic
window (vertebral interspace). On the midline in thin patients, and
within the vertebral interspace, an equal sign (“=”) or a “flying bat” sign
may be seen (Fig. 38.12). A hyperechoic band corresponding to the
ligamentum flavum and the dorsal dura is visualized that corresponds to
the upper most part of the equal sign. A second hyperechoic band,
parallel and beneath the first part of the equal sign, corresponds to the
anterior dura, the posterior longitudinal ligament, and the vertebral body
(Fig. 38.12). In thin patients, the anterior and posterior epidural spaces
(Fig. 38.12), transverse processes, and facet joints can also be seen (Fig.
38.13).
The exact level of each of the interspaces (L3–S1) can be marked on the
skin to facilitate epidural placement, and both views can be used as
points of reference to determine midline and the optimal vertebral
interspace for epidural puncture and insertion. Once the clear image of
the interspace is obtained, the image is frozen and the distance from the
skin to the epidural space can be measured using the built-in caliper
equipped on the ultrasound machine.

Tips • Both the longitudinal and transverse approaches should be used to


visualize the epidural space, as both views provide two points of
reference.
• Of the two approaches, the transverse approach provides the most
reliable distance to the epidural space.
• The ultrasound probe angle that produced the best visualization of the
epidural space is the angle that should be used for epidural needle
placement.

Ultrasound-assisted neuraxial technique: The failure rate for labor


epidural analgesia ranges from 1% to 5%. A preprocedure ultrasound
examination to determine the midline, needle direction, and epidural
space depth before labor epidural placement by resident trainees has
been shown to decrease the number of puncture attempts and epidural
failure rate when compared to traditional blind techniques. This is of
immense value in teaching institutions, as it increases the margin of
safety and facilitates the learning curve.
A preprocedure ultrasound examination is performed to establish the
neuraxial anatomy, locate the target lumbar interspace, the midline, and
the epidural space depth from the skin. Epidural placement is then
carried out using the traditional loss-of-resistance technique. This is
more commonly performed as compared to real-time ultrasonography.

Real-time ultrasound-guided neuraxial technique This consists of


actual ongoing imaging of the lumbar region during epidural
puncture. It can be performed as an out-of-plane technique where
ultrasound scanning is performed in the paramedian longitudinal
plane and a Tuohy needle is inserted in the midline. The
disadvantages are poor demarcation of structures and the need for
an assistant to help with imaging. Real-time ultrasound-guided
epidural insertion is not commonly performed, and there are very
few studies describing the same.
Role in presumed difficult cases In obese parturients, the placement
of labor epidural catheters is technically more difficult due to
obscure anatomical landmarks, and it is associated with multiple
puncture attempts as well as an increased risk of an ADP. The initial
failure rate for epidural catheter can be as high as 42% requiring
catheter replacement. Additionally, due to increased depth to the
epidural space in this subset of patients, a small change of angle at
the skin surface can result in a significant change in the epidural
needle trajectory. Therefore, careful visualization of the transducer
angle is important. In order to aid visualization through excessive
adipose tissue, the resolution and/or penetrance on the ultrasound
machine can be altered to enhance visualization.
Patients with scoliosis have axial rotation of the vertebral bodies and
angulation of spinous processes. A preprocedure ultrasound study of the
spine can help to estimate the degree of rotation and provide guidance
for needle insertion. Several interspaces should be scanned, and the best
space with normal sonoanatomy selected for the neuraxial block.
In patients with spinal instrumentation, surgical correction of bone, bone
grafting, and metal implants can distort anatomical as well as
sonographic landmarks. A preprocedure ultrasound can identify the best
interspace and provide guidance for neuraxial block placement.
References
1. Bogin IN, Stulin ID. Application of the method of 2-dimen-sional
echospondylography for determining landmarks in lumbar
punctures. Zh Nevropatol Psikhiatr Im S S Korsakova.
1971;71:1810 –1811.
2. Cork RC, Kryc JJ, Vaughan RW. Ultrasonic localization of the
lumbar epidural space. Anesthesiology. 1980;52:513–516.
3. National Institute for Health and Clinical Excellence. Ultrasound-
guided catheterisation of the epidural space. London: National
Institute for Health and Clinical Excellence 2008; ISBN 1-84629-
583-1. Available at: https://www.nice.org.uk/guidance/ipg249
(Accessed on October 7, 2014).

Suggested Readings Balki M. Locating the epidural space in


obstetric patients—ultrasound a useful tool: continuing professional
development. Can J Anesth. 2010;57:1111–1126.
Balki M, Lee Y, Halpern S, et al. Ultrasound imaging of the lumbar spine in the transverse plane:
the correlation between estimated and actual depth to the epidural space in obese parturients.
Anesth Analg. 2009;108:1876–1881.
Baraz R, Collis RE. The management of accidental dural puncture during labour epidural
analgesia: a survey of UK practice. Anesthesia. 2005;60:673–679.
Carvalho JC. Ultrasound-facilitated epidurals and spinals in obstetrics. Anesthesiol Clin.
2008;26:145–158.
Grau T, Leipold RW, Horter J, et al. The lumbar epidural space in pregnancy: visualization by
ultrasonography. Br J Anaesth. 2001;86:798–804.
Karmakar MK, Li X, Ho AM, et al. Real-time ultrasound-guided paramedian epidural access:
evaluation of a novel in-plane technique. Br J Anaesth. 2009;102:845–854.
Vallejo MC. Anesthetic management of the morbidly obese parturient. Curr Opin Anaesthesiol.
2007;20:175–180.
Vallejo MC, Phelps AL, Singh S, et al. Ultrasound decreases the failed labor epidural rate in
resident trainees. Int J Obstet Anesth. 2010;19:373–378.
Section V
Pediatric Regional Anesthesia
39 Fundamentals of Ultrasound-Guided Pediatric
Regional Anesthesia

40 Ultrasound-Guided Brachial Plexus Block in


Infants and Children

41 Ultrasound-Guided Lower Extremity Blockade in


Children

42 Ultrasound-Guided Epidural Block (Caudal,


Lumbar, and Thoracic), Truncal and Paravertebral
Blocks in Children
Part 1: Introduction
Part 2: Single-Injection Caudal Approach to the
Epidural Space
Part 3: Lumbar and Thoracic Epidural Catheters
via the Caudal Approach
Part 4: Lumbar and Thoracic Continuous Epidural
Analgesia via the Direct, Intervertebral
Approach
Part 5: Ultrasound-Guided Regional Anesthesia of
the Thorax, Trunk, and Abdomen in Infants
and Children


39
Fundamentals of
Ultrasound-Guided
Pediatric Regional
Anesthesia
TARUN BHALLA AND JOSEPH D. TOBIAS

All children have beautiful (Fig. 39.1) anatomy, which makes pediatric
regional anesthesia especially enjoyable. Although pediatric regional
anesthesia was historically preceded by descriptions in the literature of
adult-based regional anesthesia, Dr. H. Tyrell Gray in London reported
and discussed the use of spinal anesthesia in children in 1909.1 Gray
concluded that the benefits of regional anesthesia in the pediatric
population included “absolute anesthesia, no surgical shock, localized
analgesia to the area of the block, and minimal postoperative vomiting.”
In the 1930s and 1940s, pediatric regional anesthesia did not receive
significant press or popularity in the anesthesia literature, likely due to
advancement in the pharmacology involved with general anesthesia,
including tubocurarine, thiopentone, and cyclopropane.2 By the 1960s,
pediatric caudal epidural techniques were being described in both the
literature and textbooks as a reliable method to provide surgical
anesthesia.3,4 The early history of regional anesthesia in infants and
children dealt primarily with neuraxial techniques, including spinal and
epidural anesthesia. Although first described in the 1980s, peripheral and
truncal blockade in children has also become more commonplace within
the past 10 years with the introduction of ultrasound and the
development of pediatric-appropriate equipment.

Given these advancements, the last decade has witnessed an increased


use of regional anesthesia in the perioperative period in infants and
children of all ages. Although regional anesthetic techniques are used
most commonly as an adjunct to general anesthesia to either supplement
general anesthesia or to provide postoperative analgesia, many of these
regional anesthetic techniques can also be used instead of general
anesthesia in circumstances where anatomic or physiologic alterations
may make the conduct of general anesthesia more difficult or dangerous.
Furthermore, with the recent information regarding apoptosis and an
increased awareness regarding the potential neurotoxic effects of
anesthetic agents on the developing brain, there has been a resurgence in
interest in using spinal anesthesia for short surgical procedures such as
herniorrhaphy in neonates and infants.5 Alternatively, as in the adult
population, regional anesthetic techniques in conjunction with sedation
is viewed as an equal alternative to general anesthesia in some centers.
Regional anesthetic techniques may also be used during painful or
invasive procedures in situations where the conduct of general
anesthesia may not be readily possible, as a therapeutic modality to
provide sympathetic blockade in patients with vascular insufficiency,
and in the treatment of acute and chronic pain of various etiologies.
Regardless of the situation, appropriate training, the correct equipment,
and accurate placement are mandatory to ensure the success and safety
of the technique.
In pediatric-aged patients, caudal epidural anesthesia remains the most
frequently performed regional anesthetic technique.6 Although this
technique is effective for lower abdominal and lower extremity
procedure, other approaches are obviously necessary when the operative
procedure or pain process involves the upper extremity. For unilateral
extremity procedures, there has been increased use of peripheral nerve
blockade to allow for unilateral anesthesia and avoid the potential
adverse effects of neuraxial techniques, including bilateral lower
extremity motor blockade, urinary retention, and sympathectomy. The
potential advantages of peripheral block are illustrated by two surveys of
regional anesthetic techniques from the French Language Society of
Paediatric Anaesthesiologists, commonly known as Association des
Anesthesistes Reanimateurs Pediatriques d’Expression Francaise
(ADARPEF).7,8 In its first survey, central neuraxial blockade (mostly
caudal epidural blockade) accounted for 60% of the 24,409 regional
anesthetic techniques. There were 4,090 peripheral nerve blocks, 997 of
which were placed in the upper extremity. No complications were noted
related to the peripheral nerve blocks, whereas the complication rate
related to caudal epidural block was 1.5%. These results prompted the
authors to suggest that “the extremely low incidence of complications
(zero in the study) after peripheral nerve blocks should encourage
anesthesiologists to use them more often, when they are appropriate, in
the place of a central block.” A follow-up study, published in 2012 from
the ADARPEF, noted similar findings.8 The 12-month data collected
from 47 institutions included 29,870 regional anesthetic techniques used
in association with general anesthesia and 1,262 sole regional blocks.
Central blocks accounted for 34% of all of the techniques as compared
to 60% in the first survey. In children aged ≤3 years, the percentage of
central blocks was similar to the peripheral ones (45% versus 55%). In
older patients, peripheral blocks were four times more common than
central techniques. The second survey also showed an increased
occurrence of peripheral blockade of the face and trunk. Complications
were rare, occurring in only 40 patients for an overall complication rate
of 0.12%, and did not result in any sequelae. The complication rate was
six times higher for central than for peripheral blocks. The authors
continued to encourage the use of peripheral instead of central blocks,
including caudal whenever appropriate.
The findings of the ADARPEF group and the safety of regional
anesthesia have been echoed more recently with the report from the
Pediatric Regional Anesthesia Network (PRAN).9 PRAN is a multi-
institutional study focusing of the use and incidence of complications of
pediatric regional anesthesia. A total of 14,917 regional blocks,
performed on 13,725 patients, were accrued from April 1, 2007 through
March 31, 2010. There were no deaths or complications with sequelae
lasting greater than 3 months related to regional anesthesia. Single-
injection blocks had fewer adverse events than continuous blocks,
although the most frequent events (33% of all events) in the latter group
were catheter-related problems. Ninety-five percent of blocks were
placed while patients were anesthetized. Although single-injection
caudal blocks remained the most frequently performed (40%), peripheral
nerve blocks were also frequently used (35%), possibly driven by the
widespread use of ultrasound (83% of upper extremity and 69% of lower
extremity blocks). The conclusion stated that “regional anesthesia in
children as commonly performed in the United States has a very low rate
of complications, comparable to that seen in the large multicenter
European studies.”8
The ongoing research regarding the potential neurotoxicity of general
anesthetic agents on the developing brain has also prompted a further
interest in regional anesthesia in the pediatric population. In the recent
literature, gamma-aminobutyric acid antagonists and N-methyl-D-
aspartate agonists have been shown to accelerate apoptosis in the
immature brain, with concerns expressed regarding eventual
neurocognitive outcomes.10–12 These concerns have led to a renewed
interest in the use of regional anesthesia instead of general anesthesia or
at least combining a regional anesthetic technique with general
anesthesia as a means of limiting exposure to pro-apoptotic agents.
Certainly, further investigation needs to be performed before any
definitive conclusions can be drawn regarding these issues.
Ongoing advancements in technique and refinement of the available
equipment have improved the applicability of regional anesthetic
techniques in the pediatric population. The use of ultrasound-guided
blockade has gained widespread applicability in the adult population.
Given its ease of use, noninvasive modality, and potential to improve the
success rate and decrease the incidence of adverse effects, ultrasound-
guided regional anesthesia has increased in the pediatric population.13–15
Of the many differences between the practices of adult and pediatric
regional anesthesia, the major one is that in the pediatric-aged patient,
these techniques are performed in the anesthetized or deeply sedated
patient.9,16 This practice is well within the standard of care for infants
and children and, as such, a means of ensuring correct placement of the
needle is mandatory. This has been demonstrated in the more than
40,000 blocks reported in the publications of Giaufre and Polaner.7–9
The following chapter reviews basic information regarding the use of
ultrasound to guide regional anesthesia procedures in infants and
children, discusses some of the appropriate equipment needed for such a
practice, and reviews issues related to local anesthetic agents in pediatric
patients, including dosing regimens, test dosing to identify inadvertent
intravascular injection, local anesthetic toxicity, and its treatment.
Equipment for regional anesthesia in infants and children Regional anesthesia
equipment has advanced significantly over the past 20 years; however, the
preparation and setup have remained the same. Consent from the guardian and,
where appropriate, assent from the patient are obtained during the preoperative
interview. The purpose and expected outcomes of the regional anesthetic
technique are discussed. When motor blockade can be expected along with
sensory blockade, the patient and parent should be informed of this. The
procedure should be performed in a designated area, such as an induction room
or operating room. An appropriate preprocedure “time-out” should be performed
to ensure the appropriate patient, block, and site. The time-out should include a
preprocedure marking of the block site. Supportive treatment options should be
readily available, including oxygen, monitoring, and airway equipment including
medications to treat adverse effects related to local anesthetic agents, including
seizures and cardiovascular effects (hypotension or arrhythmias). Ready access
to rescue lipid emulsion to treat local anesthetic toxicity is suggested.17,18
Monitoring equipment includes those standards as outlined for intraoperative
monitoring by the American Society of Anesthesiologists, including pulse
oximetry, noninvasive blood pressure cuff, electrocardiogram (ECG), and
respiratory sensors. Most pediatric patients have undergone induction of general
anesthesia, and therefore, capnography and temperature monitoring are also
utilized. When regional anesthetic techniques are performed using sedation,
capnography measured from the nasal cannula during spontaneous ventilation is
also suggested.
Ultrasound: Ultrasound to guide peripheral blockade in children has great
utility, as the nerves or tissue planes are superficial and easily imaged.
Regional anesthetic techniques in children are often performed under
deep sedation or general anesthesia. In this setting, the ability to identify
the nerve and the needle tip may prevent adverse effects of these
techniques, including damage to or injection within nerves, puncture of
vascular or visceral structures, or pneumothorax. In smaller patients, the
safe dose of local anesthetic is often a limiting factor in the successful
performance of peripheral nerve blockade. Ultrasound-guided block is
an asset in this setting, as it has been shown to speed the onset of the
block, improve its efficacy, and—most importantly—diminish the dose
of local anesthetic agent that is necessary to provide a successful
block.19–21 With such accuracy and efficacy, ultrasound-guided
peripheral nerve blockade in children can now be used instead of caudal
epidural blockade, which was previously used for surgery of the lower
extremities and abdomen. Although neuraxial blockade (caudal and
lumbar epidural analgesia) in children is efficacious, adverse effects
including urinary retention, as well as bilateral sensory blockade and
motor weakness, may occur. If hydrophilic opioids such as morphine or
hydromorphone are added to neuraxial blockade, additional adverse
effects including nausea, vomiting, pruritus, and respiratory depression,
may occur. Ultrasound-guided lower extremity blocks, truncal blocks
(transversus abdominis, rectus sheath and ilioinguinal nerve blocks), and
paravertebral/intercostal blocks have started replacing many of the
neuraxial techniques previously used in pediatric patients.
The true echogenicity of a nerve is demonstrated only when the
ultrasound beam is directed perpendicular to the nerve’s axis.
Consequently, linear probes with parallel sound beam emission are used
for ultrasound-guided regional anesthesia. Using a transverse scanning
technique, peripheral nerves generally appear as dark structures
(hypoechoic) above the clavicle and as hyperechoic (white or bright)
structures below the clavicle. These differences in the appearance of the
nerves on ultrasound depend on the size of the nerve, the frequency of
the ultrasound, and the angle at which the ultrasound beam contacts the
nerve. Given these issues, the nerves may appear as round or oval
hypoechoic structures or as a single hyperechoic structure. With some
techniques (transversus abdominis, rectus sheath or fascia iliaca block),
the nerves are not individually identified, but rather ultrasound is used to
identify the appropriate fascial plane for hydrodissection and deposition
of the local anesthetic agent. Once hydrodissection is started, the nerve
may then be visualized within the local anesthetic solution. The needle
(see following text) can be inserted in line with the transducer so that the
needle and its tip can be visualized or at a 90-degree angle (cross-
sectional technique) to the transducer. With the latter technique, only the
tip of the needle can be seen and so the course of the needle is identified
by tissue displacement.
As with adults, various probes and transducers can be used for the
performance of regional anesthesia in infants and children, the choice
being based on the patient’s weight and the type of bock that is being
performed.22,23 There is an inverse relationship between frequency and
depth of penetration, so higher frequency probes provide better
resolution of superficial structures and can generally be used in smaller
pediatric patients given the proximity of the structures to the skin. In
general, three types of probes provide most of what is needed for
ultrasound-guided regional anesthesia in infants and children.
1. A 25-mm (hockey-stick), high-frequency (10 to 14 MHz) linear
transducer is suitable for the performance of many regional blocks in
children who weigh less than 20 kg. This probe is ideal for imaging
structures within 6 cm from the surface of skin and has a small
footprint.
2. A 35-mm linear probe with a frequency range of 6 to 13 MHz is
useful in children who weigh more than 20 kg. This probe is ideal for
imaging structures within 6 cm of the surface of the skin with a larger
footprint.
3. A curvilinear (35 to 55 mm) low-frequency probe (2 to 5 MHz) may
be useful for blocks in which more depth of penetration is necessary,
such as lumbar plexus blockade or neuraxial techniques in larger
patients. This probe can be used in larger children and adolescents to
survey surrounding anatomical structures such as muscles, bones,
and the vasculature to identify deeplying structures such as the sciatic
nerve.

Needles: As children come in all sizes, ages, and shapes, various needles
of different gauges and lengths are necessary for the successful use of
regional anesthesia in the pediatric patient. When selecting the
appropriate needle, an additional advantage of ultrasound is that the
depth of the structures to be anesthetized from the skin surface can be
estimated and thereby help in the choice of the length of the needle to be
used. In our practice, we have access to various lengths and gauges of
insulated needles, including 20G (6-inch), 21G (4-inch), 22G (2-inch),
and 24G (1-inch) needles for single-shot techniques. Additional
equipment for the placement of catheters for continuous postoperative
infusions is also available in various sizes for the pediatric population.
The majority of these catheters, which are 20G, are placed through an
18G insulated Tuohy needle. Various lengths of insulated 18G Tuohy
needles are available including 4-, 2-, and 1.5-inch varieties.

Nerve Stimulator: Although ultrasound-guided blockade provides the


greatest likelihood of safety and success with regional anesthesia, the
use of surface landmarks and a nerve stimulator are still practiced in
many centers, especially during the transition to using ultrasound.
Although some argue that the fusion of nerve stimulation and ultrasound
guidance is unnecessary, our experience has suggested that until
expertise with ultrasound has been established, nerve stimulation
provides additional assurance that the needle is in the correct location.
As the experience and use of ultrasound has increased at our institutions,
nerve stimulation is rarely if ever used.
The traditional practice is to set the nerve stimulator at 1.5 to 2.0 mA to
initially identify the nerve and then fine-tune the position to achieve
adequate nerve stimulation at ≤1 mA. Information from the pediatric
population demonstrates that the success of a block is not different when
stimulation is attained with the nerve stimulator at 0 to 0.5 mA versus
0.5 to 1.0 mA.24 It has also been suggested that use of lower thresholds
(less than 0.5 mA), may increase the risk of intraneural injection. The
addition of neurostimulation for peripheral nerve blocks is largely a
matter of preference rather than practice based on outcome studies. For
those practitioners who choose to add neurostimulation to ultrasound in
children, stimulation thresholds between 0.5 and 1.0 mA are appropriate.
The use of the nerve stimulator requires avoidance or reversal of
neuromuscular blocking agents.
Local anesthetic dosing guidelines In the pediatric-aged patient, the
majority of experience with regional anesthesia has included the use of
either 0.25% bupivacaine or 0.2% ropivacaine. In commercially
prepared solutions, bupivacaine is available with or without the addition
of epinephrine (1:200,000), whereas ropivacaine is available without
epinephrine (see discussion regarding test dosing). As the majority of
these patients are already anesthetized, the decreased degree of motor
block with the lower concentration of local anesthetic (0.2% or 0.2%
versus 0.5%) is not generally an issue. If the limited motor blockade
becomes a problem, the administration of a neuromuscular blocking
agent and controlled ventilation can be provided. When used solely for
the purpose of postoperative analgesia, several studies have
demonstrated that 0.125% bupivacaine is as effective as 0.25%
bupivacaine for postoperative analgesia, although intraoperative
conditions and onset times may vary.25–27 However, ropivacaine in a
concentrations of less than 0.2% is ineffective.28 An additional
advantage of the more dilute concentrations of local anesthetic agent is
the limitation of motor blockade and the ability to ambulate
postoperatively if outpatient surgery is planned. When catheters are
placed for continuous infusions, ropivacaine or bupivacaine at
concentrations of 1 to 1.25 mg/mL (0.1 to 0.125%) are generally
effective. In some circumstances, the regional anesthetic technique is
used instead of general anesthesia. Depending on the size of the patient,
higher concentrations of bupivacaine or ropivacaine (0.5%) may be used
to provide surgical anesthesia and a greater degree of motor blockade.
As the concentration of the local anesthetic is increased, the total volume
used should be decreased to limit the total dose of bupivacaine to ≤3
mg/kg. This would amount to 1.2 mL/kg of 0.25% bupivacaine or 0.6
mL/kg of 0.5% bupivacaine.
With regional anesthetic techniques in children, the greatest risk of
morbidity and mortality lies in local anesthetic toxicity making
calculation of the appropriate dose of great importance. Although blood
levels achieved after plexus anesthesia or peripheral nerve blockade are
less than those with neural blockade in more vascular areas (i.e.,
interpleural, intercostal, and caudal/lumbar/thoracic epidural), the total
amount of bupivacaine administered during a single-bolus injection
should be limited to ≤3 mg/kg. Various studies have investigated plasma
bupivacaine levels after regional blockade in infants and children.
Following axillary blockade in children with either 2 mg/kg (n = 21) or 3
mg/kg (n = 20) of bupivacaine, Campbell et al.29 reported a rapid
absorption with a mean time to peak concentration of 0.37 hours for
either dose and mean peak plasma concentrations of 1.35 µg/mL with 2
mg/kg and 1.84 µg/mL after 3 mg/kg. The authors suggested that even
after a 30% allowance for anticipated higher arterial levels, the plasma
levels were well below what the authors considered be the toxic
concentration of bupivacaine (4 µg/mL). Sfez et al.30 evaluated plasma
bupivacaine concentrations following penile block with either 0.1 mL/kg
of 0.5% bupivacaine (0.5 mg/kg bupivacaine) or 0.1 mL/kg of 0.25%
bupivacaine plus lidocaine 1% (0.25 mg/kg bupivacaine). The time to
peak plasma concentration was similar between the two groups (27 ± 19
minutes versus 27 ± 18 minutes), as was the duration of analgesia (6.7 ±
1.7 versus 7.4 ± 1.0 hours). Although the peak plasma concentrations
were well below the toxic range in both groups (0.25 µg/mL and less
than 0.1 µg/mL), the authors noted that the addition of lidocaine
somehow altered the absorption of bupivacaine. Although the dose of
bupivacaine used with lidocaine was one-half of the dose used without
lidocaine, the peak plasma concentration was less than one-half.
As noted previously, one of the factors that may affect the plasma
concentration achieved with any local anesthetic agent is the site of
injection. The intercostal space is one of the most vascular areas into
which local anesthetic agents are injected, and hence, high plasma
concentrations may occur following intercostal blockade. Bricker et al.31
evaluated bupivacaine plasma levels and pharmacokinetics following
single-shot intercostal injections with a dose of bupivacaine (1.5 mg/kg)
in 11 neonates (less than 1 month of age) and 11 infants (1 to 6 months
of age). Peak plasma concentrations of bupivacaine occurred within 10
minutes in the majority of the patients and were 0.82 ± 0.56 µg/mL in
neonates and 0.91 ± 0.27 µg/mL in infants. They noted no difference
between the two age groups in regards to clearance or elimination half-
life. Likewise, no difference was noted when comparing cyanotic and
acyanotic patients. Plasma concentrations have also been shown to be
well within acceptable ranges following the administration of 2 to 3
mg/kg of bupivacaine or ropivacaine for caudal epidural blockade.32–34
More recently, other investigators have demonstrated low free plasma
concentrations following the use of 1.25 mg/kg of either bupivacaine or
ropivacaine for caudal epidural block.35
When ongoing continuous infusions are administered via an indwelling
catheter, it has been suggested that the dose of bupivacaine should be
less than 0.3 to 0.4 mg/kg/hr in children and less than 0.2 to 0.25
mg/kg/hr in neonates and infants.36,37 However, even with infusions at
these rates, concern has been raised regarding the potential for
bupivacaine plasma concentrations to continue to escalate when
continuous infusions are used in younger patients for more than 24 to 48
hours. To date, these data have been derived from studies with
bupivacaine administered via a continuous epidural infusion. In eight
infants, varying in age from 3 to 13 months, bupivacaine was
administered via an epidural catheter as a bolus dose of 1.2 mg/kg
followed by an infusion of 0.36 to 0.39 mg/kg/hr for up to 44 hours.38
Although the maximum bupivacaine plasma concentration was 2.02
µg/mL, there was evidence of the potential for accumulation with more
prolonged infusions as three of eight patients had increasing levels at the
time of discontinuation of the infusion. Similar results were reported by
Larsson et al.,39 who noted increasing bupivacaine levels in five of eight
patients at 48 hours after an initial bolus dose of 1.8 mg/kg and an
infusion of 0.2 mg/kg/hr. The potential for bupivacaine toxicity with
continuous infusions appears to be magnified in the youngest pediatric-
aged patients. Luz et al.40,41 reported higher plasma bupivacaine
concentrations in infants less than 4 months of age compared to those
who were more than 9 months age and higher free bupivacaine
concentrations in infants compared to older children during continuous
epidural infusions. The potential for an increased free fraction of plasma
bupivacaine in young infants has been noted in other studies related to
decreased protein binding of local anesthetic agents.42 Decreased
clearance and an increased free fraction has also been reported in
neonates and young infants with a continuous epidural infusion of
ropivacaine; however, these studies have demonstrated a stable plasma
concentration over time without an increasing plasma concentration, as
has been demonstrated with bupivacaine.43,44 These studies used
ropivacaine in doses of 0.2 mg/kg/hr in neonates and infants and up to
0.4 mg/kg/hr in older children.
The risk of local anesthetic toxicity is greatest during the injection of the
initial bolus dose. To limit the risks of toxicity, the lowest effective
volume and concentration should be used. As demonstrated by
Willschke et al.,20 ultrasound guidance may be beneficial in this regard,
as visualization of local anesthetic deposition may allow lower volumes
to be used without affecting efficacy. In addition, many practitioners still
recommend the addition of epinephrine in a concentration of 1:200,000
(5 µg/mL) to the anesthetic solution to act as a test dose to identify
inadvertent intravascular injection. Although the addition of low-dose
epinephrine (1:200,000) is a time-honored practice as a means of
inducing tachycardia and thereby alerting the clinician should
inadvertent intravascular injection occur, there is debate regarding this
practice. The classic test dose is described at 0.1 mL/kg of the solution
with 1:200,000 epinephrine thereby providing 0.5 µg/kg of epinephrine.
The accuracy of the hemodynamic response to the epinephrine “test
dose” during halothane or sevoflurane has been questioned.45,46 The
criteria to identify inadvertent intravascular injection had been a heart
rate increase ≥20 beats/minute. However, the sensitivity of this response
has been shown to be altered by the inhalational anesthetic agent that is
used, as well as whether atropine has been administered. Although the
blunting of the heart rate response is less with sevoflurane than with
halothane, further studies have demonstrated that changes in the T wave
(increase in amplitude ≥25%) or an increase in systolic blood pressure
are more sensitive indicators of inadvertent intravascular injection than
changes in heart rate.47,48 In fact, in children anesthetized with
sevoflurane, the positive response rate to the injection of the epinephrine
test dose was shown to be 100% for a T wave amplitude increase ≥25%,
95% for a systolic blood pressure increase ≥15 mm Hg, and 71% for a
heart rate increase ≥10 beats/minute.45 More recently, use of peripheral
indicators of perfusion have been suggested as a sensitive tool for the
identification of an inadvertent intravascular injection. The newest
generation of pulse oximeters provides what is termed the
plethysmographic pulse wave amplitude (PPWA), an indicator of
peripheral perfusion. Mowafi et al.49 have demonstrated that a decrease
in the PPWA is a sensitive means of identifying intravascular injection.
In their study of 80 infants and children anesthetized with either 0.5 or
1.0 minimum alveolar concentration (MAC) of sevoflurane, they
reported that the sensitivity, specificity, positive predictive value, and
negative predictive value of the PPWA was 100% with either 0.5 or 1.0
MAC of sevoflurane. Although the use of the epinephrine test dose is
not embraced by all practitioners of regional anesthesia, there may be
other potential benefits to the addition of epinephrine. In addition to
providing the potential to alert the clinician to inadvertent intravascular
injection, the addition of epinephrine may also slow vascular absorption
of the local anesthetic agent, thereby decreasing peak plasma
concentrations. Doyle et al.50 demonstrated that the addition of
epinephrine to the solution used for fascia iliaca compartment blocks
resulted in a lower peak plasma concentration (0.35 µg/mL with
epinephrine versus 1.1 µg/mL without epinephrine), as well as a delay in
the time to reach the peak plasma concentration (45 versus 20 minutes).
Regardless of whether a test dose is used or not, the dose of local
anesthetic agent should be injected incrementally while watching the
patient’s vital signs and hemodynamic status. Aspiration prior to each
incremental injection is recommended; however, this practice does not
preclude the potential for inadvertent systemic administration.
Especially in young infants, the sacrum may be incompletely ossified.
As such, the needle can be placed into the intraosseous space resulting in
systemic injection with negative aspiration.
As the majority of pediatric patients are anesthetized when regional
procedures are performed, the early symptoms of local anesthetic
toxicity (tinnitus, altered taste sensation) do not manifest clinically. The
usual last signs and symptoms of local anesthetic toxicity (seizures and
arrhythmias followed by cardiac collapse) may be the first indications of
a problem. Treatment algorithms for local anesthetic system toxicity
(LAST) should be readily available in any location in which pediatric
regional anesthesia is being performed. In the 2012 update of
management of LAST, Neal et al.51 recommended first and foremost to
obtain help followed by an initial focus of airway management and
seizure suppression and then management of cardiac arrhythmias
followed by lipid emulsion. The initial bolus dose of 20% lipid emulsion
includes 1.5 mL/kg over 1 minute. It can be repeated once as needed.
Following the initial bolus dose, a continuous infusion is initiated at 0.25
mL/kg/min and can be doubled to obtain cardiac stability. Even if the
patient responds immediately to the bolus dose, the infusion should be
started, as the effect of the lipid emulsion may be transient and the
toxicity may recur if the infusion is not started. The upper limit is 10
mL/kg of lipid emulsion over the first 30 minutes for any patient. The
incident should be posted to the website http://www.lipidrescue.org.
Summary
Given its many benefits, the use of peripheral regional blockade
continues to increase in infants and children. A recent review article
described the benefits of regional anesthesia in children, including
improving safety, pain relief, reduction of general anesthetic requirement
thereby possibly reducing neurotoxicity, and improving physiologic
hemostasis.52 Over the past 10 years, there have been improvements in
the equipment available for this practice. Although initially used in the
practice of adult regional anesthesia, there is significant use in most
centers of ultrasound to guide regional anesthetic techniques in infants
and children. Along with this interest has come evidence-based medicine
supporting the benefits of this practice. Major benefits include not only
improved accuracy but potentially decreased morbidity. Although the
practice of peripheral blockade in infants and children has included
primarily single-shot techniques, the refinement of the equipment and
improvements in technique, including the use of ultrasound imaging, has
led to interest in the use of continuous infusions to provide prolonged
analgesia following major surgical procedures.53 These practices are
beginning in selected pediatric centers, including our own. The
availability of various disposable elastomeric devices allows the use of
continuous infusions via plexus and peripheral nerve catheters in the
outpatient setting.54–56 Future evaluations are needed in this arena to
further define and demonstrate the safety, efficacy, and benefits of these
techniques.
References
1. Gray HT. A study of spinal anaesthesia in children and infants: from
a series of 200 cases. Lancet. 1909;3:913–917.
2. Brown T. History of pediatric regional anesthesia. Pediatr Anesth.
2012;22;3–9.
3. Spiegel P. Caudal anesthesia in pediatric surgery. Anesth Analg.
1962;41:218–221.
4. Fortuna A. Caudal anaesthesia: a simple and safe technique in
paediatric surgery. Br J Anaesth. 1967;39:165–170.
5. Blaylock M, Engelhardt T, Bissonnette B. Fundamentals of
neuronal apoptosis relevant to pediatric anesthesia. Pediatr Anesth.
2010;20:383–395.
6. Yaster M, Maxwell LG. Pediatric regional anesthesia.
Anesthesiology. 1989;70:324–338.
7. Giaufre E, Dalens B, Gombert A. Epidemiology and morbidity of
regional anesthesia in children: a one-year prospective survey of the
French-language society of pediatric anesthesiologists. Anesth
Analg. 1996;83:904–912.
8. Ecoffey C, Lacroix F, Giaufré E, et al. Epidemiology and morbidity
of regional anesthesia in children: a follow-up one-year prospective
survey of the French-Language Society of Paediatric
Anaesthesiologists (ADARPEF). Paediatr Anaesth. 2010;20:1061–
9.
9. Polaner D, Taenzer A, Walker B, et al. Pediatric Regional
Anesthesia Network (PRAN): a multi-institutional study of the use
and incidence of complications of pediatric regional anesthesia.
Anesth Analg. 2012;115:1353–1364.
10. Davidson A, Soriano SG. Does anesthesia harm the developing
brain-evidence or speculation? Pediatr Anesth. 2004;14:199–200.
11. Perouansky M, Hemmings HC Jr. Neurotoxicity of general
anesthetics. Anesthesiology. 2009;11:1365–1371.
12. Sanders RD, Davidson A. Anesthetic induced neurotoxicity of the
neonate: time for clinical guidelines? Pediatric Anesth.
2009;19:1141–1146
13. Oberndorfer U, Marhofer P, Bosenberg A, et al. Ultrasonographic
guidance for sciatic and femoral nerve blocks in children. Br J
Anaesth. 2007;98:797–801.
14. Schwemmer U, Markus CK, Greim CA, et al. Sonographic imaging
of the sciatic nerve and its division the popliteal fossa in children.
Paediatr Aaesth. 2004:14:1005–1008.
15. Tsui B, Suresh S. Ultrasound imaging for regional anesthesia in
infants, children, and adolescents: a review of current literature and
its application in the practice of extremity and trunk blocks.
Anesthesiology. 2010;112:473–492.
16. Bernards CM, Hadzic A, Suresh S, et al. Regional anesthesia in
anesthetized or heavily sedated patients. Reg Anesth Pain Med.
2008;33:449–460.
17. Corman SL, Skledar SJ. Use of lipid emulsion to reverse local
anesthetic-induced toxicity. Ann Pharmacother. 2007;41:1873–
1877.
18. Shah S, Gopalakrishnan S, Apuya J, et al. Use of Intralipid in an
infant with impending cardiovascular collapse due to local
anesthetic toxicity. J Anesth. 2009;23:439–441.
19. Marhofer P, Sitzwohl C, Greher M, et al. Ultrasound guidance for
infraclavicular brachial plexus anaesthesia in children. Anaesthesia.
2004;59:642–646.
20. Willschke H, Bosenberg A, Marhofer P, et al. Ultrasonographic
guided ilioinguinal/iliohypogastric nerve block in pediatric
anesthesia—what is the optimal volume? Anesth Analg.
2006;102:1680–1684.
21. Willschke H, Marhofer P, Bosenberg A, et al. Ultrasonography for
ilioinguinal/iliohypogastric nerve blocks in children. Br J Anaesth.
2005;95:226–230.
22. Marhofer P, Frickey N. Ultrasonographic guidance in pediatric
regional anesthesia. Part 1: theoretical background. Pediatr Anesth.
2006;16:1008–1018.
23. Marhofer P, Chan VWS. Ultrasound-guided regional anesthesia:
current concepts and future trends. Anesth Analg. 2007;104:1265–
1269.
24. Gurnaney H, Ganest A, Cucchiaro G. The relationship between
intensity for nerve stimulation and success of peripheral nerve
blocks performed in pediatric patients under general anesthesia.
Anesth Analg. 2007;105:1605–1609.
25. Tobias JD. Caudal epidural block with 0.125% bupivacaine for
postoperative analgesia in children following lower extremity
orthopedic procedures. Amer J Pain Manage. 1997;7:89–91.
26. Malviya S, Fear DW, Roy WL, et al. Adequacy of caudal analgesia
in children after penoscrotal and inguinal surgery using 0.5 and 1
mL/kg of 0.125% bupivacaine. Can J Anaesth. 1992;39:449–453.
27. Wolf AR, Valley RD, Fear DW, et al. Bupivacaine for caudal
analgesia in infants and children: the optimal effective
concentration. Anesthesiology. 1988;69:102–106.
28. Luz G, Innerhofer P, Haussler B, et al. Comparison of 0.1% and
0.2% ropivacaine with bupivacaine 0.2% for single caudal
anaesthesia in children. Pediatr Anesth. 2000;10:499–504.
29. Campbell RJ, Ilett KF, Dusci L. Plasma bupivacaine concentrations
after axillary block in children. Anaesth Intensive Care.
1986;14:343–346.
30. Sfez M, Mapihan YL, Mazoit X, et al. Local anesthetic serum
concentrations after penile nerve block in children. Anesth Analg.
1990;71:423–426.
31. Bricker SRW, Telford RJ, Booker PD. Pharmacokinetics of
bupivacaine following intraoperative intercostals nerve block in
neonates and infants. Anesthesiology. 1989;70:942–947.
32. Ala-Kokko TI, Partanen A, Karinen J, et al. Pharmacokinetics of
0.2% ropivacaine and 0.2% bupivacaine following caudal blocks in
children. Acta Anaesthesiol Scand. 2000;44:1099–1102.
33. Eyres RL, Bishop W, Oppneheim RC, et al. Plasma bupivacaine
concentrations in children during caudal epidural analgesia.
Anaesth Intensive Care. 1983;11:20–22.
34. Bosenberg AT, Thomas J, Lopez T, et al. Plasma concentrations of
ropivacaine following a single-shot caudal block of 1, 2 or 3 mg/kg
in children. Acta Anaesthesiol Scand. 2001;45:1276–1280.
35. Bozkurt P, Arslan I, Bakan M. Free plasma levels of bupivacaine
and ropivacaine when used for caudal blocks in children. Eur J
Anesthesiol. 2005;22:640–641.
36. Berde CB. Convulsions associated with pediatric regional
anesthesia. Anesth Analg. 1992;75:164–166.
37. Berde CB. Toxicity of local anesthetics in infants and children. J
Pediatr. 1993;122:S14–20.
38. Peutrell JM, Holder K, Gregory M. Plasma bupivacaine
concentrations associated with continuous extradural infusions in
babies. Br J Anaesth. 1997;78:160–162.
39. Larsson BA, Lonnqvist PA, Olsson GL. Plasma concentrations of
bupivacaine in neonates after continuous epidural infusion. Anesth
Analg. 1997;84:501–505.
40. Luz G, Wieser C, Innerhofer P, et al. Free and total bupivacaine
plasma concentrations after continuous epidural anaesthesia in
infants and children. Pediatr Anesth. 1998;8:473–478.
41. Luz G, Innerhofer P, Bachmann B, et al. Bupivacaine plasma
concentrations during continuous epidural anesthesia in infants and
children. Anesth Analg. 1996;82:231–234.
42. Mazoit JX, Denson DD, Samii K. Pharmacokinetics of bupivacaine
following caudal anesthesia in infants. Anesthesiology.
1988;68:387–391.
43. Bosenberg AT, Thomas J, Cronje L, et al. Pharmacokinetics and
efficacy of ropivacaine for continuous epidural infusion in neonates
and infants. Pediatr Anesth. 2005;15:739–749.
44. Berde CB, Yaster M, Meretoja O, et al. Stable plasma
concentrations of unbound ropivacaine during postoperative
epidural infusions for 24–72 hours in children. Eur J Anesthesiol.
2008;25:410–417.
45. Desparmet J, Mateo J, Ecoffey C, et al. Efficacy of an epidural test
dose in children anesthetized with halothane. Anesthesiology.
1990;72:249–251.
46. Tobias JD. Caudal epidural block: A review of test dosing and
recognition of systemic injection in children. Anesth Analg.
2001;93:1156–1161.
47. Tanaka M, Kimura T, Goyagi T, et al. Evaluating hemodynamic and
T wave criteria of simulated intravascular injection using
bupivacaine or isoproterenol in anesthetized children. Anesth
Analg. 2000;91:567–572.
48. Kozek-Langenecker SA, Marhofer P, Jonas K, et al. Cardiovascular
criteria for epidural test dosing in sevoflurane and halothane-
anesthetized children. Anesth Analg. 2000;90:579–583.
49. Mowafi HA, Arab SA, Ismail SA, et al. Plethysmographic pulse
wave amplitude is an effective indicator for intravascular injection
of epinephrine-containing epidural test dose in sevoflurane-
anesthetized pediatric patients. Anesth Analg. 2008;107:1536–1541.
50. Doyle E, Morton NS, McNicol LR. Plasma bupivacaine levels after
fascia iliaca compartment block with and without adrenaline.
Paediatr Anaesth. 1997;7:121–124.
51. Neal JM, Mulroy MF, Weinberg GL. American Society of Regional
Anesthesia and Pain Medicine checklist for managing local
anesthetic system toxicity: 2012 version. Reg Anesth Pain Med.
2012;37:16–18.
52. Bosenberg A. Benefits of regional anesthesia in children. Pediatr
Anesth. 2012;22:10–18
53. Dadire C, Pirat P, Raux O, et al. Perioperative continuous peripheral
nerve blocks with disposable infusion pumps in children: a
prospective descriptive study. Anesth Analg. 2003;97:687–690.
54. Burnett T, Bhalla T, Sawardekar A, et al. Performance of the On-Q
pain infusion device during changes in environmental temperature.
Pediatr Anesth. 2011;21:1231–1233.
55. LeRiger M, Bettesworth J, Bhalla T, et al. Comparison of flow rate
accuracy and consistency between the On-Q, Baxter, and Ambu
pain infusion devices. World J Anesthesiol. 2014;3(1):119–123.
56. Martin D, Bhalla T, Rehman S, et al. Dual continuous peripheral
nerve catheters for treatment of chronic regional pain syndrome
type I in a pediatric patient. Pediatrics. 2013;131:e323–e326.
40
Ultrasound-Guided
Brachial Plexus
Block in Infants and
Children
DAVID P. MARTIN, JOSEPH D. TOBIAS,
STEPHEN LUCAS, SUNATHENAM SURESH,
AND PAUL E. BIGELEISEN

The brachial plexus originates from the C5 through T1 nerve roots. It is


anatomically divided into trunks, divisions, cords, and ultimately into
the terminal branches. The brachial plexus is responsible for the
complete sensory and motor innervation of the arm, except the upper
half of the medial and posterior part of the arm, which is innervated by
the intercostobrachial nerve, a branch of T2. The brachial plexus can be
anesthetized by the deposition of local anesthetic agents at one of several
locations along the plexus (interscalene, parascalene, supraclavicular,
infraclavicular, or axillary approach).1,2 Although the axillary approach
was previously the most commonly used technique to anesthetize the
brachial plexus, the use of ultrasound has allowed practitioners to use
more proximal blocks safely and often with significantly less local
anesthetic agent. The advent of ultrasound has revolutionized the
practice of brachial plexus anesthesia, allowing the various approaches
in even our youngest patients. An interscalene approach is used for
procedures involving the shoulder, and surgery of the upper extremity
distal to the shoulder or the midhumerus can be performed with a
supraclavicular or infraclavicular approach. As with most regional
anesthesia that is performed in infants and children, brachial plexus
anesthesia is most frequently used as an adjunct to anesthesia with the
block placed after the induction of anesthesia either at the start or
completion of the surgical procedure.3 In these cases, the block is used
to provide postoperative analgesia. In other circumstances, brachial
plexus blockade can be used instead of general anesthesia to provide
surgical anesthesia or even occasionally as a therapeutic modality
whereby the sympathetic blockade that accompanies the motor and
sensory blockade is used to improve regional blood flow.4–7 Although
many practitioners advocate the use of only ultrasound for brachial
plexus anesthesia, the use of a nerve stimulator combined with
ultrasound is advocated by some as a means of further localizing the
blockade of the plexus to only the desired region. For specific clinical
scenarios, the nerve stimulator and low volumes of a local anesthetic
agent can be used to provide selective blockade of specific nerves of the
plexus.
Axillary approach to the brachial plexus Background and Indications: The
terminal branches of the brachial plexus (musculocutaneous, median, radial, and
ulnar nerves) arise from the cords relatively high in the axilla. The
musculocutaneous nerve is identified as an oval hyperechoic structure cephalad
to the artery, lying between the biceps and coracobrachialis muscles. The
median, radial, and ulnar nerves or the cords that give rise to them may have
variable positions around the artery (Fig. 40.1). Although there is often wide
anatomical variation in the location of the cords or terminal branches of the
brachial plexus, typical locations in relation to the axillary artery are as follows:
(1) the median nerve lies anteriorly, (2) the radial nerve lies inferolateral to the
artery, and (3) the ulnar nerve lies anteromedial to the artery. These three nerves
appear as hyperechoic clusters around the artery. The cutaneous nerves often
have the same appearance, although depending on the size of the patient, it may
not be possible to identify these nerves. In many cases, the outlines of the nerves
become apparent as the local anesthetic solution is injected. In children, it may
be possible to anesthetize the axillary nerve as well as the radial nerve by
injecting local anesthetic posterior to the axillary artery when the block is
performed high in the axilla. Local anesthetic injected using an axillary approach
in small children frequently spreads cephalad. Thus, the axillary nerve is often
blocked when local anesthetic is injected around the radial nerve. A selective
musculocutaneous nerve block is usually required with the axillary approach as
the nerve is outside the fascial sheath that encases the other nerves. This can be
performed by identification of the musculocutaneous nerve within the body of
the coracobrachialis muscle. The musculocutaneous nerve provides innervation
to the lateral aspect of the forearm. Depending on the site of surgery, axillary
block may be sufficient for anesthesia of the upper arm, forearm, and hand in
small children. If a tourniquet is used, some practitioners perform a separate
subcutaneous injection along the upper medial aspect of the arm just below the
axilla to anesthetize the intercostobrachial nerve, which is a branch of the second
intercostal nerve. This may help to delay the onset of tourniquet pain, although
there is no evidence for this practice. One of the major limitations of this
approach is that it cannot be performed in patients who cannot abduct their arm
due to pain or restricted mobility while they are awake (Fig. 40.2).
Patient position: Supine with arm abducted 90 degrees from the body
and the elbow flexed 90 degrees.
Transducer: 25-mm linear array oscillating at 8 to 13 MHz.
Transducer orientation: Transverse in the axilla.

Needle: 22G or 24G, 1-or 2-inch, blunt or insulated needle.


Local anesthetic: The volume is dependent on the size of the patient, and
the concentration is dependent on the degree of motor blockade that is
desired. For postoperative analgesia, 0.25% bupivacaine or 0.2%
ropivacaine are effective, whereas 0.5% concentrations of either are used
to provide surgical anesthesia and profound motor blockade. In most
patients, a total volume of 0.2 to 0.3 mL/kg is sufficient. Because of the
precise targeting of the nerves, only small amounts of local anesthetic
are required for each nerve target. In all cases, the total dose of
bupivacaine or ropivacaine should be ≤3mg/kg. Epinephrine in a
concentration of 1:200,000 is added to the solution to identify
inadvertent systemic injection.
Technique: For the axillary approach, the patient’s arm is abducted 90
degrees from the body and the elbow is flexed at a 90-degree angle so
that the hand is above the level of the head or behind it. For ultrasound-
guided blockade of the axillary plexus, after the axilla is washed with a
sterile prep solution, sterile ultrasound gel is placed and the ultrasound
probe is inserted into a sterile cover. A high-frequency ultrasound probe
is oriented in the transverse plane quite high in the axilla with the
axillary artery centered on the screen. If the procedure is performed
without the use of general anesthesia, local anesthesia is provided by
subcutaneous infiltration with a 27G needle. The block needle is inserted
in line, cephalad to the probe and directed toward the musculocutaneous
nerve. The nerve is surrounded with a small volume of the local
anesthetic agent (2 to 3 mL). The needle is then withdrawn to a position
just below the skin and directed toward the median nerve. An injection
of 2 to 4 mL of local anesthetic, depending on the size of the patient, is
made until the nerve is surrounded; the needle is then advanced in the
same plane targeting the ulnar nerve, and the injection process is
repeated. Lastly, the needle is withdrawn and redirected inferiorly to the
artery, targeting the radial nerve and local anesthetic is injected. In some
patients, the artery must be pushed out of the way of the needle to reach
all four nerves through the same injection site. If slightly larger volumes
are injected around the radial nerve, the axillary nerve is usually blocked
if the needle insertion site is in the proximal axilla.
Interscalene approach to the brachial plexus Background and Indications:
Prior to the introduction of ultrasound for regional anesthesia, interscalene
blocks were used less frequently in children because of the potential for adverse
effects including pneumothorax, vascular puncture, and neuraxial spread of local
anesthetic. As the trunks of the brachial plexus are organized in a superior-to-
inferior direction between the anterior and middle scalene muscles, the lower
dermatomes of the brachial plexus may be less effectively blocked than with
more distal approaches.8 Therefore, the term ulnar sparing has been used to
describe the relative lack of reliable blockade of the caudal or lower components
of the brachial plexus. The relative degree of ulnar sparing may be reduced with
the use of ultrasound because the distance anatomically between the interscalene
and more distal access point to the brachial plexus is quite small.9 The
ultrasound shortens the distance because an optimal picture for the interscalene
approach is obtained more distal than where the landmarks for a blind technique
would dictate. In addition to the risks of pneumothorax, vertebral artery puncture
with local anesthetic toxicity is of particular concern. The vertebral artery
provides direct blood flow to the central nervous system, thus even minute
amounts of local anesthetic agents can lead to seizures in the event of arterial
injection. Due to the proximity of the vertebral column and spinal cord, there is
also a potential for epidural or intrathecal injection. These aforementioned risks
are purportedly increased with a blind, nerve-stimulator approach, and the risk of
such complications should be decreased by the use of ultrasound and direct
visualization of the site of local anesthetic placement.10 Additional issues
include blockade of surrounding nerves resulting in phrenic nerve blockade with
hemidiaphragm paralysis, recurrent laryngeal nerve block with unilateral vocal
cord paralysis, and sympathetic block with Horner syndrome.11,12 Although
phrenic/recurrent laryngeal nerve block are well tolerated in healthy, older
patients, infants and patients with respiratory dysfunction may be dependent on
the diaphragm for respiratory function.13 Additionally, airway compromise can
occur with unilateral vocal cord paralysis in infants and young children.
Especially when the block is performed high in the interscalene groove, our
clinical experience has demonstrated that cephalad spread to the cervical plexus
may also occur so that effective analgesia may be provided in these dermatomes
as well, thereby making this block effective for shoulder and perhaps even
clavicular surgery.
Patient position: Supine with the head turned away from the operator
and the side of the block.
Transducer: 25-mm linear array oscillating at 13 MHz or 35-mm linear
probe oscillating at 8 to 13 MHz for larger patients.
Transducer orientation: Transverse oblique in the supraclavicular fossa
and then transverse over the sternocleidomastoid muscle at the level of
the thyroid cartilage.

Needle: 22G or 24G, 1-or 2-inch, blunt or insulated needle.


Local anesthetic: The volume is dependent on the size of the patient, and
the concentration is dependent on the degree of motor blockade that is
desired. For postoperative analgesia, 0.25% bupivacaine or 0.2%
ropivacaine are effective, whereas 0.5% concentrations are used to
provide surgical anesthesia and profound motor blockade. In most
patients, a volume of 0.2 to 0.4 mL/kg is sufficient. There is some
literature to state that, with the addition of ultrasound guidance, the total
volume of local anesthetic required for surgical blockade is markedly
reduced.14 In all cases, the total dose of bupivacaine or ropivacaine
should be ≤3 mg/kg. Epinephrine in a concentration of 1:200,000 is
added to the solution.
Technique: For ultrasound-guided block, the skin is washed, and sterile
ultrasound gel is placed in the supraclavicular fossa. A 25-mm linear
probe oscillating at 13 MHz is placed in the supraclavicular fossa in an
oblique sagittal orientation. The subclavian artery and brachial plexus
are identified. The brachial plexus is traced cephalad to the level of the
thyroid cartilage while the probe is rotated into an axial alignment. At
this level, the roots of C5, C6, and C7 appear as hypoechoic circles
sandwiched between the anterior and middle scalene muscles (Figs. 40.3
and 40.4). In small children, the carotid artery and internal jugular vein
will appear medial to the brachial plexus and the sternocleidomastoid
muscle will overlie the scalene muscles in some patients. These
structures can be identified by placing the probe in a horizontal position
just above the clavicle/sternum in the middle of the neck. The probe is
then moved laterally to identify the carotid artery, the jugular vein, and
then the anterior and middle scalene muscles. Alternatively, the probe
can be placed in the supraclavicular fossa as for a supraclavicular
approach and then angled or moved cephalad to identify the brachial
plexus between the anterior and middle scalene muscles. The nerve roots
can also be seen adjacent to the transverse process by toggling the probe
so that it has a caudal oblique angle relative to the surface of the neck
(Fig. 40.5). The needle is inserted in-line with the ultrasound probe from
its lateral aspect and directed toward the roots. The needle can also be
inserted above the probe using an out-of-plane technique. When an out-
of-plane technique is used, a gentle bounce on the needle will displace
the tissues and aid in identification of the needle path. Once the needle is
adjacent to the roots, the local anesthetic solution is injected in aliquots
of 3 mL until the plexus is surrounded. It may be necessary to insert the
needle between the roots to ensure circumferential spread of local
anesthetic around the plexus. Care must be exercised not to inject local
anesthetic into the nerves at this level. Each nerve root is in direct
communication with the cerebrospinal fluid surrounding the spinal cord.
Depending on the site of surgery, separate injections may be made to
ensure coverage of the lower portions of the brachial plexus (C8–T1) to
ensure anesthesia of the distribution of the ulnar nerve.
Supraclavicular and infraclavicular approaches to the brachial plexus
Background and Indications: Both supraclavicular and infraclavicular
approaches to the brachial plexus have been described and used with a high
degree of success in the adult population. The supraclavicular and infraclavicular
nerve blocks occur at the level of the divisions and cords of the brachial plexus,
respectively. Prior to the advent of ultrasound for regional anesthesia, there was
a paucity of reports regarding the use of a supraclavicular approach in children.15
Ultrasonography at the level of supra and infraclavicular approaches to the
brachial plexus allows visualization of the subclavian artery, the pleura, and the
first rib. In the supraclavicular fossa, blockade of the brachial plexus may also
cause vocal cord paresis, Horner syndrome, and phrenic nerve paresis. Younger
patients such as infants who are more dependent on diaphragmatic function for
ventilation or those with diminished respiratory reserve may not tolerate phrenic
nerve blockade.
Both the supraclavicular and infraclavicular approaches can be used to
provide anesthesia for all branches of the brachial plexus. However,
given the proximity of these approaches to the pleura and vascular
structures, they do carry a risk of vascular puncture or pneumothorax. As
noted earlier, the use of these techniques has increased in the recent past
as the use of ultrasound-guided needle placement has empowered
physicians to perform these blocks. There is still a relative paucity of
data definitively demonstrating that such complications are avoided by
the use of the ultrasound. The supraclavicular and infraclavicular
approaches to the brachial plexus offer distinct advantages. The patient
is not required to move the potentially injured extremity. The arm can
remain at rest in a neutral position.16 Next, anatomically, the brachial
plexus, at the supraclavicular level, lies entirely lateral to the artery,
allowing for reduced needle passes past the artery. Both infraclavicular
and supraclavicular blockade are suitable for surgical procedures distal
to the shoulder. The infraclavicular block can be used as an alternative
approach to the brachial plexus when the supraclavicular approach is
impeded by a vascular structure such as a dorsal scapular or transverse
cervical arteries.17
Patient position: Supine with the head turned away from the practitioner.
Transducer: 25-mm linear array oscillating at 13 MHz (supraclavicular);
the 35-mm linear probe may be useful in children who weigh more than
40 kg, and the 11-mm curved array oscillating at 10 MHz may be useful
for the performance of blocks where there is limited space for the
placement of the probe, such as the supraclavicular area.
Transducer orientation: Transverse oblique in the supraclavicular fossa.
Sagittal in the deltopectoral groove (infraclavicular).
Local anesthetic: The volume is dependent on the size of the patient, and
the concentration is dependent on the degree of motor blockade that is
desired. For postoperative analgesia, 0.25% bupivacaine or 0.2%
ropivacaine are effective, whereas 0.5% concentrations are used to
provide surgical anesthesia and profound motor blockade. In most
patients, a volume of 0.2 to 0.3 mL/kg is sufficient. In all cases, the total
dose of bupivacaine or ropivacaine should be ≤3 mg/kg. Epinephrine in
a concentration of 1:200,000 is added to the solution.
Technique: For the ultrasound-guided supraclavicular approach, the skin
is prepped and sterile gel is placed in the supraclavicular fossa. The
probe is placed in an oblique, transverse orientation in the
supraclavicular fossa and the subclavian artery and brachial plexus are
identified (Fig. 40.6). The plexus should appear as a cluster of
hypoechoic structures lateral to the artery (Fig. 40.7). The first rib can be
identified immediately below the plexus as a hyperechoic streak. The
pleura has a similar appearance to the rib but lies 3 to 10 mm deep to the
plexus, whereas the rib lies immediately beneath the plexus. Classically,
the needle is inserted in line at the medial end of the probe and advanced
toward the subclavian artery. When the needle reaches the equator of the
artery, the needle is used to push the artery away from the plexus. The
needle is then advanced between the artery and plexus until the tip of the
needle is immediately above the rib. Injection of a small amount of local
anesthetic solution (0.05 to 0.1 mL/kg) should cause the entire plexus to
float off the rib. With further use of ultrasound in the pediatric
population, many practitioners have begun to advocate a lateral-to-
medial approach where the needle is advanced toward the plexus using a
very shallow entry angle to the skin. The needle is inserted deep to the
epineurium containing the brachial plexus and a popping sensation is felt
with typical blunt regional block needles. The typical first injection point
is located in the “corner-pocket” location at the posterolateral side of the
subclavian artery. The plexus is seen to lift up and away from the first
rib and often away from the artery as well. The needle is then moved
into the superior portion of the plexus, and the remainder of the local
anesthetic is deposited there. We advocate for the lateral-to-medial
approach, as it limits the risk of puncturing the subclavian artery or
causing a pneumothorax while providing optimal in-line access to the
brachial plexus. Either approach, if used safely with optimal needle
visualization, provides a safe and effective means of blocking the
brachial plexus at the supraclavicular level while preserving arterial and
pleura integrity. The injection of local anesthetic is done in small
aliquots as the needle is withdrawn. A total of 0.2 to 0.3 mL/kg of local
anesthetic solution should be sufficient to achieve a block depending
upon the size of the patient and specific trajectory of the needle.
Catheters are best placed through a needle that is immediately above the
rib after an injection has been made to float the plexus off the rib.
For ultrasound-guided infraclavicular blockade, the skin is washed and
sterile gel is placed in the deltopectoral groove, which is also called the
infraclavicular fossa. The arm may be abducted 90 degrees and flexed at
the elbow as in axillary block or left at the patient’s side as the patient’s
comfort dictates (Fig. 40.8). A curved, 11-mm probe oscillating at 8 to
13 MHz is placed in the fossa, and the axillary artery and vein are
imaged. The pectoralis major and minor muscles are identified
superficial to the axillary artery, and the cords can be seen as hypoechoic
circles (Fig. 40.9). The cords are named lateral, posterior, and medial as
they are related to the axillary artery. The rib or pleura can be seen as
hyperechoic lines posterior and inferior to the plexus and artery. The
needle is inserted in line with the transducer, cephalad to the probe, and
directed toward the cords (Fig. 40.9). Either a cephalad or caudad entry
site of the needle is acceptable; however, the cephalad-toward-caudad
approach is generally advocated, as it likely places the pleura at less risk.
Local anesthetic is injected in aliquots of 0.05 to 0.1 mL/kg superiorly,
posteriorly, and inferiorly to the cords or within the plexus itself. In the
event of poor visualization of the cords, some advocate deposition of the
local anesthetic in a “horseshoe” shape around the posterior aspect of the
artery. The needle tip is very close to the pleura in this block, and it is of
critical importance to keep the tip in view at all times. A total of 0.2 to
0.3 mL/kg of local anesthetic solution is required to achieve effective
blockade depending upon the size of the patient. Catheters are best
placed inferiorly to the plexus or in the space between the lateral and
posterior cords. The infraclavicular approach lends itself well to catheter
placement, as the muscles generally provide a good anchor for the
catheter.
An additional approach to the brachial plexus that has been added to the
armamentarium for approaches to the brachial plexus is the lateral
infraclavicular block or the coracoid block.18,19 The arm is adducted to
the trunk, and the elbow flexed at a 90-degree angle with the forearm
placed on the abdomen. The coracoid process is palpated, and the site of
needle insertion is 0.5 cm distal and medial to the coracoid (Fig. 40.10).
Although the use of ultrasound has not been reported in the literature for
this approach in children, identification of the coracoid process could
serve as an easily identifiable landmark with subsequent identification of
the hypoechoic cords of the brachial plexus as they pass 0.5 to 1 cm
caudal (inferiorly) to the coracoid. For this approach, the nerves of the
brachial plexus can be seen arranged radially around the axillary artery
in the infraclavicular space (Fig. 40.11). If a stimulating needle is used,
contractions of the upper extremity are noted. Recommendations for
needle use and local anesthetic dosing would remain the same as for any
of the previously described blocks.
Summary
The brachial plexus is derived from the ventral (anterior) branches of
spinal roots C5–C8 and T1. These spinal nerves pass through the
intervertebral foramina and course between the anterior and middle
scalene muscles, which attach respectively to the anterior and posterior
tubercles of the transverse processes of the cervical vertebrae. The
brachial plexus provides sensory and motor innervation to the majority
of the shoulder and upper extremity. A portion of the shoulder is
innervated by the cervical plexus, and the medial aspect of the upper arm
is innervated by the intercostobrachial nerve, a branch of the second
intercostal nerve, which communicates with the medial cutaneous nerve
of the arm. As the spinal roots exit the vertebral column and pass
between the anterior and middle scalene muscles, they form three trunks
(superior, middle, and inferior). As the trunks exit the interscalene
groove, they lie in a cephaloposterior position to the subclavian artery as
it courses along the upper surface of the first rib. The trunks split into
anterior and posterior divisions, which then unite to form cords (lateral,
posterior, and medial) that are named because of their relationship to the
axillary artery at the level of the coracoid process. These cords surround
the axillary artery and are blocked with the infraclavicular approach to
the brachial plexus. The cords further divide into the nerves of the
brachial plexus (musculocutaneous, ulnar, median, and radial) as they
surround the axillary artery and can be blocked with an axillary
approach to the brachial plexus. The brachial plexus can be anesthetized
by one of several approaches (interscalene, parascalene, supraclavicular,
infraclavicular, or axillary route). The site of blockade will be
determined by the location of the surgery as well as the expertise of the
anesthesia provider. With the use of ultrasound-guided techniques, it is
postulated that the risks of adverse events associated with these
techniques including vascular puncture or pneumothorax should
decrease.20 Given that, these techniques continue to play an increasingly
important role in the provision of postoperative analgesia in the
pediatric-aged patient.
References
1. Tobias JD. Brachial plexus anaesthesia in children. Pediatr Anaesth.
2001;11:265–275.
2. Dalens B, Vanneuville G, Tanguy A. A new parascalene approach to
the brachial plexus in children: comparison with the supraclavicular
approach. Anesth Analg. 1987;66:1264–1271.
3. Altintas F, Bozfurt P, Ipek N, et al. The efficacy of pre-versus post-
surgical axillary block on postoperative pain in paediatric patients.
Paeditr Anaesth. 2000;10:23–28.
4. Messeri A, Calamandrei M. Percutaneous central venous
catheterization in small infants: axillary block can facilitate the
insertion rate. Paeditr Anaesth. 2000;10:527–530.
5. Ross DM, Williams DO. Combined axillary plexus block and basal
sedation for cardiac catheterization in young children. Br Heart J.
1970;32:195–197.
6. Inberg P, Kassila M, Vilkki S, et al. Anaesthesia for microvascular
surgery in children: a combination of general anaesthesia and
axillary plexus block. Acta Anaesthesiol Scand. 1995;39:518–522.
7. Audenaert SM, Vickers H, Burgess RC. Axillary block for vascular
insufficiency after repair of radial club hands in an infant.
Anesthesiology. 1991;74:368–370.
8. Vester-Andersen T, Christiansen C, Hansen A, et al. Interscalene
brachial plexus block: area of analgesia, complications, and blood
concentrations of local anesthetics. Acta Anaesthesiol Scand.
1981;25:81–84.
9. Williams SR, Chouinard P, Arcand G, et al. Ultrasound guidance
speeds execution and improves the quality of supraclavicular block.
Anesth Analg. 2003;97:1518–1523.
10. Hadzic A, Sala-Blanch X, Xu D. Ultrasound guidance may reduce
but not eliminate complications of peripheral nerve blocks.
Anesthesiology. 2008;108;557–558.
11. Urmey WF, Talts KH, Sharrock NE. One hundred percent incidence
of hemidiaphragmatic paresis associated with interscalene brachial
plexus anesthesia as diagnosed by ultrasonography. Anesth Analg.
1991;72:498–503.
12. Sinha SK, Abrams JH, Barnett JT, et al. Decreasing the local
anesthetic volume from 20 to 10 mL for ultrasound-guided
interscalene block at the cricoid level does not reduce the incidence
of hemidiaphragmatic paresis. Reg Anesth Pain Med. 2011;36:17–
20.
13. Kempen PM, O’Donnell J, Lawler R, et al. Acute respiratory
insufficiency during interscalene plexus block. Anesth Analg.
2000;90:1415–1416.
14. Gautier P, Vandepitte C, Ramquet C, et al. The minimum effective
anesthetic volume of 0.75% ropivacaine in ultrasound-guided
interscalene brachial plexus block. Anesth Analg. 2011;113:951–
955.
15. Pande R, Pande M, Bhadani U, et al. Supraclavicular brachial
plexus block as a sole anaesthetic technique in children: an analysis
of 200 cases. Anaesthesia. 2000;55:798–810.
16. Marhofer P, Willschke H, Kettner SC. Ultrasound-guided upper
extremity block—tips and tricks to improve the clinical practice.
Paediatr Anaesth. 2012;22:65–71.
17. Murata H, Sakai A, Hadzic A, et al. The presence of transverse
cervical and dorsal scapular arteries at three ultrasound probe
positions commonly used in supraclavicular brachial plexus
blockade. Anesth Analg. 2012;115:470–473.
18. Fleishmann E, Marhofer P, Greher M, et al. Brachial plexus
anaesthesia in children: lateral infraclavicular block versus axillary
approach. Paediatr Anaesth. 2003;13:103–108.
19. Maria BDJ, Tielens LKP. Vertical lateral infraclavicular brachial
plexus block in children: a preliminary study. Pediatr Anesth.
2004;14:931–935.
20. Koscielniak-Nielsen ZJ. Ultrasound-guided peripheral nerve blocks:
what are the benefits? Acta Anaesthesiol Scand. 2008;52:727–737.
41
Ultrasound-Guided
Lower Extremity
Blockade in Children
RALPH BELTRAN, JOSEPH D. TOBIAS,

STEPHEN LUCAS, AND PAUL E. BIGELEISEN

Depending on the site of surgery or injury and the nerves that need to be
anesthetized, there are several regional anesthesia techniques of the
lower extremity that may be applicable to the pediatric population.
These techniques may be used for procedures of the foot and ankle (such
as club foot repair or tendon lengthening of the lower extremity) or for
major operations of the femur and hip (including the treatment of
traumatic femoral fractures or open reduction and internal fixation of the
hip due to traumatic or congenital conditions). In addition to their use for
postoperative analgesia, isolated blocks of the lower extremity may be
used instead of anesthesia in patients with comorbid diseases, which
may increase the risk of general anesthesia, such as in patients with
undiagnosed myopathy. In these patients, muscle biopsy can generally
be accomplished with femoral nerve block or a fascia iliaca block.
Alternatively, there are also isolated case reports of the use of regional
anesthesia to induce sympathectomy in the lower extremity for treatment
of vascular compromise of various etiologies. In the past, analgesia
following lower extremity procedures was most commonly
accomplished with the use of a caudal approach; however, given data
demonstrating the efficacy and lower risk of adverse events with
peripheral nerve blockade versus caudal epidural block, it is likely that
there will be increasing use of peripheral nerve block of the lower
extremity.
Lumbar plexus block Background and Indications: The most proximal
approach to the lumbar plexus is direct block at the lumbar plexus. The
technique provides anesthesia of the three nerves of the lumbar plexus (femoral,
lateral femoral cutaneous, and obturator) with a single injection. This approach
is also commonly referred to as the psoas compartment block. The authors
advocate a slightly more medial approach (Fig. 41.3) to needle insertion than
that described in the adult population. As with the techniques of Winnie et al.
and Chayen et al., a line is drawn through the posterior superior iliac spine
(PSIS) parallel to the spine. A second line is drawn connecting the two iliac
crests (the intercristal line). The point of intersection of these two lines is
identified and the needle is moved slightly (1 to 2 cm) medially along the
intercristal line. The needle is inserted at a 90-degree angle to the skin and
advanced using a nerve stimulator. In addition, ultrasound guidance can be used
with the needle in plane with the transducer held in the transfer or longitudinal
plane. As the psoas compartment is entered, a loss of resistance is felt and a
muscle response in the quadriceps muscle will be obtained. During needle
insertion, if the transverse process is contacted, the needle is walked off the
transverse process in either a caudad or cephalad direction. The lumbar and
sacral plexuses lie in the same anatomic planes in the paravertebral space along
the vertebral column, consequently variable analgesia of the sacral plexus
occurs. The potential to achieve anesthesia of both the lumbar and sacral plexus
makes this approach suitable for femoral osteotomies as well as surgical
procedures of the hip. Although there are limited data in the pediatric population,
lumbar plexus analgesia can be used to provide analgesia following major
orthopedic surgical procedures of the hip, such as triple osteotomy of the pelvis,
or femur surgery, such as femoral osteotomy.
Other investigators have provided recent additional insight into the use
of the lumbar plexus block in the pediatric population. Using ultrasound
guidance, Kirchmair1 demonstrated that weight rather than age provided
the best measure for estimating the depth of the lumbar plexus in a
cohort of 32 children. The patients were stratified into three age groups
(3 to 5 years, 5 to 8 years, and more than 8 years). The lumbar plexus
could be delineated in 19 of 20 cases in group 1, 17 of 20 cases in group
2, 22 of 24 cases at either L3–L4 or L4–L5. In all patients, the lumbar
plexus was situated within the posterior part of the psoas major muscle.
The strongest positive correlation existed between skin-to-plexus
distances and the children’s weight. Although there was a significant
increase in the skin-to-plexus distance in the three groups, no difference
was noted in the skin-to-plexus distance at L3–L4 versus L4–L5. The
distance from the skin to the anterior border of the lumbar plexus at L3–
L4 was 2.5 ± 0.4, 2.7 ± 0.5, and 3.2 ± 0.3 cm in the three groups,
respectively. The authors concluded that ultrasound guidance enabled
safe und successful posterior approaches to the lumbar plexus, thus
resulting in effective anesthesia and analgesia of the inguinal region.
Although this study demonstrated that the lumbar plexus was within the
body of the psoas major muscle in all cases, subsequent work from the
same group has demonstrated that the lumbar plexus occasionally is
located posterior to the psoas major muscle. In a prospective randomized
comparison of continuous epidural versus lumbar plexus block
following surgical procedures of the hip and femur in a cohort of 40
children, lumbar plexus block provided equivalent analgesia with a
superior adverse effect profile. The cohort for the study included 40
children. After the induction of general induction, 0.5 mL/kg of 0.375%
ropivacaine was injected via the epidural or lumbar plexus catheter. This
was followed postoperatively by an infusion of 0.2% ropivacaine at 0.1
mL/kg/hr for the lumbar plexus catheter or 0.2 mL/kg/hr for the
epidural. Postoperative analgesia was excellent for both continuous
block techniques with comparable pain scores and need for supplemental
analgesia. The number of children who had at least one adverse effect,
plasma ropivacaine levels, and the need to stop the local anesthetic
infusion prematurely was significantly higher in epidural group. These
data and other investigators have demonstrated the efficacy of lumbar
plexus blockade in providing analgesia following hip and femur surgery
in the pediatric population. Although rare, reported complications for
lumbar plexus blockade have included cardiac arrest during
intravascular injection, muscular and renal hematoma, epidural
anesthesia, spinal blockade, and retroperitoneal injection.
Anatomy: The sensory and motor innervation of the lower extremity is
derived from the lumbar and sacral plexuses. The lumbar plexus is
formed by the union of the anterior rami of the first four lumbar nerves
(L1–L4) with variable input from the 12th thoracic nerve (T12) and L5.
The lumbar plexus lies in the “psoas compartment” in the paravertebral
space. The anterior border of the compartment is formed by the psoas
major, and the posterior border is formed by the quadratus
lumborum/erector spinae muscles (Fig. 41.1 A to D). In many cases, the
lumbar plexus lies within the psoas muscle. As the lumbar plexus
emerges from the psoas compartment, it divides into the three nerves
that innervate the anterior portion of the proximal aspect of the lower
extremity: the femoral nerve, the lateral femoral cutaneous nerve, and
the obturator nerve. The femoral nerve provides sensory innervation to
the anterior and medial aspects of the thigh and motor innervation to the
quadriceps muscles (Fig. 41.2 A,B). The lateral femoral cutaneous nerve
is purely sensory, providing sensory innervation to the lateral aspect of
the thigh. It branches from the lumbar plexus and enters the thigh deep
to the inguinal ligament, medial to the anterior superior iliac spine. The
obturator nerve provides motor innervation to the adductors of the leg as
well as sensory innervation to part of the medial aspect of the lower
portion of the thigh. The obturator nerve also innervates the knee joint,
making it necessary to anesthetize to achieve analgesia following
procedures involving the knee.
Patient position: Lateral decubitus with knees and hips flexed as feasible
or Sims position, with side to be blocked upright.
Transducer: Depending on the size of the patient, either the 35-mm
linear array oscillating at 13 MHz or the curved 11-mm intermediate
frequency probe oscillating at 5 to 10 MHz will allow visualization of
bony structures, psoas major muscle, and in most cases the lumbar
plexus. The lower frequency probe is used for larger patients given its
increased depth of penetration.
Transducer orientation: Transverse, just below the site of needle
insertion, using an in-plane technique. Alternatively, the transducer can
be placed in the longitudinal plane with the needle inserted caudad
(distal) to the transducer (in plane).

Needle: Depending on the size of the patient and the depth of the lumbar
plexus, either a 20G, 6-inch or 21G, 4-inch needle is used. Alternatively,
an 18G insulated Tuohy needle of variable length can be used if a
catheter is going to be placed for a continuous infusion.
Local anesthetic agent: The volume is dependent on the size of the
patient, and the concentration is dependent on the degree of motor
blockade that is desired. For postoperative analgesia, 0.25% bupivacaine
or 0.2% ropivacaine are effective, whereas 0.5% concentrations are used
to provide surgical anesthesia and profound motor blockade. In most
patients, a volume of 0.2 to 0.4 mL/kg is sufficient. In all cases, the total
dose of bupivacaine or ropivacaine should not exceed 3 mg/kg.
Epinephrine in a concentration of 1:200,000 is added to the solution. For
continuous infusions, 0.2% ropivacaine is infused at 0.1 mL/kg/hr.
Technique: For use of ultrasound to guide the lumbar plexus block,
sterile preparation of the skin is performed and sterile ultrasound gel is
placed over the skin of the lumbar area at the anticipated needle
insertion point. The ultrasound probe is placed in a transverse position to
identify the bony elements of the vertebral body and the muscles that
border the psoas compartment (the quadratus lumborum/erector spinae
and the psoas major muscles) (Fig. 41.1 A,B). The needle is inserted
above (superior) to the probe. The kidney can usually be identified
superior (cephalad) to the psoas muscle. In many cases, the lumbar
plexus can be visualized within the psoas major muscle. The needle is
advanced between two of the transverse processes of the lumbar
vertebrae and can be viewed as it pierces the quadratus
lumborum/erector spinae muscles and enters the psoas compartment.
Once the psoas compartment is entered, the ultrasound transducer can be
turned into a longitudinal plane to verify the location of the tip of the
needle at the lumbar plexus. Electrostimulation may also be used to
confirm needle placement. Once the correct placement of the needle is
confirmed, the local anesthetic solution is injected until the plexus is
surrounded.
Some practitioners prefer a longitudinal approach (Fig. 41.1 C,D). In
this case, a parasagittal scan is obtained that identifies the transverse
processes and quadratus lumborum and psoas muscles. In slim patients,
the nerve plexus may also be identified. The needle is inserted in line
with the probe until the lumbar plexus compartment is entered. Needle
position may be confirmed with electrostimulation. Once proper needle
placement is confirmed, local anesthetic is injected until the nerve
plexus is surrounded.
Femoral nerve block Background and Indications: Block of the femoral
nerve can be used to provide analgesia following surgical procedures on the
anterior or lateral aspect of the thigh, knee arthroscopy, anterior cruciate
ligament reconstruction, patellar ligament realignment, and traumatic femur
fracture or following femoral osteotomies.
There are two basic approaches described for femoral nerve block. The
first involves direct block of the nerve just below the inguinal crease and
lateral to the femoral artery. This technique can also be modified to
provide what has been termed a 3-in-1 block, or the inguinal
perivascular approach, whereby block of the femoral, lateral femoral
cutaneous, and obturator nerve may be feasible (Fig. 41.2 A,B). As
originally described, the theory behind the block is that a fascial sheath
that surrounds the femoral nerve can be used as a conduit to carry local
anesthetic centrally to the lumbar plexus. This is accomplished by the
use of larger volume of local anesthetic than is used for isolated femoral
nerve block and holding pressure distal to the site of injection. Our own
cadaver studies suggest that Winnie’s 3-in-1 block was actually an
intraneural block wherein the local anesthetic traveled proximally to the
origin of the lumbar plexus within the epineurium. Another possibility,
shown by our cadaver studies, is that the local anesthetic originally
travels proximally about 10 cm within the epineurium and then ruptures
into the space containing the three nerves between the psoas and iliacus
muscles.
The second approach to the femoral nerve is the fascia iliaca block. With
this technique, a large volume of local anesthetic solution is injected
more laterally with medial spread to the femoral nerve, superiorly to the
obturator and lateral femoral cutaneous nerves. The point of needle entry
is at the junction of the outer and middle third of the line connecting the
symphysis pubis and the anterior superior iliac crest (Fig. 41.3). When
compared to nerve stimulation, ultrasound guidance in the performance
of femoral nerve block in the pediatric population has been shown to
result in a longer duration of analgesia and a decrease in the volume of
local anesthetic that is required.

Anatomy: The femoral nerve is the largest branch of the lumbar plexus.
It arises from the dorsal division of the anterior rami of L2–L4 and
descends into the pelvis lateral to the psoas major muscle, where it
passes deep to the inguinal ligament. In the anterior compartment of the
thigh, the femoral nerve divides into multiple branches supplying the
muscle, joints, and skin of the anterior thigh. The inguinal crease is
generally several centimeters caudad to the inguinal ligament. Below the
inguinal crease is the level at which the block is performed. At this level,
the nerve lies deep to the fascia lata and the fascia iliaca and is separated
from the femoral artery and vein by the iliopectineal ligament (Cooper
ligament). The anterior branch of the femoral nerve innervates the
pectineus muscle and is responsible for thigh adduction on stimulation.
The posterior branch innervates the quadriceps femoris muscles and
provides leg extension and patellar elevation on stimulation. The
superficial branch lies deep to the fascia lata and superior to the fascia
iliaca. It stimulates contraction of the sartorius muscle. Because this
branch lies above the fascia iliaca, it is not a suitable stimulation end
point for femoral nerve block. The saphenous nerve is a cutaneous
branch of the femoral nerve, which supplies innervation to the skin over
the medial aspect of the leg and foot.
Patient position: Supine.
Transducer: 25-mm linear array oscillating at 13 MHz or 35-mm linear
probe oscillating at 8 to 13 MHz for larger patients.
Transducer orientation: Transverse, below the inguinal ligament.

Needle: 22G or 24G, 1-or 2-inch, blunt or insulated needle for single-shot
technique. 18G insulated Tuohy needle for continuous block.
Local anesthetic: The volume is dependent on the size of the patient, and
the concentration is dependent on the degree of motor blockade that is
desired. For postoperative analgesia, 0.25% bupivacaine or 0.2%
ropivacaine are effective, whereas 0.5% concentrations are used to
provide surgical anesthesia and profound motor blockade. In most
patients, a volume of 0.2 to 0.4 mL/kg is sufficient. In all cases, the total
dose of bupivacaine or ropivacaine should be ≤3 mg/kg. Epinephrine in
a concentration of 1:200,000 is added to the solution.
Technique: For use of ultrasound to guide femoral nerve block, sterile
preparation of the skin is performed and sterile ultrasound gel is placed
over the skin below the inguinal ligament. A high frequency 10-to 13-
MHz linear probe is placed transversely in the inguinal crease. The
femoral artery is identified, and the nerve is imaged lateral to the artery
and deep to the fascia iliaca. The nerve is of variable size and may be
triangular. The needle is inserted at the lateral end of the probe using an
in-line technique. Once the fascia iliaca has been pierced, the local
anesthetic is injected and observed to surround the nerve. The local
anesthetic can be injected above the nerve and then the needle redirected
and an additional amount of local anesthetic injected below the nerve to
ensure that the femoral nerve is surrounded by the solution. If a fascia
iliaca block is used, the point of needle insertion is more lateral. A line is
drawn connecting the anterior pubic tubercle and the anterior superior
iliaca spine. The point of needle insertion is where the lateral one-third
of this line meets the medial two-thirds. From this point, a line is drawn
at a 90-degree angle to below the inguinal ligament. As the needle is
inserted, a double loss of resistance may be felt as the fascia lata and
fascia iliaca are penetrated. The use of ultrasound during needle
placement allows visualization of the local anesthetic solution as it is
injected deep to the fascia iliaca ligament and surrounds the femoral
nerve (Fig. 41.3).
Distal block of the femoral nerve: The femoral nerve can also be blocked more
distal at a point where it transitions into the saphenous nerve. This technique can
be combined with a sciatic block at the knee (popliteal fossa block) to provide
analgesia of the entire leg below the knee. It can be performed at the level of the
midthigh (Fig. 41.4) next to the femoral artery or above the knee where the
saphenous nerve is located between the sartorius and gracilis muscles (Fig.
41.5). A high-frequency probe is placed over the medial aspect of the thigh in a
transverse orientation. Once the nerve has been located, the needle is inserted at
the lateral aspect of the probe using an in-line approach. In smaller patients,
given the size of the nerve, ultrasound visualization may be difficult. In that
setting, the local anesthetic solution is infiltrated around the femoral artery (Fig.
41.4) or the geniculate artery (Fig. 41.5). At more distal sites as it approaches the
medical epicondyle of the femur, the saphenous nerve may be found between the
sartorius and gracilis muscles adjacent to the geniculate artery (Fig. 41.5C,D).
Sciatic nerve block Background and Indications: The sacral plexus is
formed by the anterior rami of lumbar nerves 4 and 5 and sacral nerves 1 to 3
with variable input from the fourth sacral nerve. The sacral plexus forms on the
anterior surface of the sacrum and travels distally anterior to the piriformis
muscle. The sacral plexus gives rise to the posterior cutaneous nerve of the thigh
(small sciatic nerve) and the sciatic nerve. The sciatic nerve provides sensory
and motor innervation to the posterior aspect of the thigh and knee and all of the
leg below the knee except for the skin innervated by the saphenous branch of the
femoral nerve. A sciatic nerve block is used most commonly to provide
analgesia following contracture release of the thigh, leg, or ankle or surgical
repair of fractures of the leg and ankle. The posterior approach (subgluteal and
posterior popliteal) remain the most popular, perhaps because these were the
most common approaches before the advent of ultrasound guidance. The lateral
approach at the midthigh and anterior approach have gained popularity because
of their relative ease with ultrasound guidance and because the patient’s lower
extremity does not have to be elevated to perform the block. Ultrasound
guidance offers additional benefit when using the popliteal approach given the
anatomic variability in the location of the division of the sciatic nerve into its
terminal branches, the tibial and peroneal nerves.
Anatomy: The sciatic nerve exits the pelvis through the grater sciatic
foramen. Below the piriformis muscle, the nerve descends medial to the
midpoint of a line drawn between the greater trochanter of the femur and
the ischial tuberosity (Fig. 41.6). Here, the nerve lies deep to the gluteus
maximus muscle. Inferior to the border of the gluteus maximus muscle,
the nerve is relatively superficial and is generally easily blocked with a
posterior approach (Fig. 41.7). At the level of the lesser trochanter, the
sciatic nerve lies medial and posterior to the femur with the patient in
supine, frog-leg position (Fig. 41.8). Near the apex of or within the
popliteal fossa, the sciatic nerve divides into the tibial nerve, which
passes medially down the back of the leg and the common peroneal
nerve, which travels laterally and eventually wraps around the head of
the fibula (Figs. 41.9 and 41.10). The sciatic nerve can be blocked at
various levels along its course depending on the patient and the site of
the surgical procedure.
Subgluteal sciatic block: Patient Position: Supine, lateral, or prone.
Transducer: The choice of probe is dependent on the size of the patient
and the anticipated depth of the nerve from the skin. In children ≤30 kg,
the 25-or 38-mm, high-frequency (10 to 14 MHz) linear transducer is
suitable. In larger patients weighing ≥30 kg, the curved (11-mm)
intermediate-frequency probe (5 to 10 MHz) may be useful to improve
the depth of penetration.
Transducer Orientation: Transverse in the gluteal crease.

Needle: 24G, 1-inch needle; 22G, 2 inch-needle; or 21G, 4-inch


dependent on the size of the patient and the depth of the nerve from the
skin. An 18G insulated Tuohy needle with appropriate sized catheter can
be used for continuous techniques.
Local Anesthetic: 0.1 to 0.2 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine. Epinephrine in a concentration of 1:200,000 is added to the
solution.
Technique: The greater trochanter and ischium are palpated, and the
probe is located midway or slightly lateral on the line between these two
landmarks. The sciatic nerve is seen deep to the gluteus maximus muscle
(Fig. 41.6). The probe is moved distally to the gluteal crease where the
nerve is located between the quadratus femoris (lateral) and the biceps
femoris (medial) muscles. Some practitioners prefer to place the probe in
the gluteal crease at the start of the scan. In small patients, the patient is
usually positioned supine and the thigh and leg are elevated by an
assistant (Fig. 41.7A,B). In larger patients, the lateral position is
preferred (Fig. 41.7C to E). The skin is washed, and the probe is
positioned in a transverse orientation at the gluteal crease. The nerve is
imaged with the appropriate transducer, and the needle is inserted using
an in-line or out-of-plane approach (Fig. 41.7). The needle tip is
positioned adjacent to the nerve, and the local anesthetic is injected until
the nerve is surrounded. The needle may be inserted in line or out of
plane and advanced until it is adjacent to the nerve. Local anesthetic is
injected until the nerve is surrounded.
Anterior sciatic block: Patient Position: Supine with the operative extremity
in the frog-leg position.
Transducer: 25-mm probe oscillating at 8 to 13 MHz for smaller patients
or 11-mm, 6-to 10-MHz probe for larger patients.
Transducer Orientation: Transverse, 3 to 5 cm below the inguinal
ligament.

Needle: 24G, 1-inch needle; 22G, 2-inch needle; or 21G, 4-inch


dependent on the size of the patient and the depth of the nerve from the
skin. An 18G insulated Tuohy needle with appropriate sized catheter can
be used for continuous techniques.
Local Anesthetic: 0.1 to 0.2 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine. Epinephrine in a concentration of 1:200,000 is added to the
solution.
Technique: The anterior sciatic block is one of the most difficult blocks in
adults because of the distance from the skin to the target nerve.
However, in children, the sciatic nerve is usually only 3 to 5 cm deep to
the skin. This makes ultrasound-guided block very practical in the
pediatric population. The operative lower extremity is placed in the frog-
leg position (Fig. 41.8A to C). This maneuver rotates the sciatic nerve
from a position posterior to the femur into a position where the nerve is
now medial and deep to the femur. The ultrasound probe is placed 3 to 5
cm below the inguinal crease in a transverse orientation. The hypoechoic
shadow of the lesser trochanter is imaged. Medial and anterior to the
trochanter, the femoral artery is imaged. The sciatic nerve usually lies on
a line drawn perpendicularly through the artery at a depth that is 2 to 3
cm deep to the posterior border of the femur (Fig. 41.8C). In some
patients, the hyperechoic nerve lies in a position that is deep to the femur
at a point that is midway between the femur and the femoral artery. The
appropriate length needle is inserted in line at the medial end of the
probe. The needle is directed laterally until its tip is adjacent to the
nerve. The local anesthetic agent is injected until the nerve is
surrounded.
Midfemoral sciatic nerve block: Patient Position: Supine with the thigh and
leg internally rotated.
Transducer: 11-mm curved array oscillating at 6 to 10 MHz.
Transducer Orientation: Transverse, between the vastus lateralis and
biceps femoris muscles.

Needle: 24G, 2-cm needle; 22G, 4-cm needle; or 21G, 6-cm needle
dependent on the size of the patient and the depth of the nerve from the
skin. An 18G insulated Tuohy needle with appropriate sized catheter can
be used for continuous techniques.
Local Anesthetic: 0.1 to 0.2 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine. Epinephrine in a concentration of 1:200,000 is added to the
solution.
Technique: The skin is washed and a small curved probe (6 to 10 MHz) is
positioned in a transverse orientation between the vastus lateralis and
biceps femoris muscles. The nerve is a round hyperechoic structure
imaged posterior to the femur (Fig. 41.9). The needle is inserted superior
to the probe using an in-line approach and advanced at a 45-degree angle
to the skin until it is adjacent to the nerve. Local anesthetic is inserted
until the nerve is surrounded.
Popliteal block (sciatic at the knee): Patient Position: Supine, lateral, or
prone.
Transducer: 25-or 35-mm linear array oscillating at 8 to 13 MHz or 11-
mm curved array oscillating at 6 to 10 MHz.
Transducer Orientation: Transverse.

Needle: 24G, 1-inch needle; 22G, 2-inch needle; or 21G, 4-inch


dependent on the size of the patient and the depth of the nerve from the
skin. An 18G insulated Tuohy needle with appropriate sized catheter can
be used for continuous techniques.
Local Anesthetic: 0.1 to 0.2 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine. Epinephrine in a concentration of 1:200,000 is added to the
solution.
Technique: The patient can be positioned supine, lateral, or prone. If the
patient is positioned supine, the leg is elevated and bent 90 degrees at
the hip and at the knee. A modest amount of further extension at the
knee can be used to better delineate the tendons and the boundaries of
the popliteal fossa. The skin is washed, and the transducer is positioned
in a transverse orientation slightly above the level of the popliteal crease
(Fig. 41.10A,B). The popliteal artery and vein are imaged as is the tibial
nerve, which is an oval hyperechoic structure lying superficial to the
vein (Fig. 41.10C to E). The probe is advanced cephalad until the tibial
nerve is joined by the common peroneal nerve. This is generally several
centimeters above the popliteal crease. Once the junction of the two
nerves has been identified, the needle is inserted at the lateral end of the
probe using an in-line approach. When the tip of the needle is adjacent
to the round hyperechoic sciatic nerve, local anesthetic is injected until
the nerve is surrounded.
Reference
1. Kirchmair L, Lirk P, Colvin J, et al. Lumbar plexus and psoas major
muscle: not always as expected. Reg Anesth Pain Med.
2008;33:109–114.
Bibliography
Tobias JD, Mencio GA. Regional anesthesia for club foot surgery in children. Amer J Ther.
1998;5:273–277.
Manion SC, Tobias JD. Lumbar plexus blockade in children. Amer J Pain Manage. 2005;15:120–
126.
Johnson CM. Continuous femoral nerve blockade for analgesia in children with femoral fractures.
Can J Anaesth. 1994;22:281–283.
Tobias JD. Continuous femoral nerve block to provide analgesia following femur fractures in a
paediatric ICU population. Anaesth Intensive Care. 1994;22:616–618.
Gielen M, Viering W. 3-in-1 lumbar plexus block for muscle biopsy in malignant hyperthermia
patients. Amide anaesthetics may be used safety. Acta Anaesthesiol Scand. 1986;30:581–583.
Maccani RM, Wedel DJ, Melton A, et al. Femoral and lateral femoral cutaneous nerve block for
muscle biopsies in children. Paeditr Anaesth. 1995;5:223–227.
Ion T, Cook-Sather SD, Finkel RS, et al. Fascia iliaca block for an infant with arthrogryposis
multiplex congenital undergoing muscle biopsy. Anesth Analg. 2005;100:82–84.
Vincent CR, Turchiano J, Tobias JD. Fascia iliaca block for a muscle biopsy in an infant with
undiagnosed hypotonia. Saudi J Anesth. 2008;2:22–24.
Sanchez V, Segedin ER, Moser M, et al. Role of lumbar sympathectomy in the pediatric intensive
care unit. Anesth Analg. 1988;67:794–797.
Giaufre E, Dalens B, Gombert A. Epidemiology and morbidity of regional anesthesia in children:
a one-year prospective survey of the French-language society of pediatric anesthesiologists.
Anesth Analg. 1996;83:904–912.
Ecoffey C, Lacroix F, Gianfre E, et al. Epidemiology and morbidity of regional anesthesia in
children: a follow up one year prospective survey of the French-Language Society of
Paediatric Anaesthesiologists (ADARPEF). Pediatr Anesth. 2010;20:1061–1069.
Polaner D, Taenzer A, Walker B, et al. Pediatric Regional Anesthesia Network (PRAN): a multi-
institutional study of the use and incidence of complications of pediatric regional anesthesia.
Anesth Analg. 2012;115:1353–1364.
Dalens B, Tanguy A, Vanneuville G. Lumbar plexus block in children: a comparison of two
procedures. Anesth Analg. 1988;67:750–758.
Chayen D, Nathan H, Chayen M. The psoas compartment block. Anesthesiology. 1976;45:95–99.
Winnie AP, Ramamurthy S, Durrani X, et al. Plexus blocks for lower extremity surgery.
Anesthesiol Rev. 1974;1:11–16.
Kirchmair L, Enna B, Mitterschiffhaler G, et al. Lumbar plexus in children. A sonographic study
and its relevance to pediatric regional anesthesia. Anesthesiology. 2004;101:445–450.
Dadure C, Bringuier S, Mathieu O, et al. Continuous epidural block versus continuous psoas
compartment block for postoperative analgesia after major hip or femoral surgery in children:
a prospective comparative randomized study. Ann Fr Anesth Reanim. 2010;29:610–615.
Capdevila X, Macaire P, Dadure C, et al. Continuous compartment block for postoperative
analgesia after total hip arthroplasty new landmarks, technical guidelines and clinical
evaluation. Anesth Analg. 2002;94:1606–1613.
Kirchmair L, Lirk P, Colvin J, et al. Lumbar plexus and psoas major muscle: not always as
expected. Reg Anesth Pain Med. 2008;33:109–114.
Flack S, Anderson C. Ultrasound guided lower extremity blocks. Pediatr Anesth. 2012;22:72–80.
Dadure C, Capdevila X. Peripheral catheter techniques. Pediatr Anesth. 2012;22:93–101.
Dalens B, Vanneuville G, Tanguy A. Comparison of the fascia iliaca compartment block with the
3-in-1 block in children. Anesth Analg. 1989;69:705–713.
Farid I, Heiner E, Fleissner P. Comparison of femora nerve block and fascia iliaca block for
analgesia following reconstructive knee surgery in adolescents. J Clin Anesth. 2010;22:256–
259.
Oberndorfer U, Marhofer P, Bosenberg A, et al. Ultrasonographic guidance for sciatic and femoral
nerve blocks in children. Br J Anaesth. 2007;98:797–801.
Bodner G, Bernathova M, Galiano K, et al. Ultrasound of the lateral femoral cutaneous nerve.
Normal findings in a cadaver and in volunteers. Reg Anesth Pain Med. 2009;34:265–268.
Gray AT, Collins AB, Schafhalter-Zoppoth I. Sciatic nerve block in a child: a sonographic
approach. Anesth Analg. 2003;97:1300–1302.
Ponde V, Desai AP, Dhir S. Ultrasound-guided sciatic nerve block in infants and toddlers produces
successful anesthesia regardless of motor response. Pediatr Anesth. 2010;20:633–637.
Schwemmer U, Markus CK, Greim CA, et al. Sonographic imaging of the sciatic nerve and its
division in the popliteal fossa in children. Pediatr Anesth. 2004;14:1005–1008.
Tsui BC, Santhanam S. Ultrasound imaging for regional anesthesia in infants, children, and
adolescents. Anesthesiology. 2010;112:473–492.
42
Ultrasound-Guided
Epidural Block
(Caudal, Lumbar, and
Thoracic), Truncal
and Paravertebral
Blocks in Children
Part 1: Introduction

KAREN BORETSKY, PAUL E. BIGELEISEN,


ARVIND CHANDRAKANTAN, HADI S. MOTEN,
AND MIHAELA VISOIU

Epidural anesthesia continues to occupy a major role in pediatric


regional anesthesia and represents almost half of the regional anesthetics
in children younger than 3 years of age.1 In parallel with the use of
ultrasound in all regional anesthesia, the use of ultrasound for guidance
of epidural anesthesia has evolved quickly. When imaging and accessing
the neuraxial structures in infants and children, it is important to be
aware of the developmental changes in the spine and the spinal canal
that affect the ultrasound images.
The location of the spinal cord and dural sac within the spinal canal
changes over time. During the development of the fetus in utero, the
spinal cord extends the entire length of the spinal canal. This position
ascends as the fetus develops, and in a newborn, the spinal cord extends
as low as the L3 level with termination of the dura sac at the S4 level.2
As the child grows, the vertebral column grows faster than the spinal
cord and the termination of the spinal cord moves to a progressively
higher level, L1/L2 for the spinal cord and S2 for the dura at 1 year of
age.2,3 These may be important details to keep in mind in order to avoid
complications such as inadvertent dural puncture and/or spinal cord
trauma.3
The limited ossification of the pediatric vertebral column allows better
visualization of the neuraxis in pediatric patients. The newborn skeleton
is predominantly cartilage at birth and proceeds to deposit calcium and
ossify progressively until puberty. This relative lack of ossification
allows the easy penetration of ultrasound waves and a clear view of the
neuraxial structures up until 3 months of age, after which ultrasound
imaging becomes progressively more difficult.4 Around 9 months of
age, the bony mass of the vertebral arches start to impose ultrasound
shadows over the underlying structures, but anatomic information
remains available in the ultrasound windows between adjacent vertebrae.
Ultrasound can be used in multiple ways to help with accessing the
pediatric epidural space at different levels. Prepuncture scanning gives
information about the location and depth of the epidural space and
adjacent structures. Alternatively, ultrasound imaging is used for real-
time scanning to guide and direct the needle and catheter insertion into
the epidural space. The ultrasound imaging is also used after the catheter
or needle is inserted to confirm position.
References
1. Ecoffey C, Lacroix F, Giaufre E, et al. Epidemiology and morbidity
of regional anesthesia in children: a follow-up one-year prospective
survey of the French-Language Society of Paediatric
Anaesthesiologists (ADARPEF). Paediatr Anaesth. 2010;20:1061–
1069.
2. Johr M, Berger TM. Caudal blocks. Paediatr Anaesth. 2012;22:44–
50.
3. Malas MA, Seker M, Salbacak A, et al. The relationship between
the lumbosacral enlargement and the conus medullaris during the
period of fetal development and adulthood. Surg Radiol Anat.
2000;22:163–168.
4. Willschke H, Marhofer P, Bosenberg A, et al. Epidural catheter
placement in children: comparing a novel approach using
ultrasound guidance and a standard loss-of-resistance technique. Br
J Anaesth. 2006;97:200–207.
Part 2: Single-Injection Caudal
Approach to the Epidural Space

KAREN BORETSKY, PAUL E. BIGELEISEN,


ARVIND CHANDRAKANTAN, HADI S. MOTEN,
AND MIHAELA VISOIU

Background and indications: The relative popularity of epidural


analgesia can be attributed to the relative ease with which the epidural
space can be accessed via the sacral hiatus in infants and young children.
A single injection of local anesthetic into the caudal epidural space, the
“kiddie caudal,” is appropriate for most surgeries below the umbilicus
and results in low complication rates and high success rates. Inadvertent
spinal anesthesia, with an incidence of dural puncture of 0.19%,1 and
perforation of adjacent viscera from improper position of the needle
have, however, been reported. Ultrasound guidance offers a way to
identify sacrum anatomy and perform caudal blocks in patients at risk
for dural puncture (neonates and very small infants) and patients with
difficult anatomy (older patients and patients with lumbosacral spinal
dysraphism), potentially increasing the success rate and decreasing the
rate of complications.2–4
Anatomy: The sacrum is composed of five fused sacral vertebrae (Fig.
42.1). Its dorsal surface is convex and has a raised interrupted median
crest with four (sometime three) spinous tubercles representing fused
sacral spines. The posterior surface is formed by fused laminae. Below
the fourth (or third) spinous tubercle, an arched sacral hiatus is identified
due to failure of the fifth pair of laminae to meet, exposing the sacral
canal, a continuation of the lumbar spinal canal. This caudal opening of
the canal, or sacral hiatus, is covered by the sacrococcygeal ligament,
which is an extension of the ligamentum flavum. The fifth inferior
articulate processes project caudally and flank the sacral hiatus as
prominent bony sacral cornua. The sacral canal contains the cauda
equina (including the filum terminale) and the spinal meninges, epidural
venous plexus, and adipose tissue. The lowest margin of the filum
terminale emerges at the sacral hiatus and transverses the dorsal surface
of the fifth sacral vertebra.

The sacrum structures are visualized in both the short axis (transverse)
and the midline long axis (longitudinal).
Short axis: The two bony prominences of the sacral cornua appear as
hyperechoic reversed U-shaped structures (Figs. 42.2 and 42.3). Two
hyperechoic bandlike structures lie between the two cornua. The
bandlike structure on top of the sacral cornua is the sacrococcygeal
ligament, and the bandlike structure at the bottom is the bone of the
dorsal surface of the anterior sacrum. The sacral hiatus is the hypoechoic
region between the two bandlike structures.
Long axis: With the caudad edge of the transducer resting between the
two cornua, the most prominent rounded hyperechoic structure observed
is the bony prominence of the S4 spinous tubercle of the sacrum (Figs.
42.4 and 42.5). The sacrococcygeal ligament presents like a thick band
beyond the end of the S4 spinous process, and the sacral hiatus is the
hypoechoic region under the sacrococcygeal ligament.
Patient position: Prone or lateral position with hips, knees, and neck
flexed.
Transducer: 25-to 35-mm linear array oscillating at 6 to 13 MHz.
Transducer orientation: Initially transverse over the two bony
prominences of sacral cornua to identify sacrum structure. Then the
transducer is rotated 90 degrees for real-time and/or confirmatory needle
placement.

Needle: 22G, 4-to 5-cm blunt needle.


Local anesthetic: 1 mL/kg of ropivacaine 0.2% or bupivacaine 0.25%,
maximum 20 mL.
Technique: After proper disinfection of the skin, the sterile covered
transducer is placed with a transverse orientation over the sacrum at the
level where the sacral cornua can be palpated, the S4–S5 level. The
sacral hiatus is identified as the hypoechoic space between the
hyperechoic upside down U-shaped structures of the sacral cornua deep
to the sacrococcygeal ligament. The probe is rotated to the longitudinal
position while still visualizing the sacral hiatus, and the lower edge of
the probe is placed at the level of the cornua. Under ultrasound
longitudinal view, the needle is inserted through the sacrococcygeal
ligament from a caudad to cephalad direction using an in-plane
approach. The needle appears as a hyperechoic structure positioned in
the sacral hiatus. The local anesthetic administration is observed in real
time as a turbulence moving cephalad in the sacral hiatus. A final
transition back to a transverse view may permit visualization of the local
anesthetic between the sacrococcygeal ligament and the sacrum.
Tips 1. A prone position with rolls placed transversely under the hips
achieves a flatter skin surface at the sacral hiatus area for the placement
of the ultrasound transducer.
2. In the lateral position with hips, knees, and neck flexed, the dural sac
shifts significantly cephalad, providing some safety margin to avoid
dural puncture.3
The technique can be done under the ultrasound transverse view,
using an out-of plane technique.
3. The optimal angle for needle insertion is between 10 and 38 degrees.5
4. The characteristic “give” or “pop” can be detected when the
sacrococcygeal ligament is penetrated.
5. The dosages for a single-injection caudal anesthesia are based on the
site of surgery and the patient’s age and weight. In older children
weighing more than 20 kg, the spread of local anesthetic after a
single caudal injection is less reliable.

References 1. Beyaz SG, Tokgoz O, Tufek A. Caudal epidural block


in children and infants: retrospective analysis of 2088 cases. Ann
Saudi Med. 2011;31:494–497.
2. Chen CP, Tang SF, Hsu TC, et al. Ultrasound guidance in caudal
epidural needle placement. Anesthesiology. 2004;101:181–184.
3. Koo BN, Hong JY, Kim JE, et al. The effect of flexion on the level
of termination of the dural sac in paediatric patients. Anaesthesia.
2009;64:1072–1076.
4. Visoiu M, Lichtenstein S. 25 years of experience, thousands of
caudal blocks, and no dural puncture. What happened today?
Pediatr Anesth. 2012;22:304–305.
5. Park JH, Koo BN, Kim JY, et al. Determination of the optimal angle
for needle insertion during caudal block in children using
ultrasound imaging. Anaesthesia. 2006;61:946–949.
Part 3: Lumbar and Thoracic
Epidural Catheters via the Caudal
Approach

KAREN BORETSKY, PAUL E. BIGELEISEN,


ARVIND CHANDRAKANTAN, HADI S. MOTEN,
AND MIHAELA VISOIU

Background and indications: In the smallest patients, accessing the


epidural space via a direct approach at the desired lumbar or thoracic
level is technically challenging because of the diminutive size, the close
proximity of vulnerable anatomic structures, and the fear of
complications such as spinal cord trauma. Since 1988, a technique of
accessing the epidural space via the sacral hiatus and threading a
catheter cephalad to the desired dermatomal level is frequently used on
infants and young children.1,2 Ultrasound imaging can be used to
confirm placement of the needle and catheter through the sacral hiatus
into the epidural space and to confirm passage of the catheter to the
appropriate lumbar or thoracic level.
Anatomy: The vertebral column contains 12 thoracic and 5 lumbar
vertebrae bones (Fig. 42.6). Each vertebra is composed of a body
anteriorly and an arch posteriorly, which enclose a canal housing the
spinal cord. Protruding from the posterior midline of each arch is the
spinous process, and extending laterally are the paired transverse
processes. The spinal cord lies within the vertebral canal and is covered
by three membranes. The two outermost layers, the dura mater and the
arachnoid mater, are closely adherent. The space between the arachnoid
mater, and the third layer, the pia mater, is the subarachnoid space,
which contains the cerebrospinal fluid (CSF). The spinal cord is thickest
at the thoracic level and decreases in size as it sheds nerve roots and
tapers to its termination in a bundle of loose nerve roots called the cauda
equina.3 The ligamentum flavum is a tough ligament that extends from
C1 to S1 and connects the vertebrae to each other. The potential space
created between the ligamentum flavum and the dura is the epidural
space. The anatomy and sonoanatomy of the sacral hiatus and caudal
canal are described earlier.

The lumbar and thoracic spine structures can be visualized in both the
short-axis/transverse view (Fig. 42.7) and the long-axis/longitudinal
view (Fig. 42.8).
Short axis: The spinous process is represented as a hyperechoic upside
down V-shaped structure with an acoustic shadow below (Figs. 42.6 and
42.7). The spinal cord and dura are seen as two hyperechoic concentric
circles deep to the spinous process. The hypoechoic spinal cord cased by
the hyperechoic pia is represented by the innermost circle. The CSF is
the hypoechoic concentric rim, and the dura is the outermost
hyperechoic circle. The ligamentum flavum is difficult to visualize and
usually appears indistinguishable from the dura. The ligamentum flavum
can best be seen when separated from the dura by hydrodissection into
the epidural space (Fig. 42.9).

Long axis: The long axis can be best visualized from a longitudinal
paramedian view (Fig. 42.8 and 42.10). The spinous processes are now
visualized as thick, slanted hyperechoic lines creating acoustic windows
occurring at regular intervals. Between these acoustic windows, the
elements of the spinal canal can be identified by layer. The dura mater is
represented by the topmost (anatomically posterior) hyperechoic line
between the acoustic windows. The hypoechoic layer beneath this
represents the CSF, below which the hyperechoic pia mater is seen. The
appearance of the neural element layer is dependent on the vertebral
level being viewed. It appears as either a homogenous hyperechoic area
representing the solid spinal cord or a bundle of hyperechoic linear
structures representing the cauda equina. The vertebral bodies can be
identified ventrally. The degree of acoustic shadowing cast by the
spinous processes depends on the amount of ossification.
Patient position: Prone with rolls placed transversely under the hips and
shoulders to allow free excursion of the abdomen and flatten the lumbar
curve (Fig. 42.11). Alternatively, a lateral position with the patient
flexed forward and curled inward can be used.
Transducer: 25-to 50-mm linear array oscillating at 6 to 17 MHz.
Transducer orientation • Transverse: Visualize/locate sacral cornua.
• Longitudinal: Visualize needle and catheter insertion.
• Transverse: Confirm catheter tip location.

Needle: 18G 5-cm Tuohy needle.


Local anesthetic: 0.75 to 1 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine followed by an infusion of 0.2 to 0.25 mg/kg/hr of
ropivacaine or 0.25% bupivacaine (<4 months of age) or 0.4 to 0.5
mg/kg/hr of 0.2% ropivacaine or 0.25% bupivacaine (>4 months of age).
Technique: Palpate and mark presumed position of the sacral cornua and
desired lumbar or thoracic level for final catheter tip location. Using
sterile technique, prep and drape a field to include the top of the
intergluteal crease to the desired dermatomal level for tip. Using an
ultrasound probe with a sterile sheath, locate the cornua (2 upside down
U’s) and sacral hiatus via the short-axis view and rotate the probe 90° to
a long-axis view while keeping the sacral hiatus in sight. Position the
lower end of the probe at the level of the sacral cornua. Position and
insert the needle as described above for the single-shot technique and
carefully advance under ultrasound guidance. The needle will appear as
a hyperechoic structure. Preservative-free normal saline (2 to 5 mL) is
injected and used to confirm placement and open up the epidural space.
With the probe still in a longitudinal orientation, the catheter is inserted
through the needle and visualized entering the epidural space. The
hyperechoic catheter can sometimes be continually visualized during
insertion by sliding the transducer cephalad in the longitudinal
orientation along the vertebral column as the catheter is inserted. The
ability to follow the catheter insertion in real time continually to the
desired dermatome is difficult and becomes increasingly less reliable as
the vertebrae ossify. Once the catheter is inserted to the desired
dermatomal distance, the tip location can be confirmed by direct
visualization of the tip if it contains sonoreflective material such as
metal or by administration of small amounts of preservative-free saline
under ultrasound visualization in the short-axis view and viewing the
expansion of the epidural space and displacement of tissue. With
increasing age, visualization will rely more on information contained in
ultrasound windows between vertebra. Caudal catheters should be
tunneled subcutaneously to reduce the risk of bacterial colonization to
equal that of lumbar epidural catheters.4
Tips • View both the long axis and short axis to provide two points of
reference.
• The epidural space is a potential space that is visualized best when the
space is hydrodissected.
• The lumbar spine offers better ultrasound visualization than the
thoracic spine.3
• Before insertion of the catheter, measure the distance to the desired
dermatome and insert the catheter initially the measured distance. If
the catheter does not coil, it can then be found near the desired
dermatome and small adjustments more easily made.
References
1. Bösenberg AT, Bland BA, Schulte-Steinberg O, et al. Thoracic
epidural anesthesia via caudal route in infants. Anesthesiology.
1988;69:265–269.
2. Marhofer P, Bosenberg A, Sitzwohl C, et al. Pilot study of neuraxial
imaging by ultrasound in infants and children. Paediatr Anaesth.
2005;15:671–676.
3. Malas MA, Seker M, Salbacak A, et al. The relationship between
the lumbosacral enlargement and the conus medullaris during the
period of fetal development and adulthood. Surg Radiol Anat.
2000;22:163–168.
4. Bubeck J, Boos K, Krause H, et al. Subcutaneous tunneling of
caudal catheters reduces the rate of bacterial colonization to that of
lumbar epidural catheters. Anesth Analg. 2004;99:689–693, table of
contents.
Part 4: Lumbar and Thoracic
Continuous Epidural Analgesia via
the Direct, Intervertebral Approach

KAREN BORETSKY, PAUL E. BIGELEISEN,


ARVIND CHANDRAKANTAN, HADI S. MOTEN,
AND MIHAELA VISOIU

Background and indications: For pediatric patients, the loss-of-


resistance (LOR) technique using saline solution is the classic technique
for the placement of the epidural catheters directly at lumbar or thoracic
levels. Ultrasonography with a linear high-frequency ultrasound probe
can be used to identify neuraxial structures and to measure skin-to-
epidural space and dura depth, and it may permit visualization of the
local anesthetic spread in epidural space and confirmation of catheter
position.1–3
Anatomy: Anatomy and sonoanatomy of spine were described earlier.

Needle: 18G 5-to 10-cm Tuohy needle.


Technique: After proper disinfection of the skin using skin antiseptic, the
sterile covered transducer is placed with a paramedian/longitudinal
orientation (Fig. 42.12) to identify the longitudinal view of spinous
process, dura, and epidural space (Fig. 42.13). The distance from the
skin to the ligamentum flavum/epidural space and from the skin to dura
matter should be measured. These distances should be considered when
the Tuohy needle is inserted for the LOR technique. After the placement
of the epidural catheter, the ultrasound transducer is placed with a
transverse orientation at the level where the tip of the catheter is
expected to be found. A short/transverse ultrasound view of the spine
was presented earlier. The spine is better visualized in infants than in
older children (Figs. 42.14 and 42.15). Injection of normal saline/local
anesthetic through the epidural catheter will visualize the downward
displacement of dura. If the transducer is moved longitudinally above
and below the injection sites, the local anesthetic in the epidural space is
confirmed by the displacement of the dura as the epidural space is
expanded by the injected local anesthetic.
Ultrasound imaging can be used to place an epidural catheter under
direct visualization.4 With the transducer placed with a
paramedian/longitudinal orientation, the needle is inserted using a
midline approach and can be visualized penetrating ligamentum flavum
(if visible) and lying on the top of dura. The injection of local anesthetic
through the Tuohy needle in the epidural space can be visualized during
continuous technique for LOR. A final transition back to a transverse
view may permit visualization of the local anesthetic as described
earlier. A second person is needed to manipulate the ultrasound probe.
Local anesthetic: 0.5 to 0.7 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine, for lumbar approach and 0.4 to 0.5 mL/kg for thoracic
approach, for a maximum of 1.7 mg/kg for ropivacaine and 1.5 mg/kg
for bupivacaine.
Tips 1. Detection of epidural space using ultrasound can be done as a
prescan procedure for obese pediatric patients or for the patients with a
possible abnormal vertebral column.
2. The dura mater is more easily visualized than ligamentum flavum.
3. Ultrasound provides better visualization for catheter placement in
children younger than 6 months of age.5
4. Ultrasound visualization of the neuraxial structures at the thoracic
spine is more difficult than ultrasonography at the lumbar spine
secondary to more narrower acoustic window at thoracic level.
5. A correlation of 0.88 could be estimated between ultrasound depth of
the epidural space and measured LOR.3

References 1. Vallejo MC, Phelps AL, Singh S, et al. Ultrasound


decreases the failed labor epidural rate in resident trainees. Int J
Obstet Anesth. 2010;19:373–378.
2. Kil HK, Cho JE, Kim WO, et al. Prepuncture ultrasound-measured
distance: an accurate reflection of epidural depth in infants and
small children. Reg Anesth Pain Med. 2007;32:102–106.
3. Rapp HJ, Folger A, Grau T. Ultrasound-guided epidural catheter
insertion in children. Anesth Analg. 2005;101:333–339, table of
contents.
4. Willschke H, Marhofer P, Bosenberg A, et al. Epidural catheter
placement in children: comparing a novel approach using
ultrasound guidance and a standard loss-of-resistance technique. Br
J Anaesth. 2006;97:200–207.
5. Chawathe MS, Jones RM, Gildersleve CD, et al. Detection of
epidural catheters with ultrasound in children. Paediatr Anaesth.
2003;13:681–684.
Part 5: Ultrasound-Guided Regional
Anesthesia of the Thorax, Trunk, and
Abdomen in Infants and Children

TARUN BHALLA, RALPH BELTRAN, DAVID P.


MARTIN, STEPHEN LUCAS, PAUL E.
BIGELEISEN, AND JOSEPH D. TOBIAS

Depending on the site of injury and the nerves that need to be


anesthetized, there are several approaches for regional anesthesia of the
thorax, trunk, and abdomen that are applicable to the pediatric
population. Examples of these techniques are the use of thoracic
paravertebral or intercostal blockade for analgesia following
thoracotomy, ilioinguinal/iliohypogastric blockade following inguinal
herniorrhaphy, and rectus sheath blockade following umbilical
herniorrhaphy. Additional techniques that have been used most
commonly in the adult population but are now finding their way into the
pediatric population include lumbar paravertebral blockade for analgesia
following renal surgery and the transversus abdominis plane or TAP
block for analgesia following abdominal wall incisions.
In the past, analgesia following many of these procedures was most
commonly accomplished with the use of an epidural block at the caudal,
lumbar, or thoracic level; however, given data demonstrating the efficacy
and lower risk of adverse events with peripheral nerve blockade versus
caudal epidural block, it is likely that there will be an increasing use of
peripheral nerve blockade of the lower extremity.1 Additionally,
blockade of the nerves of the thorax, trunk, and abdomen (which are
described in this chapter) also have the advantage of providing unilateral
blockade, thereby eliminating the adverse effects associated with
bilateral epidural blockade, including sympathectomy. To close out this
chapter, the use of ultrasound guidance for caudal epidural blockade is
reviewed. Although used less frequently since the advent of peripheral
blockade, there are various clinical scenarios in which the caudal
epidural block may still be the first choice, such as the former preterm
infant in whom a caudal epidural block is used to provide surgical
anesthesia during inguinal herniorrhaphy, thereby avoiding the potential
deleterious adverse effects of general anesthesia.
Transversus abdominis block Background and Indications: The lateral
abdominal wall contains three muscle layers including the external oblique, the
internal oblique, and the transversus abdominis muscles and their associated
fascial sheaths. The lower thoracic and upper lumbar nerves provide sensory
innervation of the skin, muscles, and parietal peritoneum of the anterior
abdominal wall. These nerves course in a plane between the transversus
abdominis and internal oblique muscles. Given the anatomic localization of
these nerves, McDonnell et al.2 described a unique approach that allows for
blockade of these nerves with the administration of local anesthetic agents in the
plane between the transversus abdominis and internal oblique muscles with a
single injection administered in the triangle of Petit.3–7 The triangle of Petit is
bounded posteriorly by the latissimus dorsi muscle, anteriorly by the external
oblique muscle, and inferiorly by the iliac crest. In clinical practice, the TAP
block is placed by using an ultrasound-guided technique with the ultrasound
probe placed in the axial plane just above the iliac crest. A needle is inserted in
line with the probe so that it can be demonstrated that the needle lies in the
correct fascial plane prior to injection of the local anesthetic solution. The
potential utility of the TAP block has been demonstrated by McDonnell et al.2 in
a cadaveric and radiologic evaluation. In a follow-up cadaveric evaluation,
McDonnell described the varying anatomic location of the triangle of Petit and
therefore recommends placing the block needles close to the midaxillary line as
possible when attempting a TAP block.4 Using a double pop or LOR technique
with a blunt block needle in a cadaver model, the authors demonstrated that
methylene blue could be injected between the transversus abdominis and internal
oblique muscles. The correct anatomic location of the dye was demonstrated by
dissection of the cadaver specimens. This was followed by the demonstration of
radiopaque dye in the correct fascial plane using computed tomography and
magnetic resonance imaging in three healthy adult volunteers.
In the adult population, the TAP block has been shown to provide
effective analgesia following various types of lower abdominal
procedures, including retropubic prostatectomy, Cesarean section, and
total abdominal hysterectomy.5–7 In a prospective randomized trial of 50
adults following Cesarean delivery, TAP block with 0.2 mL/kg of 0.75%
ropivacaine on each side resulted in decreased postoperative pain scores,
delayed request for postoperative analgesia, and decreased morphine use
during the initial 48 postoperative hours.5 The median time to first
request for postoperative analgesia was 90 minutes in the control group
and 220 minutes in patients who received a TAP block. Morphine use
during the initial 48 hour postoperative period was decreased by 70% in
patients who received a TAP block (66 ± 26 mg in control patients
versus 18 ± 14 mg in patients who received a TAP block, P < .001). The
same investigators evaluated the efficacy of the TAP block following
total abdominal hysterectomy in 50 women.5 After anesthetic induction
and prior to surgical incision, a bilateral TAP block was placed using 0.2
mL/kg of 0.75% ropivacaine on each side. Patients who received a TAP
block had decreased postoperative pain scores, delayed request for
postoperative analgesia, and decreased morphine use during the initial
48 postoperative hours (55 ± 17 mg in control patients versus 27 ± 20
mg in patients who received a TAP block, P < .001). There continues to
be literature published on the multifaceted nature of the TAP block.
More recently, the subcostal technique has been described. This
approach is aimed along the midaxillary line in the subcostal space,
slightly cephalad to the triangle of Petit. This may be beneficial for
surgical incisions that may traverse the dermatomal regions of nerve
roots from T9 to T11.
There is also a surge of data regarding the use of TAP block in infants
and children.8–11 Mukhtar and Singh9 reported the successful use of TAP
block to provide analgesia following laparoscopic appendectomy in four
patients 14 to 17 years of age. A bilateral TAP block was placed using 20
mL of 0.25% bupivacaine per side with ultrasound guidance. No patient
required supplemental analgesic agents for the initial 12 postoperative
hours with pain scores ranging from 0 to 2. Two patients required no
analgesic agents during their postoperative course. Unilateral TAP block
has also been shown to provide effective analgesia for inguinal hernia
repair in a cohort of eight children, and anecdotal success demonstrated
the efficacy of a TAP block in a 3.6-kg infant with VACTERL syndrome
undergoing colostomy placement on day of life 2.11,12 In the latter case,
the TAP block was chosen instead of caudal epidural blockade due to
associated vertebral anomalies. The authors have reported successful
preliminary experience with the TAP block following various surgical
procedures involving the umbilicus and lower abdomen in a cohort of 10
pediatric patients, ranging in age from 10 months to 8 years.13 Effective
postoperative analgesia was achieved in 8 of the 10 patients. The adult
and pediatric data suggest that a TAP block can be used for major open
abdominal surgeries involving the umbilicus and lower abdominal area.
The procedure is also useful for open appendectomies, colostomy
placement, and transverse lower abdominal incisions (bladder and
ureteral surgery) as well as to cover the access ports for various types of
laparoscopic procedures.
Patient position: Supine.
Transducer: 25-or 35-mm linear transducer oscillating at 8 to 13 MHz.
Transducer orientation: Transverse between the ribs and iliac crest in
the triangle of Petit in the midaxillary line.

Needle: 24G, 1-inch or 22G, 2-inch blunt or insulated needle. An 18G


insulated Tuohy needle with appropriate sized catheter can be used for
continuous techniques.
Local anesthetic: <0.1 to 0.2 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine. Epinephrine in a concentration of 1:200,000 is added to the
solution.
Technique: The skin between the iliac crest and the lower margin of the
12th rib over the triangle of Petit in the midaxillary line is washed, and
sterile gel is applied. A 25-or 35-mm linear transducer oscillating at 10
to 13 MHz is oriented in the transverse plane. The external oblique,
internal oblique, and transversus abdominis muscles are identified. The
bowel can be seen immediately below the transversus abdominis muscle.
For surgeries of the upper abdominal wall, the probe is moved cephalad
and anterior until it is immediately below the costal margin. For lower
abdominal surgeries, the probe is placed immediately above the iliac
crest. The needle is inserted in line and anterior to the end of the
ultrasound probe. The needle is directed to the fascial plane between the
internal oblique and transversus abdominis muscles (Figs 42.16 to
42.18). Once the proper fascial space has been entered, the local
anesthetic is injected in divided doses. The procedure is repeated on the
opposite side for midline or bilateral incisions. A catheter may be
inserted after the plane between the muscles has been dilated.
Rectus abdominis (paraumbilical or rectus sheath) block Background and
Indications: Smith et al.14 first utilized the rectus sheath (RS) block for adult
laparoscopic gynecologic procedures. In 1996, Ferguson et al.15 suggested the
use of the RS block to provide analgesia for umbilical hernia repair in the
pediatric population. Courreges et al.16 described a new technique for providing
analgesia for children undergoing umbilical hernia repair, naming it the
paraumbilical block. The difference in these techniques is mainly based on the
site of needle injection, with the latter block being placed using four injections
(two on each side) above and below the level of the umbilicus rather than two
injections (one on each side) above the level of the umbilicus.
Willschke et al.17 were the first to report the use of ultrasound-guided
RS blocks in the pediatric population. The study started with a
sonoanatomic evaluation in 30 children, which noted a poor correlation
between the depth of the posterior RS and the weight of the child, body
surface area, or the height. In the second part of the study, the authors
demonstrated the efficacy of RS block placed using ultrasound guidance
for the deposition of 0.1 mL/kg of 0.25% levobupivacaine in providing
analgesia following umbilical hernia surgery in 20 children.
The periumbilical area is innervated by the right and left
thoracoabdominal intercostals nerves, which are derived from the
anterior rami of spinal roots T8–T12. These nerves travel in the plane
between the internal oblique and the transversus abdominis muscles. The
transversus abdominis muscle ends in a tendon sheath that forms the
posterior wall of the RS, which runs posteriorly to the rectus abdominis
muscle. The anterior wall of the RS is formed by the merger of the
tendons of the internal and external oblique muscles. The anterior and
posterior walls of the rectus sheath encase the rectus abdominis muscle
starting at the linea alba laterally and joining in the midline of the
abdomen into a single sheath. The internal costal nerves travel in the
space behind the rectus abdominis muscle and the posterior wall of the
RS. The nerve perforates the sheath near the midline of the abdomen
after traveling behind the rectus abdominis muscle. It ends as the
anterior cutaneous branch supplying the area around the midline of the
abdomen. Alternatively, the anterior cutaneous branch arises before the
rectus abdominis muscle and travels over the top of the muscle in the
subcutaneous space. Given this anatomy, these nerves can be blocked by
the placement of local anesthetic agent above and behind the rectus
abdominis muscle.
De Jose Maria et al.18 recently described a new approach to the classic
technique. The modified technique included avoidance of the epigastric
vasculature as well as isolating the 10th intercostal nerve at the lateral
edge of the rectus abdominis muscle before the anterior cutaneous
branching points. Ten children scheduled for umbilical hernia repair
were included and underwent a bilateral RS block as described earlier.
The needle was localized to the lateral edge of the rectus muscle at the
junction of the aponeurosis of the internal oblique and transversus
abdominis muscles, using ultrasound guidance followed by the
administration of 0.1 mL/kg of 0.25% bupivacaine. Effective analgesia
was achieved in all patients without the need for supplemental
intraoperative or postoperative opioids. Gurnaney et al.19 described a
prospective randomized and observer blinded study comparing RS
blocks versus local anesthetic infiltration by surgeon in a total of 52
patients. The study demonstrated that the ultrasound-guided RS block
provides superior analgesia in the perioperative period compared with
infiltration of the surgical site after umbilical hernia repair. There was a
statistically significant difference in the perioperative opioid medication
consumption between the local infiltration group (mean: 0.13 mg/kg,
confidence interval [0.09 to 0.17 mg/kg]) and the RS block group (mean:
0.07 mg/kg, confidence interval [0.05 to 0.09 mg/kg]) (P = .008). When
the investigators compared the postoperative opioid consumption
between the local infiltration group (mean: 0.1 mg/kg, 95% confidence
interval [0.07 to 0.13 mg/kg]) and the RS block group (mean: 0.07
mg/kg, 95% confidence interval [0.05 to 0.09 mg/kg]) (P = .09), there
was a trend towards statistical significance between the two groups.
Patient position: Supine.
Transducer: 25-or 35-mm linear transducer oscillating at 8 to 13 MHz.
Transducer orientation: Transverse, lateral to the midline.

Needle: 24G, 1-inch or 22G, 2-inch blunt or insulated needle. An 18G


insulated Tuohy needle with appropriate sized catheter can be used for
continuous techniques.
Local anesthetic: 0.1 to 0.2 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine. Epinephrine in a concentration of 1:200,000 is added to the
solution.
Technique: The skin over the umbilicus is washed, and sterile gel is
placed medial and lateral to the umbilicus. The ultrasound probe
oscillating at 10 to 13 MHz is placed in an axial orientation 2 to 5 cm
lateral to the midline and slightly above the level of the umbilicus. Using
ultrasound, the anterior and posterior borders of the RS are identified.
Note that the peritoneum and bowel are immediately below the posterior
boundary of the posterior RS. The needle is inserted in line at the lateral
end of the probe (Figs. 42.19 and 42.20). Note, the needle will pass
through the rectus abdominis muscle during performance of this block.
The needle is directed to the edge of the sheath and the local anesthetic
is deposited between the posterior RS and the posterior border of the
rectus abdominis muscle. Half of the local anesthetic is placed posterior
to the muscle, and the other half is placed anteriorly. Alternatively, the
entire volume can be placed behind the rectus abdominis muscle, as it
will generally spread in the space and cover the muscle anteriorly and
posteriorly. The procedure is repeated on the contralateral side.
Ilioinguinal and iliohypogastric block Background and Indications:
Ilioinguinal/iliohypogastric (IL-IH) nerve block provides analgesia to the
inguinal area for inguinal hernia repair, orchiopexy, and hydrocelectomy. This
block has been shown to be equally effective as a caudal block for these
procedures.20–22 However, the IL-IH block cannot be used as the sole anesthetic
instead of general anesthesia for groin surgery, as the stress response and visceral
pain from peritoneal traction and manipulation of the spermatic cord are not
covered by an IL-IH block. The ilioinguinal (L1) and iliohypogastric (T12 and
L1) nerves originate from the lumbar plexus and provide innervation to the
scrotum and inner thigh. The roots travel in the space between the iliacus and
psoas muscle. More distally along their course, these nerves pierce the
transversus abdominis muscle near the anterior superior iliac spine (ASIS) and
travel in the plane between the transversus abdominis and internal oblique
muscles. Complications from IL-IH blockade are rare; however, there have been
reports of colonic and small bowel puncture when these blocks are placed
blindly without ultrasound guidance. Recently, a case report was published
regarding a large retroperitoneal hematoma on an adult occurring subsequent to
an IL-IH block.23
The block was performed using only a landmark technique, which
resulted in inadvertent perforation of the right deep circumflex iliac
artery. The patient was also receiving antiplatelet medications (aspirin
and dipyridamole), which likely resulted in hematoma formation. An IL-
IH block may also result in motor block of the quadriceps muscle if the
local anesthetic solution is placed below the inguinal ligament.
Weintraud et al.24 demonstrated that when the IL-IH block is placed
blindly, the local anesthetic is placed in the correct location only 14% of
the time. Willschke et al.25 demonstrated that with ultrasound guidance,
the effective volume can be reduced to 0.075 mL/kg. Fredrickson et al.26
reported improved analgesia with the utility of IL-IH nerve block versus
TAP block after inguinal surgeries in infants and children.
Position: Supine.
Transducer: 25-or 35-mm linear probe oscillating at 13 MHz.
Transducer orientation: Transverse oblique.

Needle: 24G, 1-inch or 22G, 2-inch blunt or insulated needle.


Local anesthetic: 0.1 to 0.15 mL/kg of 0.2% ropivacaine or 0.25%
bupivacaine with epinephrine 1:200,000.
Technique: The skin lateral to the ASIS to the umbilicus is washed, and
sterile gel is applied. The ultrasound probe oscillating at 13 MHz is
placed immediately lateral to and slightly above the ASIS. The plane
between the transverses abdominis and internal oblique muscles is
identified. In larger children, it may be possible to identify the nerves as
identified deep to the internal oblique muscles. The nerves are close
beside each other and often have an “owl’s eyes” appearance. The
needle is inserted in line at the medial or lateral end of the probe and
directed toward the nerves (Figs. 42.21 and 42.22). The local anesthetic
agent is injected around the nerves. If the nerves cannot be readily
identified, local anesthetic is placed between the internal oblique and
transversus abdominis muscles. Given the anatomic variation in the
course of the nerves, it they cannot be identified, it may be beneficial to
inject a second aliquot of local anesthetic between the external and
internal oblique muscles (Fig. 42.23).
Intercostal blockade
Background and Indications: Intercostal nerve block may be useful for
providing analgesia after thoracotomy, upper abdominal procedures, or
rib fractures, or for chest tube pain. For these indications, intercostal
block may be used either in the perioperative arena or in an emergency
room or intensive care unit setting. Intercostal blocks are not effective
for intraperitoneal procedures, as they do not block the celiac plexus.
The intercostal nerves arise paravertebrally from the thoracic spinal
nerves (T1–T12) and may be blocked in their position between the
intercostal muscles in a groove that is found underneath the
corresponding rib and shared with the intercostal vessels. Gray and
white rami communicantes branch off from the spinal nerves before
entering the intercostal space and adjoin the sympathetic ganglia to form
the thoracic sympathetic chain. A second branch, the posterior cutaneous
nerve, travels posteriorly, innervating the paraspinous musculature. This
branch is generally not anesthetized by intercostal block.
Although an intercostal block may be performed at any location along
the lower border of the rib, an approach to the intercostal space at the
posterior axillary line will provide adequate analgesia for thoracotomy,
is relatively simple to perform, and is the most common approach. A
more anterior approach can be used for midline procedures such as open
pectus excavatum repair or sternotomy. The patient can be in either the
supine or sitting position. Given the cross-innervation of the various
interspaces, the block should be performed at the interspace of the
surgical incision or injury and two interspaces above and below.
Complications of intercostal block include pneumothorax, vascular
puncture, and epidural or spinal local anesthetic spread. Spread of local
anesthetic to the epidural or spinal spaces may occur if the injection
travels through a dural sleeve covering the spinal root. This complication
is rare but may be more common in the posterior approach compared to
more anterior approaches. Another complication that must be considered
during intercostal nerve blockade is the risk of local anesthetic toxicity.
Because of the proximity of vessels to the intercostal nerves, there may
be increased risk of local anesthetic toxicity from systemic uptake or
inadvertent vascular puncture compared to other peripheral nerve blocks.
In fact, the peak blood concentration of the local anesthetic agent is
second only to interpleural analgesia when considering the various
regional anesthetic techniques performed in anesthetic practice.
Anterior intercostal block: The nerves to the sternum take origin from
the roots of T1–T6. After exiting the lateral spinous foramina, the nerves
travel in the space between the internal and inner intercostal muscles
along with the intercostal artery and vein below the inferior edge of the
rib. Anterior intercostal block is most useful for analgesia after pectus
excavatum repair.
Patient position: Supine.
Transducer: 25-mm linear array oscillating at 13 MHz.

Needle: 24G, 1-inch or 22G, 2-inch blunt or insulated needle.


Local anesthetic: Total dose of 0.2 to 0.5 mL/kg of 0.2% ropivacaine or
0.25% bupivacaine with epinephrine 1:200,000 divided in the various
interspaces.
Technique: The skin is washed, and sterile gel is applied. A 25-mm
linear probe oscillating at 13 MHz is initially applied in the sagittal
plane several centimeters lateral to the expected site of the pectus repair.
The ribs are imaged as bright hyperechoic structures. Between the ribs,
the pleura appears as hyperechoic streaks approximately 2 to 10 mm
deep to the anterior surface of the rib. The probe is then rotated so that it
is aligned along the rib. Toggling the probe facilitates visualization of
the pleura and intercostal muscles (Figs. 42.24 and 42.25). The needle is
inserted in line at the lateral end of the probe and directed to the plane
between the internal and innermost intercostal muscles at the lower
border of the rib. Once the proper plane has been entered, the local
anesthetic is injected. The block must be repeated at several levels and
on the contralateral side.
Posterior intercostal block: The nerves to the thorax and upper
abdominal wall arise from nerve roots T1–T12. After exiting the lateral
spinous foramina, the nerves travel in the space between the internal and
innermost intercostal muscles along the intercostal artery and vein.
Posterior intercostal block can be used for analgesia after thoracotomy
or sternotomy.
Patient position: Prone, supine, or sitting.
Transducer: 25-mm linear array oscillating at 13 MHz.

Needle: 24G, 1-inch or 22G, 2-inch blunt or insulated needle. An 18G


insulated Tuohy needle can be used if a continuous catheter technique is
planned.
Local anesthetic: Total dose of 0.2 to 0.5 mL/kg of 0.2% ropivacaine or
0.25% bupivacaine with epinephrine 1:200,000 divided in the various
interspaces.
Technique: Depending on the size of the patient and whether he or she is
anesthetized or not, the patient can be placed supine, prone, or in the
sitting position. The block can be placed at the posterior axillary line if
the patient is supine or more centrally if the patient is sitting or prone.
The skin is washed, and sterile ultrasound gel is applied. The ultrasound
transducer is initially oriented in the sagittal plane, and the rib and pleura
are identified. The transducer is then rotated so that it is along the long
axis of the rib. In this orientation, rotating the transducer back and forth
allows the practitioner to identify the rib and pleura. When the patient
breathes, the visceral and parietal pleura can usually be identified as they
slide past each other. In thin patients, the fascia between the intercostal
muscles can also be identified (Figs. 42.26 and 42.27). After sterile
preparation, the needle (length dependent on the age of child) is inserted
through the skin, immediately below the lower border of the rib. The
needle is introduced at the lateral border of the transducer using an in-
plane approach. The needle is then advanced until it lies approximately 2
mm superficial to the pleura. A small amount of local anesthetic is
injected. If a tissue plane appears, this is the plane between the internal
and inner intercostal muscles. Once this plane has been identified, the
total dose of local anesthetic for that interspace can be injected. At the
conclusion of this injection, a catheter may be inserted 2 to 3 cm beyond
the end of the needle if a continuous block is desired. If radiopaque dye
is injected through the catheter and a chest x-ray is obtained, then the
dye usually spreads in a spindle like pattern in the paravertebral space.
Paravertebral block
Background and Indications: Thoracic paravertebral nerve blockade can
provide perioperative analgesia for unilateral procedures (including
thoracotomy) and upper abdominal surgery (including renal surgery).27
The advantages to paravertebral block include the ability to provide
unilateral analgesia without adverse effects, which may be associated
with central neuraxial techniques.28–30 Continuous paravertebral
blockade has been shown to be superior to epidural anesthesia for
unilateral renal surgery and may also be used for analgesia during
inguinal surgery in children.28–32 The paravertebral space is a wedge-
shaped space along the vertebral column that is bound by parietal pleura,
superior costotransverse ligament, and the intercostal membrane. It
contains the intercostal nerve and dorsal ramus, rami communicantes,
and sympathetic chain. There is free communication between adjacent
levels of the paravertebral space, making it possible to achieve multiple
levels of analgesia with the use of a continuous-catheter technique or a
single injection. The exception to the free communication is at the level
of T12 where the insertion of the psoas muscle may inhibit flow to the
lumbar areas of the paravertebral space. This anatomical variant
mandates a two-injection technique for inguinal procedures, one above
and the other below the level of T12. Complications from paravertebral
block in children are rare but may include vascular puncture, pleural
puncture, and pneumothorax. In older children or adults, hypotension
may also occur from sympathectomy.
Overall, with the increasing utility of ultrasound, the placement of
paravertebral blocks has increased for both intraoperative and
postoperative pain relief in the pediatric population. There has been a
shift from a paradigm of block placement as an alternative to neuraxial
anesthesia in patients with comorbid conditions to a primary form of
analgesia with increasing success.33,34
Patient position: Lateral.
Transducer: 25-or 35-mm linear array oscillating at 8 to 13 MHz.
Needle: 22G, 2-inch or 21G, 4-inch blunt or insulated needle. An 18G
insulated Tuohy needle can be used if a continuous catheter technique is
planned.
Local anesthetic: Total dose of 0.3 to 0.5 mL/kg of 0.2% ropivacaine or
0.25% bupivacaine with epinephrine 1:200,000.
Classical technique: The skin is washed, and sterile ultrasound gel is
applied. The transducer is initially placed in the midline along the back,
and the spinous process is identified. It is then moved slightly laterally to
identify the transverse processes. Slight angulation of the tip of the
probe back to the midline allows the identification of the superior
costotransverse ligament and thereby will guide the depth of needle
insertion (Figs. 42.28 and 42.29). The spinous process of the desired
primary dermatome to be blocked is identified. The point of needle
insertion is lateral to midline across from this spinous process, at a
distance that is approximately equal to the distance from spinous process
to spinous process. Alternatively, the approximate lateral distance in mm
from the spinous process to the point of needle insertion, (0.12 × body
weight in kg) + 10.2.33 After sterile preparation and draping, the needle
is inserted and advanced using an LOR technique to saline. With the
needle perpendicular to the skin, the transverse process of the lamina is
contacted and the needle then walked over the cephalad margin of this
process. The approximate depth from the skin to the paravertebral space
in millimeters, (0.48 × body weight in kg) + 18.7. As the transverse
process is passed, there should be a light loss of resistance felt as entry is
gained into the paravertebral space. Local anesthetic may be injected
after negative aspiration for blood or CSF and a catheter threaded if
desired. In infants and children, only 2 to 4 cm of the catheter should be
threaded into the paravertebral space to avoid placing the tip laterally
into an intercostal space, which would result in analgesia of a single
dermatome. If radiopaque dye is injected through the catheter and a
chest x-ray is obtained, the catheter can be seen in the paravertebral
space.
Modified proximal lateral technique: The skin is washed, and sterile
ultrasound gel is applied. The transducer is initially placed in the midline
along the back and the spinous process identified. The transducer is then
moved slightly laterally to identify the transverse processes. The probe is
then turned perpendicular to the longitudinal plane of the spinous
process. The needle is introduced in plane in a medial direction (Figs.
42.30 and 42.31). The LOR technique is used while advancing under
ultrasound guidance to the paravertebral space.
Summary
Truncal and core blocks are effective regional anesthetic techniques to
provide sufficient analgesia in abdominal surgery in children. For
abdominal wall blocks, the superiority of the ultrasonographic guidance
technique over the landmark technique has been proven, however, only
in a limited number of studies. The future of ultrasound-guided pediatric
regional anesthesia is following the adult practice and is becoming more
standard of care for many types of procedures.
References
1. Giaufre E, Dalens B, Gombert A. Epidemiology and morbidity of
regional anesthesia in children: a one-year prospective survey of the
French-language society of pediatric anesthesiologists. Anesth
Analg. 1996;83:904–912.
2. McDonnell JG, O’Donnell BD, Farrell T, et al. Transversus
abdominis plane block: a cadaveric and radiological evaluation.
Reg Anesth Pain Med. 2007;32:399–404.
3. O’Donnell BD, McDonnell JG, McShane AJ. The transversus
abdominis plane (TAP) block in open retropubic prostatectomy.
Reg Anesth Pain Med. 2006;31:91.
4. Jankovic ZB, du Feu FM, McConnell P. An anatomical study of the
transversus abdominis plane block: location of the lumbar triangle
of Petit and adjacent nerves. Anesth Analg. 2009;109:981–985.
5. McDonnell JG, O’Donnell BD, Curley G, et al. Analgesic efficacy
of transversus abdominis plane block (TAP) after abdominal
surgery: a prospective, randomized controlled trial. Anesth Analg.
2007;104;193–197.
6. McDonnell JG, Curley G, Carney J, et al. The analgesic efficacy of
transversus abdominis plane block after Cesarean delivery: a
randomized controlled trial. Anesth Analg. 2008;106:186–191.
7. Carney J, McDonnell JG, Ochana A, et al. The transversus
abdominis plane block provides effective postoperative analgesia in
patients undergoing total abdominal hysterectomy. Anesth Analg.
2008;107:2056–2060.
8. Barrington MJ, Ivanusic JJ, Rozen WM, et al. Spread of injectate
after ultrasound-guided subcostal transversus abdominis plane
block: a cadaveric study. Anaesthesia. 2009;64:745–750.
9. Mukhtar K, Singh S. Transversus abdominis plane block for
laparoscopic surgery. Br J Anaesth. 2008;102:143–144.
10. Mai CL, Young MJ, Quraishi SA. Clinical implications of the
transversus abdominis plane block in pediatric anesthesia. Pediatr
Anesth. 2012;22:831–840.
11. Frederickson M, Seal P, Houghton J. Early experience with the
transverses abdominis plane block in children. Pediatr Anesth.
2008;18:891–892.
12. Hardy CA. Transversus abdominis plane block in neonates: is it a
good alternative to caudal anesthesia for postoperative analgesia
following abdominal surgery? Pediatr Anesth. 2008;19:56.
13. Tobias JD. Preliminary experience with transversus abdominis
plane block for postoperative pain relief in infants and children.
Saudi J Anesth. 2009;1:2–6.
14. Smith BE, Suchak M, Siggins D, et al. Rectus sheath block for
diagnostic laparoscopy. Anaesthesia. 1988;43:947–948.
15. Ferguson S, Thomas V, Lewis I. The rectus sheath block in
paediatric anaesthesia: new indications for an old technique.
Paediatr Anaesth. 1996;6: 463–466.
16. Courreges P, Poddevin F, Lecoutre D. Paraumbilical block: a new
concept for regional anaesthesia in children. Paediatr Anaesth.
1997;7:211–214.
17. Willschke H, Bosenberg A, Marhofer P, et al. Ultrasonography-
guided rectus sheath block in paediatric anaesthesia—a new
approach to an old technique. Br J Anaesth. 2006;97:244–249.
18. De Jose Maria B, Gotzens V, Marbrok M. Ultrasound-guided
umbilical nerve block in children: a brief description of a new
approach. Pediatr Anesth. 2007;17:44–50.
19. Gurnaney HG, Maxwell LG, Kraemer FW, et al. Prospective
randomized observer-blinded study comparing the analgesic
efficacy of ultrasound-guided rectus sheath block and local
anaesthetic infiltration for umbilical hernia repair. Br J Anaesth.
2011;107:790–795.
20. Hannallah RS, Broadman LM, Belman AB, et al. Comparison of
caudal and ilioinguinal/iliohypogastric nerve blocks for control of
post-orchiopexy pain in pediatric ambulatory surgery.
Anesthesiology. 1987;66:832–834.
21. Casey WF, Rice LJ, Hannallah RS, et al. A comparison between
bupivacaine installation versus ilioinguinal/iliohypogastric nerve
block for postoperative analgesia following inguinal herniorrhaphy
in children. Anesthesiology. 1990;72:637–639.
22. Fisher QA, McComiskey CM, Hill JL, et al. Postoperative voiding
interval and duration of analgesia following peripheral or caudal
nerve blocks in children. Anesth Analg. 1993;76:173–177.
23. Parvaiz MA, Korwar V, McArthur D, et al. Large retroperitoneal
haematoma: an unexpected complication of ilioinguinal nerve block
for inguinal hernia repair. Anaesthesia. 2012;67:73–84.
24. Weintraud M, Marhofer P, Boenberg A, et al.
Ilioinguinal/iliohypogastric blocks in children: where do we
administer the local anesthetic without direct visualization? Anesth
Analg. 2008;106:89–93.
25. Willschke H, Bosenberg A, Marhofer P, et al. Ultrasonographic-
guided ilioinguinal/iliohypogastric nerve block in pediatric
anesthesia: what is the optimal volume? Anesth Analg.
2006;102:1680–1684.
26. Fredrickson M, Paine C, Hamill J. Improved analgesia with the
ilioinguinal block compared to the transversus abdominis plane
block after pediatric inguinal surgery: a prospective randomized
trial. Pediatr Anesth. 2010;20:1022–1027.
27. Lonnqvist PA. Continuous paravertebral block in children: initial
experience. Anaesthesia. 1992;47:607–611.
28. Lonnqvist PA, MacKenzie J, Soni AK, et al. Paravertebral
blockade. Failure rate and complications. Anaesthesia.
1995;50:813–815.
29. Lonnqvist PA, Olsson GL. Paravertebral vs. epidural block in
children: effects on postoperative morphine requirements after renal
surgery. Acta Anesthesiol Scand. 1994;38:346–349.
30. Naja ZM, Raf M, El Rajab M, et al. Nerve stimulator-guided
paravertebral blockade combined with sevoflurane sedation versus
general anesthesia with systemic analgesia for post-herniorrhaphy
pain relief in children. Anesthesiology. 2005;103:600–605.
31. Karmakar M. Thoracic paravertebral block. Anesthesiology.
2001;95:771–780.
32. Naja ZM, Raf M, Rajab M, et al. A comparison of nerve stimulator
guided paravertebral block and ilioinguinal nerve block for
analgesia after inguinal herniorrphaphy in children. Anaesthesia.
2006;61:1064–1068.
33. Lonnqvist PA, Hesser U. Location of the paravertebral space in
children and adolescents in relation to surface anatomy assessed by
computed tomography. Paediatr Anaesth. 1992;2:285–289.
34. Bhalla T, Sawardekar A, Dewhirst E, et al. Ultrasound-guided trunk
and core blocks in infants and children. J Anesth. 2013;27:109–123.
Section VI
Pain Blocks
43 Ultrasound-Guided Maxillary and Mandibular
Block
Part 1: Maxillary Nerve Block
Part 2: Mandibular Nerve Block

44 Ultrasound-Guided Stellate Ganglion Block

45 Ultrasound-Guided Cervical Sympathetic Block

46 Celiac Ganglion Block: Endoscopic and


Transabdominal Approaches

47 Ultrasound-Guided Superior Hypogastric Plexus


Block

48 Ultrasound-Guided Genitofemoral Nerve Block

49 Ultrasound-Guided Pudendal Nerve Block

50 Neuraxial Anatomy Relevant to the Two-


Dimensional Ultrasound Examination

51 Ultrasound-Guided Third Occipital Nerve Block


and Cervical Medial Branch Block

52 Ultrasound-Guided Cervical Nerve Root Injections

53 Ultrasound-Guided Lumbar Transforaminal


Epidural Steroid Injection
54 Ultrasound-Guided Lumbar Facet Medial Branch
Block and Intra-articular Facet Joint Injection

55 Ultrasound-Guided Botulinum for Spasticity and


Applications in Physiatry

56 Ultrasound-Guided Sacroiliac Joint Injection

57 Ultrasound-Guided Piriformis Block

58 Ultrasound-Guided Subacromial Bursa Injection

59 Ultrasound-Guided Bicipital Tendon Sheath


Injection

60 Ultrasound-Guided Glenohumeral Joint Injection

61 Ultrasound-Guided Supraspinatus Tendon Injection

62 Ultrasound-Guided Acromioclavicular Joint


Injection

63 Ultrasound-Guided Sternoclavicular Joint Injection

64 Ultrasound-Guided Intra-articular Elbow Injection

65 Ultrasound-Guided Wrist Injections

66 Ultrasound-Guided Ulnar Triquetral Injection

67 Ultrasound-Guided Carpal Tunnel Injection

68 Ultrasound-Guided Intra-articular Hip Injection

69 Ultrasound-Guided Patellar Tendon Injection

70 Ultrasound-Guided Intra-articular Ankle Injection

71 Ultrasound-Guided Tarsal Tunnel Injection


72 Ultrasound-Guided Peripheral Nerve Stimulation

73 Ultrasound-Guided Intrathecal Pump Management


43
Ultrasound-Guided
Maxillary and
Mandibular Block
Part 1: Maxillary Nerve Block

PAUL E. BIGELEISEN AND MILENA MORENO

Background and indications: Maxillary nerve block is used primarily


for the diagnosis of trigeminal neuralgia and for postoperative pain relief
in children having cleft palate repairs. The block can also be used as the
sole block for facial surgery in patients who are poor candidates for
general surgery.

Anatomy: The maxillary nerve is the second division (V2) of cranial


nerve V (trigeminal nerve). The nerve exits the base of the skull through
the foramen rotundum and travels into the pterygopalatine fossa (Fig.
43.1). The nerve is anterior and deep to the lateral pterygoid plate. The
nerve supplies sensory innervation to the skin of the face, including the
cheeks, upper lip, and nasolabial folds, as well as sensory innervation to
the maxillary sinuses, upper teeth, and the hard and soft palate.
Patient position: Supine with the head turned away from the operator.

Transducer: 25-or 38-mm linear probe oscillating at 10 to 13 MHz.

Transducer orientation: Transverse oblique below the zygoma.

Needle: 22G, 5-cm needle with a blunt tip.

Local anesthetic: Ropivacaine 0.2%, bupivacaine 0.25%.

Technique: The patient is positioned supine with the head turned away
from the practitioner. The patient is sedated, and the skin is washed. The
transducer is sited below the zygoma, and the lateral pterygoid plate is
imaged (Fig. 43.2). Anterior and deep to the pterygoid plate, the nerve is
seen as a round or triangular hyperechoic structure. The needle is placed
at the posterior end of the probe and advanced through the skin toward
the nerve. Once the needle is adjacent to the nerve, 2 to 5 mL are
injected until the nerve is surrounded with local anesthetic. The patient
should be monitored closely for 30 minutes after the injection because
spread of local anesthetic to the brain stem or other cranial nerves is a
possibility. For small children, the dose should be reduced and the nerve
must be blocked bilaterally for palate repair.

Part 2: Mandibular Nerve Block

PAUL E. BIGELEISEN AND MILENA MORENO

Background and indications: Mandibular nerve block was traditionally


done using a blind approach in which the practitioner inserted a needle
inferior to the zygoma until the pterygoid plate was contacted. The
needle was then withdrawn and advanced posterior and 1 cm deep to the
pterygoid plate. Some practitioners sought a paresthesia in the lower jaw
as they inserted the needle. Once the needle was assumed to be in the
correct position, 5 mL of local anesthetic was injected. Another
technique was to insert the needle inferior to the zygoma until the
pterygoid plate was contacted. Once bony contact was made, 10 mL of
local anesthetic was injected. This approach had a high success rate. In
many cases, the local anesthetic spread beyond the intended distribution
to anesthetize the maxillary nerve and other cranial nerves, including
cranial nerves VI and VII. The use of ultrasound to facilitate mandibular
nerve block has not been studied, but it has the potential to increase
success rate and decrease the incidence of intravascular injection.
The block is usually employed as a diagnostic or therapeutic procedure
for trigeminal neuralgia. The block may also be used for dental
procedures or repair of mandibular fractures.

Anatomy: The mandibular nerve is the third division (V3) of cranial


nerve V (trigeminal nerve) (Fig. 43.3). It exits from the base of the skull
through the foramen ovale and runs parallel to the posterior margin of
the lateral pterygoid plate, where it divides into two branches. The
anterior division is primarily motor and supplies the muscles of
mastication. The posterior division is sensory and supplies the skin over
the lower jaw, cheek, anterior ear, and the skin above the ear. It also
innervates mucous membranes overlying the lower jaw, the cheek, the
mandible, and the teeth. The lingual branch of the posterior division
joins the chorda tympani and supplies taste to the anterior two-thirds of
the tongue.
Patient position: Supine with the head turned away from the operator.

Transducer: 25-mm linear probe oscillating at 13 MHz; 11-mm curved


probe oscillating at 10 MHz.

Transducer position: Transverse inferior to the zygoma and anterior to


the ramus of the mandible.

Needle: 22G, 5-cm needle with a blunt tip.

Local anesthetic: Ropivacaine 0.2%, bupivacaine 0.25%.

Technique: The skin is washed. If a linear probe is chosen, the probe is


sited superior to the mandible in a transverse orientation (Fig. 43.3A).
The condyle of the mandible is imaged. Anterior to the condyle, the
alveolar artery and vein may be seen next to the nerve (Fig. 43.3B,C).
The needle is inserted superior to the probe using an out-of-plane
technique. The needle is advanced toward the nerve, which is ovoid and
hyperechoic. Local anesthetic (3 mL) is injected until the nerve is
surrounded. If a curved probe is chosen, the probe is placed in an
oblique transverse position below the zygoma and anterior to the
mandibular condyle (Fig. 43.4A). The lateral pterygoid plate is imaged
(Fig. 43.4B). Posterior to the pterygoid plate, the maxillary artery, or its
branch, the alveolar artery is imaged. Deep to the artery, and posterior to
the pterygoid plate, the nerve is seen as a hyperechoic round or oval
structure. The needle is inserted posterior to the probe using an in-line
technique. The needle is advanced until it is adjacent to the nerve. Local
anesthetic (3 mL) is injected until the nerve is surrounded.
44
Ultrasound-Guided
Stellate Ganglion
Block
YASUYUKI SHIBATA, TORU KOMATSU, NIZAR

MOAYERI, AND GERBRAND J. GROEN

Background and indications: Ultrasound-guided stellate ganglion block


was first introduced in 1995 with promising results.1 The main
advantages of ultrasound over the traditional technique are direct
visualization of the stellate ganglion, adjacent structures, and
visualization of the spread of the local anesthetic. Although more
clinical reports are needed, the use of ultrasound in stellate ganglion
block is reported to reduce the risk of vascular puncture, reduce the
amount of the injected anesthetic solution, and provide a faster onset
time.1 An important prerequisite is good clinical experience of the
anesthesiologist in the practice of ultrasound.
Stellate ganglion block is usually used for the diagnosis and treatment of
complex regional pain syndrome of the upper extremity.2,3 Other uses
include treatment of acute or postherpetic neuralgia of the face4–6 and
neck and poor perfusion of the upper extremity, as seen with Raynaud
and other vascular diseases.7,8

Anatomy: The cervical sympathetic ganglia consist of the superior,


middle, and inferior ganglia. The inferior cervical ganglion is located at
the level of C7 and C8. It joins with the first thoracic ganglion to form
the stellate ganglion. The cervical ganglia lie deep to the prevertebral
fascia. The prevertebral fascia overlies the vertebral body, the longus
colli muscle, and the anterior scalene muscle (Figs. 44.1 and 44.2). The
three cervical ganglia give off grey rami communicantes that travel
through and around the longus colli to join with spinal nerves C1–C8
and T1. The carotid artery lies anterior and lateral to the cervical
ganglia. The recurrent laryngeal nerve is sandwiched between the
thyroid and trachea. At the level of C6, the transverse process has a bony
process (Chassaignac tubercle), which may be palpated with deep
pressure by retracting the carotic artery laterally. This has traditionally
been the landmark used to perform stellate block before ultrasound was
available.
No major changes have been suggested on the initial reported procedure
for ultrasound-guided stellate ganglion block. The use of curved instead
of linear transducers is advocated because linear transducers are
generally too large for this procedure. In addition, similar high success
rates are reported after injecting the local anesthetic around the fascia1
compared to injecting the local anesthetic under the fascia of the longus
colli muscle.9

Patient position: Supine, with a towel between the shoulders.

Transducer: Small curved array (11 mm) oscillating at 5 to 10 MHz.

Transducer position: Transverse position between the carotid artery and


trachea at the level of the sixth cervical vertebra.

Needle: 22G, 5-cm blunt needle.

Local anesthetic: 0.2% ropivacaine or 0.25% bupivacaine (3 to 10 mL).


Procedure: The patient is placed supine with a towel or small pillow
placed behind the shoulders. This allows the practitioner to extend the
neck and jaw. The patient is asked to open his or her mouth slightly,
which makes it easier for the patient to relax the neck muscles during the
procedure. After aseptic preparation of the skin, an 11-mm curved
transducer oscillating at 8 to 10 MHz is placed adjacent to the trachea to
identify the carotid artery and transverse process of C6 (Figs. 44.3 to
44.5). In most patients, the practitioner will need to apply some pressure
with the transducer between the carotid artery and trachea to retract the
carotid artery laterally. Once the transverse process is identified, the
longus colli muscle covered with the prevertebral fascia can usually be
seen immediately anterior to the transverse process. A 5-cm, 22G short-
bevel needle is placed between the trachea and transducer and advanced
in line until the tip of the needle is anterior to the longus colli muscle.
After aspiration, 5 to 10 mL of local anesthetic is injected until the
longus colli muscle is surrounded with local anesthetic (Fig. 44.5). In
obese patients, it may be difficult to identify the longus colli muscle. In
this case, the practitioner should advance the needle until it contacts the
transverse process. The needle should then be withdrawn 1 to 2 mm, and
the injection should proceed under ultrasound guidance. The triad of
anhydrosis, meiosis of the pupil, and ptosis of the upper lid, usually
within 5 min, are evidence of successful block.
Tips 1. Because the spinal nerve travels in the gutter of the transverse
processes, paresthesiae may occur when the needle is advanced toward
the transverse process of C6. When this occurs, the needle should be
withdrawn 3 to 4 mm and then the injection may proceed.
2. The brachial plexus lies close by. Some of the local anesthetic may
spill over onto the nerve roots or trunks, causing a partial plexus
block.
3. The recurrent nerve also lies nearby, and ipsilateral vocal cord
paralysis may also occur. This is thought to occur when local
anesthetic is injected or spreads superficial to the prevertebral fascia
and reaches the space between the trachea and esophagus where the
recurrent laryngeal nerve resides.
4. In rare instances, epidural block has been reported as well as spread
beneath the prevertebral fascia producing bilateral stellate block.
References
1. Kapral S, Krafft P, Gosch M, et al. Ultrasound imaging for stellate
ganglion block: direct visualization of puncture site and local
anesthetic spread. A pilot study. Reg Anesth. 1995;20:323–328.
2. Ackerman WE, Zhang JM. Efficacy of stellate ganglion blockade
for the management of type 1 complex regional pain syndrome.
South Med J. 2006;99:1084–1088.
3. Chaturvedi A, Dash HH. Sympathetic blockade for the relief of
chronic pain. J Indian Med Assoc. 2001;99:698–703.
4. Colding A. The effect of regional sympathetic blocks in the
treatment of herpes zoster. Acta Anesth Scand. 1969;13:133–141.
5. Olson ER, Ivy HB. Stellate block for trigeminal zoster. J Clin
Neuroophthalmol. 1981;1:53–55.
6. Winnie AP, Hartwell PW. Relationship between time of treatment of
acute herpes zoster with sympathetic blockade and prevention of
postherpetic neuralgia: clinical support for a new theory of the
mechanism by which sympathetic blockade provides therapeutic
benefit. Reg Anesth. 1993;18:277–282.
7. Yildirim V, Akay HT, Bingol H, et al. Pre-emptive stellate ganglion
block increases the patency of radial artery grafts in coronary artery
bypass surgery. Acta Anaesthesiol Scand. 2007;51:434–440.
8. Gupta MM, Bithal PK, Dash HH, et al. Effects of stellate ganglion
block on cerebral haemodynamics as assessed by transcranial
Doppler ultrasonography. Br J Anaesth. 2005;95:669–673.
9. Shibata Y, Fujiwara Y, Komatsu T. A new approach of ultrasound-
guided stellate ganglion block. Anesth Analg. 2007;105:550–551.
45
Ultrasound-Guided
Cervical Sympathetic
Block

MICHAEL GOFELD

Background and indications: Stellate ganglion block technique was first


described in the 1930s. First implemented by Leriche, and further
refined by Findley and Patzer, the method has been practiced without
major modifications since that time.1 It is a common intervention in the
diagnosis and management of sympathetically maintained pain and
vascular insufficiency of upper extremities. In addition, the block has
been advocated in a variety of medical conditions, such as phantom pain,
postherpetic neuralgia, cancer pain, cardiac arrhythmias, orofacial pain,
and vascular headache.2
The anatomy and position of the stellate ganglion has been confirmed by
dissection, magnetic resonance, and computed tomography.3–8 The
stellate ganglion, or cervicothoracic ganglion, is described as a structure
1 to 2.5 cm in length, approximately 1 cm in width, and 0.5 cm thick. It
is present in about 80% of the population as a fusion of the inferior
cervical ganglion and the first thoracic ganglion. The shape of the
ganglion may be fusiform, triangular, or globular.6 The ganglion sits just
anterior to the transverse process of C7 and superior or anterior to the
neck of the first rib.
An inferior C7 approach for stellate ganglion block has been described.9
The more common approach is administered according to anatomical
landmarks: the prominent anterior tubercle of the C6 vertebra
(Chassaignac tubercle), the cricoid cartilage, and the carotid artery.4 At
the level of C6, only the traversing sympathetic fibers or middle cervical
ganglion can be found.10 Thus, the procedure at this level should be
named the cervical sympathetic block. Ultrasound-guided stellate
ganglion blockade was described in 19954 but has recently gained
popularity. A new lateral approach to the cervical sympathetic trunk has
been recently described and validated against fluoroscopy.11,12

Anatomy: The cervical sympathetic trunk is situated on the lateral


surface of the longus colli muscle beneath the deep cervical fascia. The
cervical sympathetic trunk lies medial to the anterior scalene muscle;
lateral to the longus colli muscle, esophagus, trachea, and recurrent
laryngeal nerve; and anterior to the transverse process of C6. A classic
anterior neck sonogram at the level of the cricoid cartilage reveals the
trachea medially, the carotid artery laterally, and the thyroid gland
between those two structures. Immediately medial to the carotid artery,
the inferior thyroid artery or the recurrent laryngeal nerve can be located
(Fig. 45.1). The longus colli muscle is usually seen posterior to the
thyroid gland on the anterolateral surface of the C6 vertebral body. On
the left side of the neck, the esophagus is positioned just anterior to the
longus colli muscle. Thus, insertion of a block needle in the usual
trajectory would result in traversing the thyroid gland and esophagus.
Although displacement of the carotid artery and thyroid by a pediatric-
type curved transducer creates a passage for needle placement (Fig.
45.2), this approach requires application of pressure that may be very
unpleasant and does not eliminate the chance for vascular and visceral
damage.
The lateral approach has the advantage of in-plane needle placement and
sparing of visceral and vascular structures from damage. A lateral axial
scan usually reveals the C6 transverse process with the prominent
anterior and shorter posterior tubercles as well as the exiting C6 nerve
root (Fig. 45.3). At the level of C6, the longus colli muscle is seen as an
oval structure adjacent to the base of the transverse process and vertebral
body. Sometimes, a caudal part of the longus capitis muscle can be seen
as well. In 30% of patients, the middle cervical ganglion can be
visualized as a spindle-shaped structure situated on the posterolateral
surface of the longus colli muscle.10 Otherwise, a widening of the tissue
plane below the deep cervical fascia is appreciated (Fig. 45.4). Scanning
caudally and slightly dorsally brings the C7 transverse process into the
view. The C7 transverse process has only one tubercle, and its nerve root
is situated anterior to the transverse process (Fig. 45.4).
Patient position Anterior Approach: Supine with a towel beneath the
shoulders, neck extended, and mouth open.

Lateral Approach: Lateral decubitus position with the operative side up.

Transducer Anterior Approach: Small curved transducer (C11)


oscillating at 5 to 10 MHz.
Lateral Approach: Small curved transducer (C11) or linear transducer
(L25) oscillating at 10 to 13 MHz Transducer orientation
Anterior Approach: The transducer is placed at the level of the cricoid
cartilage medially to the carotid artery in the axial plane (Fig. 45.5).

Lateral Approach: The probe is positioned at the level of the cricoid


cartilage above the sternocleidomastoid muscle in the axial plane (Fig.
45.5). Usually, the transducer is moved slightly laterally or medially,
cephalad, or caudad to provide the optimal image of the C6 transverse
process and the longus colli muscle.

Needle: 22G, 5-to 10-cm, blunt needle.

Local anesthetic: 0.2% ropivacaine or 0.25% bupivacaine (3 to 10 mL),


normal saline for test injection.

Other equipment 1. 3-mL, 5-mL, and 10-mL lock tip syringes,


three-way stopcock, extension tubing, 25-g hypodermic needle.

Technique: Cardiovascular monitors are placed and an intravenous


catheter is inserted on the contralateral side of the block. A temperature
monitor can be attached to the palm on the block side. Sedation is
usually not required. An antiseptic is utilized to prepare the skin of the
block area. The ultrasound transducer is covered by a sterile transparent
adhesive or sleeve and positioned at the level of the cricoid cartilage.
The block needle is attached to the connecting tubing and to two
syringes connected via a three-way stopcock. A 5-mL syringe is filled
with NaCl (0.9%), and a 10-mL syringe is filled with local anesthetic.
The tubing and needle are flushed with NaCl (0.9%).
When the anterior approach is chosen, the patient is positioned supine
with the neck slightly extended. The pediatric curved transducer is
placed medial to the carotid artery, and gentle pressure is applied to
separate the carotid artery and the thyroid gland. This technique also
decreases the distance from the skin to the target. The longus colli
muscle is seen as a flattened muscular structure anterior to the C6
vertebral bone shadow. The color Doppler mode should be applied at
this time to identify the inferior thyroid and vertebral arteries. Skin
anesthesia is performed at the point inferior (caudad) to the transducer
and the block needle is advanced using an out-of-plane technique aiming
at the fascial plane of the longus colli muscle. When the needle tip is
seen immediately beneath the deep cervical fascia, 0.5 to 1 mL of NaCl
(0.9%) is injected. The spread of hypoechoic solution in this fascial
plane typically creates a double hyperechoic line with the deep cervical
fascia superior and the colli muscle fascia inferior (Fig. 45.5). If no
spread is observed, the needle tip is positioned outside of the ultrasound
beam or within a vessel and the needle should be repositioned. Needle
repositioning is also required if an intramuscular injection is seen. In this
case, only one hyperechoic line is visualized. Moreover, intramuscular
injection is usually associated with discomfort and a retropharyngeal
sensation of pressure. Once the spread is seen within the fascial plane,
the analgesic solution is administered. Usually, an injection of 5 mL will
result in spread from C4 to T1, reliably blocking the midneck
sympathetic trunk and the stellate ganglion.12
If the lateral approach is performed, the patient is placed in the lateral
decubitus position, with the block side up. A high-frequency linear
transducer is positioned in the axial orientation. The C6 anterior tubercle
is brought into view. If the nerve root is also seen, it may be damaged by
an in-plane approach (Fig. 45.6). Skin anesthesia is performed posterior
to the probe, and the block needle is advanced under continuous in-plane
viewing. The needle traverses the sternocleidomastoid and occasionally
the anterior scalene and the longus capitis muscles. The needle is
advanced beneath the deep cervical fascia and lateral to the longus colli
muscle. Verification of the correct spread of NaCl solution is performed
prior to injection of local anesthetic solution (Fig. 45.7).
After the block, the patient should remain in the head-up supine position.
Usually, the first sign of a successful sympathetic blockade is Horner
syndrome, which can be seen several minutes following the injection.
The patient should be closely monitored for adverse effects.
Summary of evidence: Traditionally, the block was done as a blind
injection and was associated with complications, such as intravascular
injection, hematoma formation, temporary paralysis of the recurrent
laryngeal nerve, and esophageal injury.2,13,14 Ultrasound guidance can
effectively reduce the volume of local anesthetic to achieve reliable
block and may also prevent puncture of blood vessels and
esophagus.15,16 Ultrasound guidance has advantages compared to
fluoroscopic guidance. Fluoroscopic guidance relies on identification of
a bony landmark (transverse process) and a contrast dye injection. A
successful fluoroscopic block also relies on the assumed thickness and
anatomical location of the longus colli muscle. There is some
controversy about the location of the sympathetic trunk location with
respect to the deep cervical fascia. The majority of anatomy atlases and
regional anesthesia texts cite its position as superficial to the deep
cervical fascia.3 The subfascial position of the cervical sympathetic
trunk has been confirmed by both anatomical dissection10,12 and
ultrasound imaging.12 In addition, these texts imply that the longus colli
muscle is very thin and advocate withdrawing the needle 2 to 5 mm once
the transverse process has been identified and contacted by the needle
using fluoroscopy. However, the longus colli muscle is actually about 10
mm in thickness.12 Thus, a standard fluoroscopic approach would likely
result in an intramuscular injection. Block of the cervical sympathetic
trunk would then only be accomplished by overflow or diffusion of the
injectate. Because the cervical sympathetic trunk is adjacent to soft
tissues (longus colli muscle, thyroid, esophagus), ultrasound guidance
should logically be a better technique.

References 1. Bonica JJ. The Management of Pain. Philadelphia, PA:


Lea and Febiger;1953: 410–432.
2. Elias M. Cervical sympathetic and stellate ganglion blocks. Pain
Physician. 2000;3:294–304.
3. Moore K, Dalley A. Clinically Oriented Anatomy. 5th ed.
Philadelphia, PA: Lippincott Williams & Wilkins;2006:1051, 1082.
4. Raj PP. Stellate ganglion block. In: Waldman SD, Winnie AP, eds.
Interventional Pain Management. Philadelphia, PA: Saunders;
1996.
5. Ellis H, Feldman S. Anatomy for Anesthetists. 3rd ed. Oxford:
Blackwell Scientific Publications;1979:256–262.
6. Hogan Q, Erickson SJ. Magnetic resonance imaging of the stellate
ganglion: normal appearance. Am J Roentgenol. 1992;158:655–659.
7. Perlow S, Vehe KL. Variations in the gross anatomy of the stellate
and lumbar sympathetic ganglia. Am J Surg. 1935;30:454–458.
8. Erickson SJ, Hogan QH. CT-guided injection of the stellate
ganglion: of technique and efficacy of sympathetic blockade.
Radiology. 1993;188:707–709.
9. Abdi S, Zhou Y, Patel N, et al. A new and easy technique to block
the stellate ganglion. Pain Physician. 2004;7:327–331.
10. Kiray A, Arman C, Naderi S, et al. Surgical anatomy of the cervical
sympathetic trunk. Clin Anat. 2005;18:179–185.
11. Gofeld M. Ultrasonography in pain medicine: a critical review. Pain
Practice. 2008;8:226–240.
12. Gofeld M, Bhatia A, Abbas S. Three-dimensional ultrasonography
of the cervical sympathetic trunk and validation of a new
ultrasound-guided technique. Reg Anesth Pain Med. 2008;32:2.
13. Higa K, Hirata K, Hirota K, et al. Retropharyngeal hematoma after
stellate ganglion block: Analysis of 27 patients reported in the
literature. Anesthesiology. 2006;105:1238–1245.
14. Mahli A, Coskun D, Akcali DT. Aetiology of convulsions due to
stellate ganglion block: a review and report of two cases. Eur J
Anaesthesiol. 2002;19:376–380.
15. Kapral S, Krafft P, Gosch M, et al. Ultrasound imaging for stellate
ganglion block: direct visualization of puncture site and local
anesthetic spread. A pilot study. Reg Anesth. 1995;20:323–328.
16. Narouze S, Vydyanathan A, Patel N. Ultrasound-guided stellate
ganglion block successfully prevented esophageal puncture. Pain
Physician. 2007;10:747–752.
46
Celiac Ganglion
Block: Endoscopic
and Transabdominal
Approaches
PAUL E. BIGELEISEN

Background and indications: Block of the celiac ganglion is usually


used for the diagnosis of visceral pain syndromes or the treatment of
pain from visceral malignancies.
The traditional approach to celiac ganglion block used a bilateral
posterior approach in which the needle was placed anterior to the aorta
using a transcrural approach. This was done either blind or with
fluoroscopy. Subsequently, practitioners used a transaortic approach,
again with a blind, fluoroscopic, or computed tomography (CT)–guided
technique. Other practitioners have used a transabdominal approach
using fluoroscopy, CT, or ultrasound guidance.
The advent of endoscopic ultrasound has made the celiac block and
neurolysis simple and very safe. An endoscopic ultrasound probe
consists of two imaging modalities placed into one probe. The first part
of the probe is a flexible endoscope, which is identical to any flexible
endoscope used for upper gastroenterological procedures. The second
part of the probe is an ultrasound imaging device, which functions in the
same way as a transesophageal echocardiogram (see Fig. 46.2) Thus, the
practitioner must be familiar with the basics of gastrointestinal
endoscopy as well as transesophageal echocardiography in order to
employ the technique. Many university gastroenterology departments
have endoscopic ultrasound machines, but the technology has not
diffused into the private community at this time.

Anatomy: The celiac ganglion is predominantly a sympathetic plexus


with afferent pain fibers from the foregut. Foregut organs from the distal
esophagus to the splenic flexure are innervated by afferents traveling
through the celiac ganglion (Fig. 46.1). Efferent fibers from the vagus
also traverse the ganglion. The ganglion surrounds the celiac artery and
spreads over the anterior surface of the aorta at the level of the 12th
thoracic vertebra and the first lumbar vertebra.

Transducer Endoscopic: 120-cm endoscopic ultrasound probe oscillating


at 2 to 5 MHz.
Transabdominal: 40-to 60-mm curved array oscillating at 2 to 5 MHz.

Transducer orientation Endoscopic: The probe is passed through the


mouth and placed alongside the lesser curvature of the stomach.

Transabdominal: Transverse, 2 cm above the umbilicus.

Needle Endoscopic: 22G flexible needle (150 cm in length).

Transabdominal: 20G, 10-to 15-cm blunt needle.

Local anesthetic: 0.2% to 0.5% ropivacaine (10 to 20 mL).

Technique Endoscopic: After appropriate sedation and topicalization of


the hypopharynx, the endoscopic ultrasound probe is passed orally into
the stomach of the patient and then placed alongside the lesser curvature
of the stomach under endoscopic guidance (Fig. 46.2). The ultrasound
part of the probe is then used to visualize the aorta and celiac artery (Fig.
46.3). Once the celiac artery is identified, a flexible, 22G needle is
passed through the wall of the stomach and the tip of the needle is
placed adjacent to the celiac artery. Local anesthetic (10 mL) is then
injected through the needle around the celiac artery under direct
ultrasound visualization.
Transabdominal: For practitioners who do not have access to an
endoscopic ultrasound device, celiac plexus block can be performed
from the anterior approach using a traditional ultrasound platform.
Because the needle traverses the peritoneum and small bowel, the patient
should receive broad-spectrum antibiotic prophylaxis prior to the block.
The author’s preference is 2 g of cefoxitin. The abdomen is washed with
sterile cleanser and a 60-cm curved array oscillating at 2 to 5 MHz is
placed 2 cm cephalad to the umbilicus (Fig. 46.4). The pulsatile aorta is
identified and then the practitioner must scan cephalad until the celiac
artery is identified (Fig. 46.5). Once the celiac artery is identified, the
skin and abdominal wall are anesthetized with local anesthetic. A 10-cm,
20G needle is advanced through the abdominal wall in line with the
probe. When the needle tip is adjacent to the celiac artery, 10 mL of
local anesthetic is injected under ultrasound visualization. If the celiac
artery cannot be identified, the practitioner may inject the local
anesthetic immediately anterior to the aorta. In this case a volume of 30
mL of local anesthetic is required.
Bibliography
1. Wiersema MJ, Wiersema LM. Endosonography-guided celiac
plexus neurolysis. Gastrointest Endosc. 1996;44:656–662.
2. Hoffman BJ. EUS-guided celiac plexus neurolysis. Gastrointest
Endosc. 2002;56:S26–S28.
3. Caratozzolo M, Lirici MM, Consalvo M, et al. Ultrasound guided
alcoholization of the celiac plexus for pain control in oncology.
Surg Endosc. 2004;11:239–244.
47
Ultrasound-Guided
Superior Hypogastric
Plexus Block
PAUL E. BIGELEISEN

Background and indications: Traditionally, hypogastric plexus block


was performed by anesthesiologists from a posterior approach using
fluoroscopy or computed tomography (CT) guidance. This procedure
was technically difficult because it required the practitioner to advance
the needle between the wing of the posterior iliac crest and vertebral
body into a position anterior to the vertebra at the L5–S1 junction.
Radiologists often used an anterior approach with fluoroscopy or CT
guidance in which the needle was advanced through the abdominal wall
until it was immediately anterior to the anterior longitudinal ligament of
the vertebrae. This same approach is easy to perform with ultrasound
and has the advantage over fluoroscopy that the aorta and iliac arteries
are readily imaged. Because this approach requires the needle to traverse
the small bowel and peritoneum, the patient should be treated with
broad-spectrum antibiotic prophylaxis prior to the procedure. The
author’s preference is 2 g of cefoxitin.
Block of the superior hypogastric plexus is used as a diagnostic
procedure or therapeutic treatment of pain syndromes related to the
testes, ovaries, uterus, cervix, vagina, bladder, sigmoid colon, and
rectum. Ultrasound guidance can be used for the anterior approach to the
superior hypogastric plexus. When neurolytic blocks are intended, the
practitioner must be aware that bowel and bladder incontinence or
dysfunction may ensue. For this reason, neurolytic blocks are usually
reserved for patients with terminal cancer of the above-named organs.

Anatomy: The superior hypogastric plexus contains contributions from


the intermesenteric plexus, which descends over the aortic bifurcation. It
also receives branches from the two lower lumbar splanchnic nerves.
The plexus carries visceral efferents to and somatic afferents from the
testes, ovaries, uterus, cervix, vagina, bladder, sigmoid colon, and
rectum. The plexus lies anterior to the fourth and fifth lumbar and the
sacrum (Fig. 47.1). Below the junction of the lumbar vertebrae and
sacrum, the plexus bifurcates into the right and left hypogastric nerves.

Transducer: 40-to 60-mm curved array oscillating at 2 to 5 MHz.

Transducer orientation: Transverse below the umbilicus.

Needle: 20G needle. In slim patients, a 10-cm needle will suffice. In


larger patients, a needle 15 to 20 cm in length will be required.
Local anesthetic: 10 to 20 mL of 0.2% to 0.5% ropivacaine.

Technique: The patient’s abdomen is washed with sterile solution. The


probe is placed about 2 cm cephalad to the umbilicus in a transverse
orientation (Fig. 47.2A). The spine is seen as a white, hyperechoic
curved line (Fig. 47.2B). The pulsatile aorta is identified anterior to the
spine. In many cases, the celiac artery can also be seen immediately
anterior to the aorta. The inferior vena cava is seen to the right of the
aorta. The probe is moved caudad until the practitioner views the aorta
bifurcating into the right and left common iliac arteries (Fig. 47.3A,B).
The spine may now be seen lying between the iliac arteries. The needle
is introduced through the skin, in line with the probe. Under direct
visualization, the needle is advanced until its tip contacts the anterior
surface of the L4 vertebra (Fig. 47.3B). For unilateral blocks, the
practitioner will need to place the needle tip to the right or left of the
midline of the vertebra. Once the vertebra has been contacted, the needle
is withdrawn 2 to 3 mm and local anesthesia is injected under direct
vision. A hypoechoic puddle of local anesthetic should appear anterior to
the spine (Fig. 47.3C). When unilateral block is desired, the dose should
be limited to 3 to 5 mL of local anesthetic.
48
Ultrasound-Guided
Genitofemoral Nerve
Block

MICHAEL GOFELD

Background and indication: The genitofemoral nerve is a mixed nerve


formed from the roots of L1 and L2. During its course, the nerve divides
into femoral and genital branches. The genital branch supplies sensory
innervation to the pubic mons and the scrotum or labia majora. The
femoral branch supplies the skin inferior to the medial aspect of the
inguinal ligament. Combination of the genitofemoral nerve block with
the ilioinguinal-iliohypogastric nerve blocks during inguinal
herniorrhaphy has been described. This addition may be advantageous
by means of blocking nociception during periods of sac traction. It may
also be used independently for surgical incisions of the scrotal skin.
Another application of the genitofemoral nerve block is genitofemoral
nerve entrapment neuralgias occurring as a consequence of inguinal
surgery.

Anatomy: The genitofemoral nerve receives contributions from the L1


and L2 ventral rami. At the level of the third and fourth lumbar
vertebrae, the genitofemoral nerve passes through the psoas muscle and
continues extraperitoneally, on the ventral surface of the muscle, deep to
the psoas fascia. The nerve typically divides into femoral and genital
branches at a location lateral to the common and external iliac arteries,
although the point of division is variable. After this split, the genital
branch passes along the psoas major and through the internal inguinal
ring of the transversalis fascia. In males, it descends posterior to the
spermatic cord to innervate the cremaster muscle and part of the scrotal
skin. In females, the genital branch travels with the round ligament to
supply sensory innervation to the pubic mons and labia majora. The
relationship of the genital branch to the spermatic cord is extremely
variable with the nerve running outside, inside, ventral, or dorsal to the
spermatic cord. The nerve may also incorporate with the cremaster
muscle. Alternatively, the femoral branch of the genitofemoral nerve
courses superficially outside the inguinal canal, piercing the fascia lata
to terminate in the skin overlying the anterior superior thigh and groin.

Patient position: Supine.

Probe: A broadband high-frequency linear array transducer.

Probe position: Transducer placed in axial orientation, inferior to the


inguinal ligament.

Needle: 22G Quincke-type spinal needle or a stimulating needle.

Local anesthetic: 2 mL of 0.5% bupivacaine or ropivacaine with


epinephrine inside the spermatic cord and 2 mL of 0.5% bupivacaine or
ropivacaine with epinephrine outside the spermatic cord.

Technique: Only the genital branch can be reliably blocked. The skin is
prepared with antiseptic solution, and the transducer is covered with a
sterile sheath. Using color Doppler imaging, the femoral artery is
identified in long axis, caudal to the inguinal ligament. It is expected that
the femoral artery will be located one-third of the distance from the
pubic tubercle to the anterior superior iliac spine. With the femoral
artery in longitudinal view, the probe is moved cephalad until the artery
is visualized, changing its plane to more deep as it approaches the
inguinal ligament. Here, the spermatic cord typically lies superficially
and medially to the femoral artery. The transducer is rotated to the short
axis, and the spermatic cord is identified as an oval or circular
hyperechoic structure (Fig. 48.1). Often, the transducer should be
panned medially toward the pubis. The spermatic cord is identified
within the inguinal ring (Fig. 48.2). After the optimal probe position is
found, the target is brought to the center of the ultrasound image. The
needle is introduced using either an out-of-plane or in-plane technique
(Fig. 48.3). When the needle tip enters the spermatic cord, the syringe is
aspirated. If no blood is withdrawn, local anesthetic is injected in small
increments while observing distention of the tissues with each aliquot
and collection of the hypoechoic fluid encircling the cord structures.
Injection of local anesthetic inside and outside the spermatic cord is
recommended due to anatomic variability of the location of genital
branch of the genitofemoral nerve.
Tips 1. There is significant variability in the cutaneous branching
patterns of the genitofemoral nerve. In some cases, the ilioinguinal nerve
may contribute to the motor and sensory innervation distributions typical
for the genitofemoral nerve.
2. In cases of chronic pain, if diagnostic blockade of the genitofemoral
nerve is successful, surgical neurectomy of the genitofemoral nerve
may be considered in an attempt to provide prolonged relief.
Alternatively, peripheral nerve or spinal cord stimulation may be
recommended.
Suggested Readings 1. Campos N, Chiles J, Plunkett A. Ultrasound-
guided cryoablation of genitofemoral nerve for chronic inguinal
pain. Pain Phys. 2009;12:997–1000.
2. Moore K, Dalley A, Agur A. Clinically Oriented Anatomy. 6th ed.
Philadelphia, PA: Lippincott Williams & Wilkins; 2009.
3. Peng P, Tumber P. Ultrasound-guided interventional procedures for
patients with chronic pelvic pain— a description of techniques and
review of literature. Pain Phys. 2008;11:215–224.
4. Rab M, Ebmer A, Dellon A. Anatomic variability of the ilioinguinal
and genitofemoral nerve: implication for the treatment of groin
pain. Plas Reconstr Surg. 2001;108:1618–1623.
5. Sasaoka N, Kawaguchi K, Yoshitani H, et al. Evaluation of
genitofemoral nerve block, in addition to ilioinguinal and
iliohypogastric nerve block, during inguinal hernia repair in
children. Br J Anaesth. 2005;94:243–246.
49
Ultrasound-Guided
Pudendal Nerve
Block
RACHAEL SEIB AND PHILIP
PENG

Background and indications: Pudendal neuralgia refers to chronic


pelvic pain where pain is experienced in the regions innervated by the
pudendal nerve,1 specifically the clitoris, penis, and perianal area.1–3 Its
prevalence is estimated at approximately 1% of the general population,4
with a female to male preponderance of 2.5:1.5 The pain is typically
burning or tingling in character, and it affects either the entire area
supplied by the pudendal nerve or only a particular branch.6 Classically,
the pain is exacerbated by sitting and alleviated when pressure is
removed from the nerve, as in standing or lying on the unaffected side.7
Anecdotally, the pain is better with sitting on a toilet seat versus a chair.6
Although the cause of pudendal neuralgia is often not readily
identifiable, risk factors include bicycle riding, vaginal delivery,
countertraction devices in orthopedic surgery, pelvic trauma, and certain
kinds of athletic activity.8-12 A pudendal nerve block is an important tool
in the diagnosis of this condition and may also serve therapeutic value.

Anatomy: The pudendal nerve contains both motor and sensory fibers.13
Its sensory fibers supply the clitoris, penis, vulva, and perianal area, and
its motor fibers supply the external anal sphincter and deep muscles of
the urogenital triangle.14,15 It is formed from the anterior rami of the
second, third, and fourth sacral nerves.13 Exiting the pelvis through the
greater sciatic foramen,16 the pudendal nerve is accompanied by the
internal pudendal artery on its medial side (Fig. 49.1). At the level of the
ischial spine, the nerve is situated between the sacrospinous and
sacrotuberous ligaments in what is referred to as the interligamentous
plane.17,18 At this level, 30% to 40% of pudendal nerves will be two-or
three-trunked.15,19

Within the interligamentous plane, the pudendal artery is located lateral


to the pudendal nerve in the vast majority of cases (76% to 100%).20
Following its passage between the two ligaments, the pudendal nerve
travels anteriorly to enter the pelvis through Alcock canal.15,19,21 Alcock
canal (also known as the pudendal canal) is a fascial tunnel created by
the duplication of the obturator internus muscle under the plane of the
levator ani muscle on the lateral wall of the ischiorectal fossa (Fig.
49.2).22 At the level of Alcock canal or just subsequent to it, the
pudendal nerve sprouts three terminal branches: the dorsal nerve of the
clitoris, the inferior rectal nerve, and the perineal nerve, providing the
sensory branches to the skin of the perianal area, labia majora, and
clitoris, respectively (Fig. 49.2).5

The path of the pudendal nerve, sandwiched between the sacrotuberous


and the sacrospinous ligaments at one end and within Alcock canal at
the other, makes it susceptible to entrapment.1,23 Furthermore, the course
of the dorsal nerve of the penis under the subpubic arc exposes the nerve
in cyclists to compression by the nose of the saddle of a bicycle.24 The
configuration of the nerve in the pelvis also makes it susceptible to
stretch during vaginal delivery.25
Although the best imaging technique for Alcock canal is computed
tomography scan,18 ultrasound is useful in imaging the interligamentous
plane at the ischial spine level. To reveal the structures at this level, the
key is to recognize the ischial spine (Fig. 49.3, probe position C). A few
details can be very helpful in identifying the ischial spine: (1) The spine
appears as a straight hyperechoic line, and the ischium cephalad is seen
as a curved line because it forms the posterior aspect of the acetabulum
(Fig. 49.4); (2) the sacrospinous ligament appears as a hyperechoic line
in continuity with the medial end of the ischial spine but in contrast to
bone does not cast an echogenic shadow behind it (Fig. 49.4); (3) the
sacrotuberous ligament is seen as a light hyperechoic line ventral to the
gluteus maximus muscle overlapping the fascia ventral to this muscle
and appears parallel and dorsal to the sacrospinous ligament (Fig. 49.4);
and (4) the internal pudendal artery can be localized with the use of
color-flow Doppler in close proximity to the ischial spine.20 Another
arterial pulsation is often seen lateral to the tip of the ischial spine and is
accompanied by the sciatic nerve. This is the branch of the inferior
gluteal artery. Mistaking this artery for the pudendal artery will result in
sciatic nerve block.
Patient position: Prone with a pillow to support the pelvis.

Transducer: 40-to 60-mm curved array oscillating at 2 to 5 MHz.

Transducer orientation: Initially transverse over sacrum and then as the


probe is moved laterally, it should be aligned with the direction of the
piriformis muscle (Fig. 49.2). Instead of placing the transducer over the
sciatic notch, the transducer should be placed more laterally and tilted
medially. This gives the best visualization of the sciatic notch.

Needle: 22G, 120-mm, or 3.5-inch spinal needle.

Local anesthetic: 4 mL of 0.25% bupivacaine + 40 mg


methylprednisolone.
Other equipment: 3-mL syringe; 25G hypodermic needle; 2% lidocaine
for skin anesthesia; short extension tubing flushed with normal saline; 5
to 10 mL of normal saline for confirmation of needle placement in the
interligamentous plane.

Technique The patient is placed in the prone position. A curvilinear


probe with a low frequency (2 to 5 MHz) is required to visualize the
structures given its location deep to the gluteus maximus muscle.27 The
ultrasound transducer should be first placed in a transverse direction
lateral to the posterior superior iliac spine to visualize the ilium as a
hyperechoic line descending laterally (Figs 49.3 and 49.4A). After the
iliac wing has been identified, the angle of the probe is rotated such that
it would be in line with the long axis of the piriformis muscle as it runs
from the sacrum to the greater trochanter (Fig. 49.3).27 As scanning
continues caudally, the hyperechoic line of the ilium starts to disappear
in the medial aspect of the image at the level of the sciatic notch. The
lateral aspect of the ultrasound image continues to display a curved
hyperechoic line representing the ischium.26 At this level, two separate
muscular layers are identified: the gluteus maximus and the piriformis
(Fig. 49.4B). Scanning slightly more caudally, the curved line of the
ischium becomes straighter as it transitions to the ischial spine (Fig.
49.4C).26 At this point, the sacrospinous ligament should be seen as a
slight hyperechoic structure extending from the tip of the ischial spine,
projecting toward the sacrum (Fig. 49.4C).27 However, in contrast to
bony structures, the sacrospinous ligament will not cast an anechoic
shadow behind it. The sacrotuberous ligament may also be seen at this
level deep to the gluteus maximus muscle. The piriformis muscle will no
longer be visualized. The pudendal artery lies slightly medial to the tip
of the ischial spine and the pudendal nerve situated medial to the artery
(Fig. 49.4D).26 Owing to its small diameter and the depth in the gluteal
region, the nerve itself is only clearly visualized under ultrasonography
about 30% of the time.20 It is rare (<2%) that neither the nerve nor the
artery is visualized.28
The success of the block depends on optimization of the ultrasound
image before needle insertion. The image must clearly show the
interligamentous plane formed by the sacrotuberous and sacrospinous
ligaments as well as the pudendal artery (Fig. 49.4).27 The purpose of
image optimization is to ensure that the injected solution is well
contained between the ligaments where the nerve is potentially
compressed.27 The image quality of these soft tissues should not be
compromised by tilting the probe to find the needle. The needle must
instead be reinserted or repositioned until it is within the path of the
ultrasound beam. The needle is inserted 2 cm medial to the probe so the
tip comes to lie medial to the pudendal artery (Figs. 49.3 and 49.4D).27
This enhances visualization of the needle. As the needle passes through
the sacrotuberous ligament, increased resistance against the needle is
felt, and a “pop” may be felt as it passes through the ligament.27 The
needle tip should now be between the sacrotuberous and sacrospinous
ligaments.27 A small volume of normal saline may be injected to serve
as a contrast.20 The solution will appear as a hypoechoic collection in
the plane between the sacrotuberous and sacrospinous ligaments. A
solution containing local anesthetic and steroid (5 mL) is injected when
the final needle position is deemed satisfactory. If solution spreads
laterally past the pudendal artery, sciatic nerve involvement becomes
more likely, and the needle should be repositioned more medially.

Summary of evidence The pudendal nerve may be blocked using


transvaginal,29 transperineal,30,31 and transgluteal approaches.32
The transgluteal approach is the most popular, as it allows a
fluoroscopy-guided approach using the ischial spine as a landmark.
Fluoroscopic approaches to the interligamentous plane have been
described and validated.32 The main disadvantage with fluoroscopy
is that the interligamentous plane is not visualized, and the
practitioner must rely on surrogate landmarks and clinical
judgment about needle depth. Furthermore, because the proximity
of the sciatic nerve to the needle tip is not known, there is a real risk
of sciatic nerve block.
The feasibility of ultrasound-guided visualization of the pudendal nerve
has been tested in humans and cadavers.16,20,28 Only one study actually
tested an ultrasound-guided pudendal nerve block.20 In that study, using
the technique described earlier, the pudendal nerve was visualized 88%
of the time. The pudendal artery was more consistently visualized, seen
in 16 of 17 subjects. This and other studies confirm that the internal
pudendal artery can be used as a surrogate landmark for the pudendal
nerve when the latter is not visualized. The internal pudendal artery is
located within 10 mm medial to the ischial spine.20 Pulsations lateral to
the ischial spine probably represent the inferior gluteal artery and an
injection performed here will likely result in a sciatic nerve block.20
Although Rofaeel et al.20 noted that use of ultrasound resulted in more
medially placed needles than under the traditional fluoroscopic-guided
procedure, a sensory block was achieved in all patients in spite of, or
perhaps because of, this. A recent randomized controlled trial comparing
pudendal nerve blocks performed under ultrasound with those performed
using ultrasound guidance33 demonstrated that the ultrasound-guided
block afforded comparable accuracy to the fluoroscopy-guided
procedure but had the added advantage of allowing visualization of the
sciatic nerve and the spread of injectate.

References
1. Robert R, Prat-Pradal D, Labat JJ, et al. Anatomic basis of chronic
perineal pain: role of the pudendal nerve. Surg Radiol Anat.
1998;20:93–98.
2. Benson JT, Griffis K. Pudendal neuralgia, a severe pain syndrome.
Am J Obstet Gynecol. 2005;192:1663–1668.
3. Amarenco G, Kerdraon J, Bouju P, et al. Treatments of perineal
neuralgia caused by involvement of the pudendal nerve. Rev
Neurol. 1997;153:331–334.
4. Spinosa JP, de Bisschop E, Laurencon J, et al. Sacral staged reflexes
to localize the pudendal compression: an anatomical validation of
the concept [in French]. Rev Med Suisse. 2006;2:2416–2421.
5. Peng PWH, Antolak Jr SJ, Gordon AS. Pudendal neuralgia. In:
Pukall C, Goldstein I, Goldstein A, eds. Female Sexual Pain
Disorders. Hoboken, NJ: Wiley-Blackwell;2009:112–118.
6. Hibner M, Desai N, Robertson LJ, et al. Pudendal neuralgia. J
Minim Invasive Gynecol. 2010;17:148–153.
7. Ramsden CE, McDAniel MC, Harmon RL, et al. Pudendal nerve
entrapment as a source of intractable perineal pain. Am J Phys Med
Rehabil. 2003;82:479–484.
8. Leibovitch I, Mor Y. The vicious cycling: bicycling related
urogenital disorders. Eur Urol. 2005;47:277–287.
9. Allen RE, Hosker GL, Smith AR, et al. Pelvic floor damage and
childbirth: a neurophysiological study. Br J Obstet Gynaecol.
1990;97:770–779.
10. Amarenco G, Ismael SS, Bayle B, et al. Electrophysiological
analysis of pudendal neuropathy following traction. Muscle Nerve.
2001;24:116–119.
11. Soulie M, Vazzoler N, Seguin P, et al. Urological consequences of
pudendal nerve trauma during orthopedic surgery: review and
practical advice. Prog Urol. 2002;12:504–509.
12. Antolak S, Hough D, Pawlina W, et al. Anatomical basis of chronic
pelvic pain syndrome: the ischial spine and pudendal nerve
entrapment. Med Hypotheses. 2002;59:349–353.
13. Juenemann K-P, Lue TF, Scmidt RA, et al. Clinical significance of
sacral and pudendal nerve anatomy. J Urol. 1988;139:74–80.
14. Schraffordt SE, Tjandra JJ, Eizenberg N. Anatomy of the pudendal
nerve and its terminal branches: a cadaver study. ANZ Journal of
Surgery 2004;74:23–26.
15. Shafik A, Doss SH. Pudendal canal: surgical anatomy and clinical
implications. Am Surg. 1999;65:176–180.
16. Gruber H, Kovacs P, Piegger J, et al. New, simple, ultrasound-
guided infiltration of the pudendal nerve: topographic basics. Dis
Colon Rectum. 2001;44:1376–1380.
17. Mahakkanukrauh P, Surin P, Vaidhayakarn P. Anatomical study of
the pudendal nerve adjacent to the sacrospinous ligament. Clin
Anat. 2005;18:200–205.
18. Hough DM, Wittenberg KH, Pawlina W, et al. Chronic perineal
pain caused by pudendal nerve entrapment: anatomy and CT-guided
perineural injection technique. AJR Am J Roentgenol.
2003;181:561–567.
19. O’Bichere A, Green C, Phillips RK. New, simple approach for
maximal pudendal nerve exposure: anomalies and prospects for
functional reconstruction. Dis Colon Rectum. 2000;43:956–960.
20. Rofaeel A, Peng P, Louis I, et al. Feasibility of real-time ultrasound
for pudendal nerve block in patients with chronic perineal pain. Reg
Anesth Pain Med. 2008;33:139–145.
21. Thompson JR, Gibbs S, Genadry R, et al. Anatomy of pelvic
arteries adjacent to the sacrospinous ligament: importance of the
coccygeal branch of the inferior gluteal artery. Obstet Gynecol.
1999;94:973–977.
22. Amarenco G, Lanoe Y, Ghnassia RT, et al. Alcock’s canal syndrome
and perineal neuralgia [in French]. Rev Neurol (Paris).
1988;144:523–526.
23. Robert R, Labat JJ, Bensignor M, et al. Decompression and
transposition of the pudendal nerve in pudendal neuralgia: a
randomized controlled trial and long-term evaluation. Eur Urol.
2005;47:403–8.
24. Sedy J, Nanka O, Belisova M, et al. Sulcus nervi dorsalis
penis/clitoridis: anatomic structure and clinical significance. Eur
Urol. 2006;50:1079–1085.
25. Ashton-Miller JA, Delancey JOL. Functional anatomy of the female
pelvic floor. Ann N Y Acad Sci. 2007;1101:266–296.
26. Peng PWH, Tumber PS. Ultrasound-guided interventional
procedures for patients with chronic pelvic pain—a description of
techniques and review of the literature. Pain Physician.
2008;11:215–224.
27. Bellingham G, Peng PWH. Ultrasound-guided interventional
procedures for chronic pelvic pain. Tech Reg Anesth Pain Manag.
2009;13:171–178.
28. Kovacs P, Gruber H, Piegger J, et al. New, simple, ultrasound-
guided infiltration of the pudendal nerve: ultrasonographic
technique. Dis Colon Rectum. 2001;44:1381–1385.
29. Bowes WA. Clinical aspects of normal and abnormal labour. In:
Resnick R, Creasy RK. Maternal-Fetal Medicine: Principles and
Practice. 2nd ed. Philadelphia, PA: WB Saunders;1989;510–546.
30. Naja Z, Ziade MF, Lonnqvist PA. Nerve stimulator-guided pudendal
nerve block decreases posthemorrhoidectomy pain. Can J Anaesth.
2005;52:62–68.
31. Imbelloni LE, Viera EM, Gouveia MA, et al. Pudendal block with
bupivacaine for postoperative pain relief. Dis Colon Rectum.
2007;50:1656–1661.
32. Prat-Pradal D, Metge L, Gagnard-Landra C, et al. Anatomical basis
of transgluteal pudendal nerve block. Surg Radiol Anat.
2009;3:289–293.
33. Bellingham GA, Bhatia A, Chan CW, et al. Randomized controlled
trial comparing pudendal nerve block under ultrasound and
fluoroscopic guidance. Reg Anesth Pain Med. 2012;37:262–266.
50
Neuraxial Anatomy
Relevant to the Two-
Dimensional
Ultrasound
Examination

DAVID BELAVY

The anatomy and sonoanatomy of the vertebral column vary between the
cervical, thoracic, lumbar, and sacral spine, but the key anatomical
landmarks are common to both landmark, two-dimensional (2D) and
three-dimensional (3D) ultrasound-guided procedures. The bones of the
vertebral column are well visualized in ultrasound imaging, but the
ligaments and neural elements are less well seen.
The anatomical features of lumbar vertebrae can usually be well defined
using ultrasound. The vertebral body supports the vertebral arch, which
is made of the two pedicles and two laminae (Fig. 50.1 and 50.2). The
posterior surface of the vertebral body and the vertebral arch enclose the
vertebral foramen. Within the vertebral foramen are the targets for
anesthetic interventions, the epidural and intrathecal spaces (Fig. 50.3).
The processes supported by the vertebral arch form important
anatomical landmarks for imaging. The spinous processes mark the
midline, and the articular processes mark the lateral limit of the lamina.
Further lateral to the articular processes are the transverse processes,
which are important landmarks for paravertebral anesthesia (Fig. 50.3).
The identification of the vertebral anatomy forms the basis of a proposed
systematic approach to the 2D examination (Table 50.1).1 The
examination begins laterally from the transverse processes and moves
medially over the articular processes and laminae. The level for insertion
is then determined by counting up from the sacrum or 12th rib. This
approach is used to mark a needle insertion point and determine an angle
of needle approach for neuraxial anesthesia, the so-called scan and mark
approach.
The sequence of views acquired is less important than the application of
a systematic and efficient examination.
Between spinous processes of adjacent vertebrae run the thick cord of
the supraspinous ligament and the membranous interspinous ligaments.
During midline landmark-guided neuraxial procedures, the needle is
passed through these ligaments into the ligamentum flavum, which runs
between the laminae of adjacent vertebrae. Ultrasound imaging through
these structures is usually poor because of their density, so neuraxial
ultrasound is generally performed in a paramedian location with the
probe directed obliquely toward the midline.
Identifying the space between the laminae of adjacent vertebrae using
ultrasound is important in identifying a suitable approach for needle
insertion. The presence of a space between laminae can be confirmed by
visualization of deeper structures. This “acoustic window” to deeper
structures increases confidence that a needle can successfully be inserted
between the laminae. The posterior border of the vertebral body and
intervertebral discs, which are covered by the posterior longitudinal
ligament, are the deepest structures that can be seen between the
laminae. (Fig. 50.3). The spinal cord or cauda equina can be seen in
some patients, particularly children, through the vertebral foramen The
sonoanatomy of thoracic vertebrae differs in that their spinous processes
and laminae will overlap the vertebra below to a greater extent. As a
consequence, the interlaminar space and acoustic window to deeper
structures is smaller and the transverse interlaminar view (Fig. 50.4)
often does not demonstrate the vertebral canal. Just off the midline, in
sagittal view (Fig. 50.5), the thoracic articular processes may be seen.
Slightly more lateral with the probe in the sagittal plane and a slight
medial orientation (Fig. 50.6), the interlaminar space can be seen. The
ribs can also be seen laterally with the tubercle of the rib articulating
with the transverse process by using an oblique transverse orientation.
This is relevant to paravertebral anesthesia (Fig. 50.7).
The third to the sixth cervical vertebrae have short bifid spinous
processes. C7 differs, with a long spinous process that is not bifid. The
laminae are thinner than at lower levels. The spinal cord is often visible,
even in adults in cervical paramedian view (Fig. 50.8) and cervical
oblique interlaminar view (Fig. 50.9). The transverse processes are
pierced by the vertebral artery and vein, which may be visible with color
Doppler ultrasound. The ligamentum flavum between laminae is often
indistinct but can be seen in some patients.
Reference 1. Chin KJ, Karmakar MK, Peng P. Ultrasonography of
the adult thoracic and lumbar spine for central neuraxial blockade.
Anesthesiology. 2011;114:1459–1485.
51
Ultrasound-Guided
Third Occipital
Nerve Block and
Cervical Medial
Branch Block
DANIEL L. KRASHIN AND MICHAEL GOFELD

Background and indication


The third occipital nerve (TON) and cervical facet joints are common
sources of cervicalgia and headache. These conditions can arise
spontaneously or secondary to trauma or to surgery. Given the complex
innervation of the head and neck, diagnostic blocks are the only reliable
way to precisely identify the source of pain arising from the TON or
cervical zygapophyseal joints. If two separate blinded diagnostic blocks
on separate days are positive in a concordant fashion, the patient may be
a candidate for neurotomy or radiofrequency ablation.
The TON supplies the C2–C3 facet joints and also a small area of skin
inferior to the occiput. The lower cervical medial branches innervate the
facet joints and the cervical multifidus muscles. Specific pain referral
patterns have been identified for each level of facet joint, but there is
considerable variation between patients.
Radiofrequency neurotomy has been an evidence-based treatment of
choice for chronic pain related to whiplash syndrome, although it is more
widely used for the management of neck pain attributed to degenerative
changes. Isolated cervical facet joint arthropathy may also be treated
with intra-articular injections of local anesthetic and corticosteroid to
treat acute inflammation. Medial branch analgesic injections should be
performed to confirm the pain generator and exclude other conditions.

Anatomy
The cervical zygapophyseal joints have uniquely variable innervation.
The C2–C3 joint is exclusively innervated by TON, which is essentially
the posterior ramus of the C3 root, much thicker than the ventral ramus.
TON crosses the C2–C3 facet joint, and its course can vary from the
apex of the C3 superior articular process (SAP) to adjacent to the C2–C3
intervertebral foramen. The most common location is the lower half of
the convexity of the C2–C3 facet joint. Care must be taken to avoid
confusing this nerve with the C3 medial branch. The nerve is usually
elevated over the surface of the bone by 1 to 2 mm.
Each cervical facet joint below the C2–C3 level is innervated by
articular branches of the medial branch nerves from the levels above and
below that joint. These medial branches arise from the cervical dorsal
rami, which come off the spinal nerve and pass over the base of the
transverse process. The medial branches come off these dorsal rami and
wrap medially around the articular pillars, held in place by fascia and by
the semispinalis capita tendon. This fixed location allows the medial
nerves to be targeted in an area with clear bony landmarks that is not
adjacent to the spinal nerve or vertebral artery. The medial branch of the
C7 dorsal ramus is slightly cephalad in its course compared to the other
cervical levels, and crosses over the SAP of that vertebra.
The zygapophyseal joints themselves are formed by the articulation of
the superior and inferior articular processes of adjacent cervical
vertebrae. The orientation of the facet joints changes from being 45
degrees superior to the transverse plane at the C2–C3 level to nearly
vertical at the cervicothoracic junction.
Third occipital nerve block

Patient Position: Lateral decubitus position with head on a pad to keep


neck in a neutral position.

Transducer: A broadband high-frequency linear array transducer.

Transducer Position: Transverse, just caudal to mastoid process on either


side. The vertebral artery is identified just caudal to this and is followed
further caudally to the C2 transverse process, and further caudally to
C2–C3 facet joint. The apex of this joint is identified and used as a target
for needle insertion. Then the transducer is rotated 90 degrees to a
longitudinal position. Gentle adjustment of the transducer is typically
required to obtain a clear “wave-shaped” line of the articular processes
and joints. The “valleys” represent center of articular pillars, and the
“peaks” correspond to the joints themselves (Fig 51.1). The most cranial
“peak” that found is the C2–C3 joint. Cranially to it, the characteristic
image will reflect descending C2 lamina and the C1–C2 interspinous
space. The cross section of the TON may be visible crossing the C2–C3
joint in this plane (Fig. 51.2). This area is marked on the skin.
Needle: 2-to 3.5-inch, 22G needle.

Local Anesthetic: 0.5 mL of 0.5% bupivacaine or 0.5 mL of 1% lidocaine.

Technique: After sterile skin preparation, a needle is advanced via an out-


of-plane technique with the transducer again held in a transverse
orientation. The needle tip should be seen exactly at the C2–C3 facet
joint line or adjacent to the cross-sectional image of the TON. At this
point, 0.5 mL of 0.5% bupivacaine or 0.5 mL of 1% lidocaine is injected
and the needle is withdrawn. Alternatively and when the patient’s
anatomy is favorable (long flexible neck), the injection can be performed
in an in-plane cephalad to caudad approach (Fig. 51.3).

Tips
1. The area of skin just inferior to the external occipital prominence is
innervated by the TON and can be used to test the adequacy of the
block.
2. The TON may be very difficult to visualize in bariatric patients.
Cervical medial branch block

Patient Position: Lateral decubitus position.

Probe: A broadband high-frequency linear array transducer.

Probe Position: Transverse or longitudinal to the cervical spine.

Needle: 2-to 3.5-inch 22G Quincke-type spinal needle.

Local Anesthetic: 0.3 mL of 0.5% bupivacaine or 0.3 mL of 2% lidocaine.

Technique: The vertebrae are identified using the transducer held in a


paravertebral coronal orientation to identify the appropriate
intervertebral space by counting and identifying transverse processes.
Usually, the C6 bifurcated transverse process with the prominent anterior
(Chassaignac) tubercle is identified first (Fig. 51.4). Caudal scanning is
performed to find the C7 transverse process, which has only one tubercle
and is oriented more coronally (Fig. 51.5). The rest of the cervical spinal
levels are found by counting transverse processes cephalad.
Alternatively, the count can be performed starting from the C2–C3 joint
in the longitudinal view as described previously, and the corresponding
articular pillars are counted and marked. The later method is useful
when the injection performed in an out-of-plane technique.
The ultrasound transducer is positioned at a 90-degree angle to the long
axis of the vertebra so that both the transverse processes and the articular
pillars are seen. An imaginary line from the transverse process will
intersect the target for the needle at the middle of the articular pillar. The
needle is inserted dorsoventrally, using an in-plane technique under
ultrasound visualization, until it contacts the articular pillar posterior to
the transverse process (Fig. 51.6). During the injection of local
anesthetic, the surrounding tissue should be monitored for visible
expansion.

An out-of-plane injection is performed similarly to the TON block,


aiming the needle tip to the center of articular pillar (the “valley”). It is
important to perform the injection in a ventral to dorsal direction
standing in front of the patient. In case the needle “missed” the pillar, it
will advanced into the posterior neck muscles. If the needle was inserted
in a dorsoventral direction, it may inadvertently cause damage of the
anterior neck nerve and vascular structures (Fig. 51.7).

Tips

1. Insert the needle from anterior to posterior to decrease the risk of


vascular injury.
2. Identifying the proper level is all-important for reliable results;
practitioners must count vertebrae and mark the location before then
performing the block using a sterile technique.
Bibliography
Eichenberger U, Greher M, Kapral S, et al. Sonographic visualization and ultrasound-guided block
of the third occipital nerve: prospective for a new method to diagnose C2-C3 zygapophysial
joint pain. Anesthesiology. 2006;104:303–308.
Galiano K, Obwegeser AA, Bodner G, et al. Ultrasound-guided facet joint injections in the middle
to lower cervical spine: a CT-controlled sonoanatomic study. Clin J Pain. 2006;22:538–543.
Galiano K, Obwegeser AA, Bodner G, et al. Ultrasound-guided periradicular injections in the
middle to lower cervical spine: an imaging study of a new approach. Reg Anesth Pain Med.
2005;30:391–396.
Gofeld M. Ultrasonography in pain medicine: a critical review. Pain Pract. 2008;8:226–240.
Narouze S, Peng PW. Ultrasound-guided interventional procedures in pain medicine: a review of
anatomy, sonoanatomy, and procedures. Part II: axial structures. Reg Anesth Pain Med.
2010;35:386–396.
Narouze S, Vydyanathan A. Ultrasound-guided cervical facet intra-articular injection. Tech Reg
Anesth Pain Manag. 2009;13:133–136.
Narouze S, Vydyanathan A, Kapural L, et al. Ultrasound-guided cervical selective nerve root
block: a fluoroscopy-controlled feasibility study. Reg Anesth Pain Med. 2009;34:343–348.
Siegenthaler A, Schliessbach J, Curatolo M, et al. Ultrasound anatomy of the nerves supplying the
cervical zygapophyseal joints: an exploratory study. Reg Anesth Pain Med. 2011;36:606–610.
52
Ultrasound-Guided
Cervical Nerve Root
Injections
DANIEL L. KRASHIN AND MICHAEL GOFELD

Background and indications: Ultrasound-guided cervical transforaminal


injections have been described by Galiano et al.1 who performed the first
cadaveric study. The technique has been criticized due to the potential
risk of spinal cord injury and intravascular (radicular artery) injection
with the ventrodorsal oblique approach. Narouze et al.2 studied
ultrasound-guided cervical nerve root injections with confirmatory
fluoroscopy. In this study, the target was the posterior aspect of the
neural foramen, anterior to the superior articular process, as seen in the
oblique view. Using this technique, he was able to demonstrate needle
placement within 3 mm of the target in all patients in the lateral oblique
view and within 8 mm in the anterior–posterior (AP) view. Because
cervical foramina are oriented obliquely ventrally but the needle is
inserted obliquely dorsally, a proper transforaminal spread may not
always be possible, and this method should be rather named as the
ultrasound-guided selective nerve root injection. Nevertheless, it may be
actually preferential practice when a diagnostic analgesic blockade is
indicated in the case of cervical radiculopathy.

Anatomy: Anatomy is as described previously. The Chassaignac tubercle


is the most prominent bony protuberance of the C6 transverse process,
and it is readily palpable on the anterior neck. An axial image at this
location will show the typical appearance of the anterior and posterior
tubercles with the exiting nerve root in between. Identification of the
remaining exiting nerves can be performed by scanning caudad, for C7,
and cephalad for nerve roots C3–C5. A color Doppler should routinely
be used to localize radicular arteries and to avoid vascular injury.

Patient position: Lateral decubitus with head placed on the pillow and
the upper shoulder rotated backward to allow better access to the lateral
neck.

Probe: A broadband high-frequency linear array transducer.

Probe position: Short axis to the corresponding vertebra.

Needle: 25G Quincke-type spinal needle.

Local anesthetic: 0.5 to 1 mL of 2% lidocaine and a water-soluble


corticosteroid (e.g., preservative-free dexamethasone).

Technique: Intravascular access should be established, and vital signs


must be monitored. Although only small amount of local anesthetic is
injected, inadvertent spread into the neuraxial space or an intra-arterial
injection may happen. Skin is prepped and draped, and the ultrasound
transducer is placed into a sterile sleeve. Sterile conductive gel is
applied. Short-axis scanning is performed, and spinal levels are
identified. At the target level, the transverse process is visually
positioned in the center of ultrasound image. The exiting nerve root is
seen between the anterior and posterior tubercles or, at the C7 level, just
anteriorly to the transverse process. Prior to injection, a color Doppler
test is performed to exclude any aberrant vessels that can be trespassed
by the needle. In addition, the location of radicular arteries is identified.
The block needle is inserted in an in-plane method aiming to the exiting
nerve root. To achieve reliable block and to avoid complications, the
needle should be positioned adjacent to the nerve between the 9 and 12
o’clock position (Fig. 52.1). One mL of local anesthetic with or without
10 mg of dexamethasone is slowly injected.

Tips
1. Always start orientation scanning with finding the C6 and C7
vertebrae in their short axis.
2. Apply color Doppler before injection to avoid vascular damage and
intravascular injection.
3. Place the needle tip at 6 to 9 o’clock position adjacent to the exiting
nerve root.
References
1. Galiano K, Obwegeser AA, Bodner G, et al. Ultrasound-guided
periradicular injections in the middle to lower cervical spine: an
imaging study of a new approach. Reg Anesth Pain Med.
2005;30:391–396.
2. Narouze S, Vydyanathan A, Kapural L, et al. Ultrasound-guided
cervical selective nerve root block: a fluoroscopy-controlled
feasibility study. Reg Anesth Pain Med. 2009;34:343–348.
Suggested Readings
Eichenberger U, Greher M, Kapral S, et al. Sonographic visualization and ultrasound-guided block
of the third occipital nerve: prospective for a new method to diagnose C2-C3 zygapophysial
joint pain. Anesthesiology. 2006;104:303–308.
Galiano K, Obwegeser AA, Bodner G, et al. Ultrasound-guided facet joint injections in the middle
to lower cervical spine: a CT-controlled sonoanatomic study. Clin J Pain. 2006;22:538–543.
Gofeld M. Ultrasonography in pain medicine: a critical review. Pain Pract. 2008;8:226–240.
Narouze S, Peng PW. Ultrasound-guided interventional procedures in pain medicine: a review of
anatomy, sonoanatomy, and procedures. Part II: axial structures. Reg Anesth Pain Med.
2010;35:386–396.
Narouze S, Vydyanathan A. Ultrasound-guided cervical facet intra-articular injection. Tech Reg
Anesth Pain Manag. 2009;13:133–136.
Siegenthaler A, Schliessbach J, Curatolo M, et al. Ultrasound anatomy of the nerves supplying the
cervical zygapophyseal joints: an exploratory study. Reg Anesth Pain Med. 2011;36:606–610.
53
Ultrasound-Guided
Lumbar
Transforaminal
Epidural Steroid
Injection
KACEY A. MONTGOMERY AND MICHAEL

GOFELD

Background and indications: Lumbar transforaminal epidural steroid


injections are common interventional procedures used in the treatment of
lower extremity radicular pain. Injections are performed at the level that
corresponds with the dermatomal pain pattern described by the patient.
Nerve root irritation can be a result of direct compression from a
herniated disk, spondylosis causing neural foraminal narrowing, or disk
rupture causing a chemical irritation.

Anatomy: Transforaminal epidural steroid injections are typically done


for radicular pain. Placement of a corticosteroid along the appropriate
nerve root rather than in the dorsal aspect of the epidural space provides
better results in treating radicular pain. The lumbar foramina, where the
nerve roots exit from, face laterally and are bordered by the vertebral
body superiorly, the inferior articular process (IAP) posterosuperiorly,
the superior articular process (SAP) posteroinferiorly and the pedicles
above and below. Injury to the artery of Adamkiewicz, entering the
spinal canal from T7–L4, can cause anterior spinal artery syndrome and
paraplegia if injured in the performance of injections, and thus, one
should ensure not to traverse or inject particulate steroids into this artery
as well as the smaller radicular vessels.

Patient position: Prone with towel rolls placed under the abdomen for
alleviation of lumbar lordosis.

Probe: Broadband curved or linear array transducer depending on the


depth expected.

Probe position: Initial probe orientation is performed as outlined


previously. For the procedure itself, the transducer is placed in the
transverse/short-axis position.

Needle: 25G 3.5-or 5-inch spinal needle.

Local anesthetic: 2 mL of 1% lidocaine with 10 mg of dexamethasone


(nonparticulate steroid).

Technique: The skin is cleansed and prepped in the usual sterile fashion
with utilization of a sterile ultrasound probe cover. A sonographic
evaluation for the appropriate level is done as described previously. In
the short axis, the lumbar vertebra should be identified between two
corresponding transverse processes (Fig. 53.1A,B). The identification of
the spinous process, lamina, and the dorsal part of the vertebral body is
necessary. The exiting spinal nerve may be seen as a faint double-
hyperechoic structure and should be avoided by moving the transducer
slightly caudad or cephalad. Color Doppler must be applied prior to
needle placement to identify radicular vessels or the arteria magna. In
such case, the transducer position should be modified. Most commonly,
the intervertebral disk rather than the bony surface is seen. The target is
the medial aspect of the visible dorsal vertebra or intervertebral disk.
The disk usually appears as more homogenic anechoic contour. Utilizing
an in-plane approach, the needle is guided to the most medial aspect of
the vertebral body followed by the chosen injectate as listed previously
(Fig. 53.1B,C).

Tips

1. Devastating complications can result with the use of particulate


steroids. Because one of the limitations of ultrasound-guided
procedures is the inability to assess the intravascular injection of
contrast dye preceding the injectate, the use of nonparticulate steroid
is recommended.
2. One other drawback to ultrasound-guided transforaminal injections is
the inability to assess the extent of epidural spread.
3. The most medial visible part of the vertebral body or intervertebral
disk is chosen as the target point. This will ensure the needle
placement into the ventral foramen and prevent inadvertent
advancement into the neuraxial compartment.
Suggested Readings
Galiano K, Obwegeser AA, Bodner G, et al. Real-time sonographic imaging for periradicular
injections in the lumbar spine: a sonographic anatomic study of a new technique. J Ultrasound
Med. 2005;24:33–38.
Gofeld M. Ultrasonography in pain medicine: a critical review. Pain Pract. 2008;8:224–240.
Gofeld M, Bristow S, Chiu S, et al. Ultrasound-guided lumbar transforaminal injections:
feasibility and validation study. Spine. 2012;37:808–812.
Narouze S, Peng P. Ultrasound-guided interventional procedures in pain medicine: a review of
anatomy, sonoanatomy, and procedures. Part II: axial structures. Reg Anesth Pain Med.
2010;35:386–396.
Sato M, Simizu S, Kadota R, et al. Ultrasound and nerve stimulation-guided l5 nerve root block.
Spine. 2009;34:2669–2673.
Sekhadia M, Benzon H. Selective nerve root blocks and transforaminal epidural steroid injections.
In: Benzon H, Raja S, Liu S, et al., eds. Essentials of Pain Medicine Philadelphia, PA:
Elsevier-Saunders;2011:314–321.
54
Ultrasound-Guided
Lumbar Facet Medial
Branch Block and
Intra-articular Facet
Joint Injection
KACEY A. MONTGOMERY AND MICHAEL

GOFELD

Background and indications: Zygapophyseal joint (facet joint)


syndrome is a common cause of axial low back pain among patients that
present to chronic outpatient pain clinics. Patients may complain of back
pain, buttock pain, or pain referred to the posterior thigh or inguinal
area. Lumbar spondylosis or facet joint osteoarthritis–mediated pain is
transmitted via lumbar medial branch nerves. These nerves can be
blocked for diagnostic purposes and eventually ablated if indicated and
confirmed as the source of pain. Alternatively, for therapeutic purposes,
intra-articular facet joint steroid injections can be performed.

Anatomy: Each vertebra is connected to the vertebra above and below by


facet joints posteriorly and intervertebral discs anteriorly. The facet joint
is formed by the inferior articular process (IAP) from the vertebra above
and the superior articular process (SAP) from the vertebra below and can
accumulate a volume of approximately 1 to 1.5 mL. The superior facet
surface has a concave shape that articulates with the convex surface of
the inferior facet (Fig. 54.1A to C). These articulations allow flexion but
limit rotation. A segmental nerve root exits the corresponding neural
foramen and divides into the ventral and dorsal rami. The dorsal ramus
has the medial, intermediate, and lateral branches. The facet joint is
innervated by the medial branch at that corresponding level and the level
above that must be anesthetized in order to block pain generated by a
single facet joint. If the pathology exists at a single lumbar level, the
median branch at that level as well as at the level above it must be
blocked to provide pain relief because of the dual innervation of the
joint.
The location of the medial branch nerve is uniformly located at the base
of the SAP between the neural foramen and the mammillo-accessory
ligament. After innervating the multifidus muscle, the medial branch
gives off the superior and inferior articular branches to the
corresponding facet joint above and below (see Fig. 54.1A, B). The L5
medial branches have a variable course. Thus, the L5 dorsal ramus is
targeted at the junction of S1 SAP and the sacral ala (see Fig. 54.1C).

Patient position: Prone with towel rolls placed under the abdomen to
alleviate lumbar lordosis.

Probe: Broadband low-frequency curved array transducer.

Probe position: Short axis/transverse.

Needle: 22G or 25G 3.5-inch or 5-inch spinal needle.

Local anesthetic: For intra-articular facet joint injections, 1 mL of a


mixture of local anesthesia and depo-steroid (i.e., 0.5 mL of 2%
lidocaine and 0.5 mL of methylprednisolone [80 mg/mL]). For medial
branch block, 0.5 mL of local anesthesia is injected at the target site,
either 0.5% bupivacaine or 2% lidocaine.

Technique: Sterile prep and drape of skin and ultrasound probe


performed per usual standard. After the liberal application of a sterile
ultrasound gel, midline sagittal/longitudinal scanning is undertaken to
map corresponding spinous processes starting at the easily identifiable
sacrum and counting levels while moving from caudad to cephalad (Fig.
54.2). Spinous processes, lamina, interlaminar spaces, and facet joints
are identified during the scanning process. With the transducer in a
longitudinal orientation, the probe is moved laterally until a “saw-tooth”
hyperechoic line is identified, which corresponds to the SAP and IAP
(Fig. 54.3). For confirmation, transverse scanning is done in the short
axis and levels are reidentified and marked. Inferiorly and lateral to the
spinous processes are the facet joints and further inferiorly and lateral
are the transverse processes (TPs).
For intra-articular joint injections, the opening between the SAP and IAP
are identified (Fig. 54.4). If an opening is not readily seen, the target
location is just medial to the SAP. The transducer is placed in the short
axis at the corresponding lumbar level. After identification of the gap
corresponding to the intra-articular facet joint injection, a 22G, 3.5-inch
or 5-inch Quincke-type spinal needle is inserted via an in-plane
approach (Fig. 54.5). Once the needle touches the medial aspect of the
SAP or enters the joint space, 1 mL of the previously stated medication
is injected.
For medial branch blocks and L5 dorsal ramus nerve injections,
longitudinal scanning is performed to identify the correct level first at
midline and then moving laterally until the TPs are identified (Fig.
54.6). Rotating the transducer to the transverse/short-axis orientation
will bring into view the junction of the SAP and TP. A 22G or 25G, 3.5-
inch or 5-inch spinal needle is directed using an in-plane approach
aiming at the junction of the SAP and TP (Fig. 54.7). Once the bone
contact was made, the transducer should be rotated longitudinally and
shadows of TPs should be identified. Gentle agitation of the needle is
exercised and the tip must be seen at the cephalad aspect of the TP. One-
half milliliter of anesthetic is injected. The L5 dorsal ramus may present
a challenge secondary to the presence of a high iliac crest. If the iliac
crest does present a challenge, the block can be done utilizing an out-of-
plane approach, where the transducer is placed in short axis at L5–S1
and the needle is inserted and advanced in a caudocephalad direction
until contact is made at the desired target, the junction of the S1 SAP and
the sacral ala. In addition to the ultrasound guidance, nerve stimulation
using a 21G, 70-mm or 100-mm insulated needle may help to localize
the nerve. At 1 mA stimulation, the needle is advanced until contractions
of the lower multifidus muscles are seen.
Tips 1. Ultrasound-guided imaging may be particularly difficult in
patients with anatomical variations such as sacralized L5 vertebral
bodies, scoliosis, and obesity but can be done technically correct.
2. If medial branch nerve blocks provide relief of the patient’s pain for
the duration of time corresponding to the local anesthesia used,
radiofrequency ablation can be considered.

Suggested Readings Galiano K, Obwegeser AA, Walch C, et al.


Ultrasound-guided versus computed tomography-controlled facet
joint injections in the lumbar spine: a prospective randomized
clinical trial. Reg Anesth Pain Med. 2007;32:317–322.
Gofeld M. Ultrasound-guided lumbar zygapophysial joint and nerve block. In: Bigeleisen P, ed.
Ultrasound-Guided Regional Anesthesia and Pain Medicine: Techniques and Tips.
Philadelphia, PA: Lippincott Williams & Wilkins; 2014:240–251.
Gofeld M, Bristow S, Chiu S. Ultrasound-guided injection of lumbar zygapophyseal joints: an
anatomic study with fluoroscopy validation. Reg Anesth Pain Med. 2012;37:228–231.
Greher M, Kirchmair L, Enna B, et al. Ultrasound-guided lumbar facet nerve block. Accuracy of a
new technique confirmed by computed tomography. Anesthesiology. 2004;101:1195–1200.
Narouze S, Peng P. Ultrasound-guided interventional procedures in pain medicine: a review of
anatomy, sonoanatomy, and procedures. Part II: axial structures. Reg Anesth Pain Med.
2010;35:386–396.
Sato M, Simizu S, Kadota R, et al. Ultrasound and nerve stimulation-guided L5 nerve root block.
Spine. 2009;34:2669–2673.
Yun D, Kim H, Yoo S, et al. Efficacy of ultrasonography-guided injections in patients with facet
syndrome of the low lumbar spine. Ann Rehab Med. 2012;36:66–71.
55
Ultrasound-Guided
Botulinum for
Spasticity and
Applications in
Physiatry

TODD BEERY

Physiatrists have recognized the clinical use of ultrasound for several


decades. The use and relevance of ultrasound in the field is
demonstrated by a group of physiatrists establishing the American
Institute for Ultrasound in Medicine.1 The use of ultrasound has
expanded with improved technology and resolution. These
improvements have led to the implementation of ultrasound in the
detection of musculoskeletal and neuromuscular pathology along with
guidance for therapeutic procedures.2
The objective of this chapter is to briefly review some of the advantages
and disadvantages of ultrasound compared to other imaging modalities
and present the clinical applications of ultrasound in the field of
physiatry.
Ultrasound is an appealing modality for its convenience,
noninvasiveness, and portability. It is relatively inexpensive compared to
computed topography and magnetic resonance imaging (MRI). In
contrast to conventional radiography, ultrasound does not expose
patients to radiation. Ultrasound has no known contraindications and
results in minimal artifact from metallic objects.3
There are several limitations to ultrasound as an imaging modality. The
imaging window of ultrasound is relatively small. It provides detailed
soft-tissue imaging of a focal area; however, it can be tedious and
difficult to evaluate a widespread anatomical area. In addition,
ultrasound has limited resolution at greater depths, making visualization
of obese patients more challenging. Ultrasound lacks the ability to
penetrate bone, which can prevent views of deep body regions or central
intra-articular regions.4 Lastly, the use of ultrasound is operator-
dependent and can be technically demanding. It requires skill and
experience along with a solid knowledge base of anatomy.
What are some clinical indications for ultrasound in physiatry?
• A diagnosis of musculoskeletal pathology
• Interventional procedure guidance
• A diagnosis and treatment of residual limb pain in amputees
• To aid in the diagnosis of entrapment neuropathies and complement
nerve conduction studies and electromyography
• To assist in the treatment of targeted spasticity management in central
nervous system pathology
A diagnosis of musculoskeletal pathology
Ultrasound can serve as an initial imaging modality for several
musculoskeletal conditions. It can be performed quickly and
conveniently in the office setting. It is well tolerated by patients, and it
can be a useful tool in efficiently assessing musculoskeletal conditions.
Its use in the outpatient clinic has been associated with a decreased need
for office visits, improved clinical outcomes, and increased patient
satisfaction.5
Musculoskeletal ultrasound offers unique aspects that are absent in other
imaging modalities. A comparison of static and dynamic evaluations can
be performed. Real-time imaging can be observed during provocative
physical exam maneuvers. Ultrasound offers the ability for
sonopalpation of painful locations. The precise localization of the area of
maximal tenderness can be identified and visualized by palpating under
the transducer. A contralateral comparative exam of extremities can
easily be performed if indicated. The use of these techniques can aid in
the diagnosis of muscle, tendon, ligament, and nerve pathologies.6
Interventional procedure guidance
Ultrasound has become an invaluable tool for interventional procedures.
It facilitates direct visualization of the needle trajectory and final
placement. There are a wide variety of interventional procedures that
encompass the use of ultrasound. These include therapeutic injections to
joints, bursa, tendons, nerve sheaths, and muscles.7 The specifics of
injections techniques are discussed in other chapters. There is evidence
that ultrasound results in more accurate needle placement when
compared to anatomical landmarks for a variety of injection sites.8–13
However, there is not convincing evidence that ultrasound guidance
translates into improved clinical efficacy.14–16
A diagnosis and treatment of residual limb pain in amputees
Residual limb pain in amputees can be debilitating. The etiology of the
pain generator is vast and can be difficult to diagnose. Ultrasound can
effectively assess the potential causes of residual limb pain. Fluid
collections from adventitious bursae or neuromas can be identified with
ultrasound.17 In addition to diagnosing residual limb pain, ultrasound
also has a role in treatment. An ultrasound-guided neurolysis for
symptomatic neuromas can be performed to ensure sclerosing agents are
injected to the precise location.18
To aid in the diagnosis of entrapment neuropathies and complement nerve
conduction studies and electromyography
Ultrasound can be used as complimentary tool for nerve conduction
studies (NCSs) and electromyography (EMG). It can provide additional
structural information to aide in the neuromuscular evaluation.
Ultrasound of some peripheral nerves can provide better spatial
resolution than MRI, allowing for the identification of individual nerve
fascicles.19 In many instances, the use of ultrasound is unnecessary for
routine NCS and EMG studies; however, there are specific instances
where this imaging modality can be useful.
Ultrasound provides direct visualization of neurologic and vascular
structures that can be traced proximally and distally along upper and
lower extremities. This feature can assist in confirming entrapment
neuropathies when used in combination with conventional NCS. In
addition to localizing pathology, ultrasound imaging can lead to the
etiology of compressive neuropathies by visualizing fibrous fascial
bands causing entrapment.20
The mapping of peripheral nerves with ultrasound can assist in
identifying the course of nerves to ensure accurate and reliable NCS are
performed. This technique is useful in nerves that have greater anatomic
variability or are technically demanding. Examples of nerves that have
been studied include the lateral femoral cutaneous, sural, and saphenous
nerves.21–23 This technique is also advantageous in instances of
postsurgical changes when the conventional anatomical course of a
nerve has been altered, such as ulnar nerve transposition.24
EMG can be performed in combination with ultrasound. Similar to
NCSs, most routine studies can properly be conducted without the use of
ultrasound. Many muscles can be reliably identified and tested without
direct visualization. Ultrasound, however, can serve as an adjunct tool in
certain circumstances. Although most electromyographers are confident
in localizing selected muscles, some studies have shown that needle
placement is not as accurate as expected using palpation.25–27
Ultrasound can help in confirming needle placement in deeper muscles
or in obese patients where common bony landmarks are not easily
identified through palpation.28
Ultrasound provides an added safety measure to avoid vascular
structures during EMG studies. Power Doppler can identify vascular
structures, which is especially beneficial in patients who are
anticoagulated. Postprocedure scanning can also be performed for
surveillance if a hematoma is suspected.29,30
Ultrasound can assist in imaging for muscles that have close proximity
to vital organs. EMG of the diaphragm to assess phrenic nerve function
can be directly visualized with the use of ultrasound.31 Other muscles,
which would otherwise be difficult to localize without ultrasound,
include the thoracic paraspinals, serratus anterior, and rhomboids. In this
setting, ultrasound may prevent pneumothorax (Fig. 55.1).32,33 Lastly,
ultrasound can serve as confirmation for EMG needle placement for
patients who are uncooperative or who cannot provide voluntary
contraction in cases of severe denervation or spasticity.

To assist in the treatment of targeted spasticity management in central nervous


system pathology
Ultrasound assists in the management spasticity for pathology related to
spinal cord injury, traumatic brain injury, cerebral vascular accident,
multiple sclerosis, and cerebral palsy. It permits accurate injections of
botulinum toxin, allowing visualization of the targeted muscles. It may
prevent injecting vascular or neurologic structures, and allows for real-
time imagining of the injectate, ensuring it is given within the muscle
belly.34 Ultrasound guidance is equal to or superior to electric
stimulation or EMG for accuracy of antispasticity injectates.34 An
advantage of ultrasound relative to EMG when targeting muscles for
spasticity management is that ultrasound eliminates the component of
patient participation in the form of muscle recruitment activity to
confirm accurate needle placement, which may apply in patients who are
unable to comply with requests.
Conclusion
Musculoskeletal ultrasound provides detailed imaging of soft-tissue
structures and has broad applications in the diagnosis and treatment of
musculoskeletal and neuromuscular pathology. With proper training and
experience, ultrasound is an invaluable clinical tool for physiatrists.

References
1. Valente C, Wagner S. History of the American Institute of
Ultrasound in Medicine. J Ultrasound Med. 2005;24:131–142.
2. Ozackar L, Tok F, Kesikburun S, et al. Musculoskeletal sonography
in physical and rehabilitation and medicine: results of the first
worldwide survey study. Arch Phys Med Rehabil. 2010;91:326–
331.
3. Erickson SJ. High-resolution imaging of the musculoskeletal
system. Radiology. 1997;205:593–618.
4. Kremkau F. Diagnostic Ultrasound: Principles and Instruments. 6th
ed. Philadelphia , PA: WB Saunders; 2002:428.
5. Sivan M, Brown J, Brennan S, et al. A one-stop approach to the
management of soft tissue and degenerative musculoskeletal
conditions using clinic based ultrasonography. Musculoskeletal
Care. 2011;9:63–68.
6. Klauser AS, Tagliafico A, Allen GM, et al. Clinical indications for
musculoskeletal ultrasound: a Delphi-based consensus paper of the
European Society of Musculoskeletal Radiology. Eur Radiol.
2012;22:1140–1148.
7. Louis LJ. Musculoskeletal ultrasound intervention: principles and
advances. Radiol Clin North Am. 2008;46:515–533.
8. Hashiuchi T, Sakurai G, Morimoto M, et al. Accuracy of the biceps
tendon sheath injection: ultrasound-guided or unguided injection? A
randomized controlled trial. J Shoulder Elbow Surg. 2011;20:1069–
1073.
9. Curtiss HM, Finnoff JT, Peck E, et al. Accuracy of ultrasound-
guided and palpation-guided knee injections by an experienced and
less-experienced injector using a superolateral approach: a
cadaveric study. PM R. 2011;3:507–515.
10. Muir JJ, Curtiss HM, Hollman J, et al. The accuracy of ultrasound-
guided and palpation-guided peroneal tendon sheath injections. Am
J Phys Med Rehabil. 2011;90:564–571.
11. Berkoff DJ, Miller LE, Block JE. Clinical utility of ultrasound
guidance for intra-articular knee injections: a review. Clin Interv
Aging. 2012;7:89–95.
12. Gilliland CA, Salazar LD, Borchers JR. Ultrasound versus anatomic
guidance for intra-articular and periarticular injection: a systematic
review. Phys Sportsmed. 2011;39:121–131.
13. Davidson J, Jayaraman S. Guided interventions in musculoskeletal
ultrasound: what’s the evidence? Clin Radiol. 2011;66:140–152.
14. Soh E, Li W, Ong KO, et al. Image-guided versus blind
corticosteroid injections in adults with shoulder pain: a systematic
review. BMC Musculoskelet Disord. 2011;12:137.
15. Dogu B, Yucel SD, Sag SY, et al. Blind or ultrasound-guided
corticosteroid injections and short-term response in subacromial
impingement syndrome: a randomized, double-blind, prospective
study. Am J Phys Med Rehabil. 2012;91:658–665.
16. Elkousy H, Gartsman GM, Drake G, et al. Retrospective
comparison of freehand and ultrasound-guided shoulder steroid
injections. Orthopedics. 2011;34.
17. Van Geffen GJ, Bruhn J, Gielen MJ, et al. Pain relief in amputee
patients by ultrasound-guided nerve blocks. Eur J Anaesthesiol.
2008;25:424–425.
18. Lim KB, Kim YS, Kim JA. Sonographically guided alcohol
injection in painful stump neuroma. Ann Rehabil Med.
2012;36:404–408.
19. Rosenberg I. Echotexture of peripheral nerves: correlation between
US and histologic findings and criteria to differentiate tendons.
Radiology. 1995;197:291–296.
20. Sucher BM. Ultrasound imaging of the carpal tunnel during median
nerve compression. Curr Rev Musculoskelet Med. 2009;2:134–146.
21. Scheidegger O, Küffer AF, Kamm CP, et al. Reproducibility of
sensory nerve conduction studies of the sural nerve using
ultrasound-guided needle positioning. Muscle Nerve. 2011;44:873–
876.
22. Boon AJ, Bailey PW, Smith J, et al. Utility of ultrasound-guided
surface electrode placement in lateral femoral cutaneous nerve
conduction studies. Muscle Nerve. 2011;44:525–530.
23. Watson JC, Pingree MJ, Boon AJ, et al. A novel ultrasound-guided
proximal saphenous nerve conduction study. Muscle Nerve.
2009;40:731.
24. Boon AJ, Smith J, Harper CM. Ultrasound applications in
electrodiagnosis. PM R. 2012;4:37–49.
25. Haig AJ, Goodmurphy CW, Harris AR, et al. The accuracy of
needle placement in lower limb muscles: a blinded study. Arch Phys
Med Rehabil. 2003;84:877–882.
26. Chiodo A, Goodmurphy C, Haig A. Cadaver evaluation of EMG
needle insertion techniques used to target muscles of the thorax.
Spine (Phila, Pa 1976). 2006;31:E241–E243.
27. Chiodo A, Goodmurphy C, Haig A. Cadaveric study of methods for
subscapularis muscle needle insertion. Am J Phys Med Rehabil.
2005;84:662–665.
28. Boon AJ, Oney-Marlow TM, Murthy NS, et al. Accuracy of
electromyography needle placement in cadavers: non-guided vs.
ultrasound guided. Muscle Nerve. 2011;44:45–49.
29. Boon AJ, Gertken JT, Watson JC, et al. Hematoma risk after needle
electromyography. Muscle Nerve. 2012;45:9–12.
30. Lynch SL, Boon AJ, Smith J, et al. Complications of needle
electromyography: hematoma risk and correlation with
anticoagulation and antiplatelet therapy. Muscle Nerve.
2008;38:1225–1230.
31. Smith J, Finnoff JT. Diagnostic and interventional musculoskeletal
ultrasound: part 2. Clinical applications. PM R. 2009;1:162–177.
32. Honet JE, Honet JC, Cascade P. Pneumothorax after
electromyographic electrode insertion in the paracervical muscles:
case report and radio-graphic analysis. Arch Phys Med Rehabil.
1986;67:601–603.
33. Miller J. Pneumothorax. Complication of needle EMG of thoracic
wall. N J Med. 1990;87:653.
34. Alter KE. High-frequency ultrasound guidance for neurotoxin
injections. Phys Med Rehabil Clin N Am. 2010;21:607–630.
56
Ultrasound-Guided
Sacroiliac Joint
Injection
KACEY A. MONTGOMERY AND MICHAEL

GOFELD

Background and indications: The sacroiliac (SI) joint is another


common source of pain affecting patients presenting to chronic
outpatient pain clinics. These patients commonly complain of buttock
pain, possibly posterior thigh pain, and have a pinpoint tenderness
overlying the superior portion of the SI joint.

Anatomy: The SI joint is, as its name implies, composed of the junction
of the articular surfaces of the sacrum and ileum. The diarthrodial joint
is enclosed in a fibrous capsule (Fig. 56.1). It normally has minimal
mobility and is typically a stable joint secondary to extensive support by
surrounding muscles and fascia. Innervation to the posterior aspect of
the SI joint is primarily via the lateral branches of the L4–S2 nerve
roots, in addition to some contributions from the superior gluteal nerve
and S3 nerve roots. The anterior SI joint is innervated by L2–S2.
Patient position: Prone with towel rolls placed under the abdomen for
alleviation of lumbar lordosis.

Probe: Broadband low-frequency curved array transducer.

Probe position: The SI joint can be accessed from a superior or inferior


approach.
For the superior approach, the probe is placed in a transverse/axial
orientation at the level of the posterior superior iliac spine (PSIS).
For the inferior approach, the transducer is placed in a transverse/axial
orientation distally to PSIS at the level of second sacral foramina.

Needle: 22G 3.5-inch or 5-inch spinal needle.

Local anesthetic: 1.5 mL of choice of local anesthesia, such as 2%


lidocaine, mixed with 0.5 mL of methylprednisolone (80 mg/mL).

Technique: The skin is prepped and draped in usual sterile fashion,


ultrasound is covered in a sterile cover, and ultrasound gel is applied
liberally to the skin. For the superior approach, transverse/axial scanning
is undertaken first, visualizing the spinous process of the fifth lumbar
vertebra. Moving caudally, the sacrum will be identified. This is the
approximate level of the S1 foramen. Sliding the transducer laterally, the
posterior surface of the ileum will come into view with anechoic shadow
of the PSIS (Fig. 56.2). The needle is advanced using an in-plane
approach toward the junction between the PSIS and sacrum. Usually, a
resistance of entering firm iliosacral ligament is felt.

For the inferior approach to the SI joint, transverse/axial scanning is


undertaken from the level of PSIS until the bony protuberance cannot be
seen anymore. Where the sacrum and ileum meet corresponds with the
posterior, caudal portion of the SI joint. Typically, the S2 foramen as
well as the lateral sacral crest is seen medially to the joint opening. The
needle is then directed toward the target utilizing an in-plane approach
beginning at the medial aspect of the transducer (Fig. 56.3).
Tips

1. A limitation of ultrasound approach is the inability to identify an


intra-articular injection versus a periarticular injection.
2. The bony artifact tends to limit the ability to reliably detect an
intravascular injection.
Suggested Readings
Klauser A, De Zordo T, Feuchtner G, et al. Feasibility of ultrasound-guided sacroiliac joint
injection considering sonoanatomic landmarks at two different levels in cadavers and patients.
Arthritis Rheum. 2008;59:1618–1624.
Narouze S, Peng P. Ultrasound-guided interventional procedures in pain medicine: a review of
anatomy, sonoanatomy, and procedures. Part II: axial structures. Reg Anesth Pain Med.
2010;35:386–396.
Pekkafahli MZ, Kiralp MZ, Basekim CC, et al. Sacroiliac joint injections performed with
sonographic guidance. J Ultrasound Med. 2003;22:553–559.
57
Ultrasound-Guided
Piriformis Injection
KACEY A. MONTGOMERY AND MICHAEL

GOFELD

Background and indications: Piriformis syndrome is characterized as


primarily buttock pain with or without sciatica-type radiation that is
aggravated with abduction or internal rotation of the hip. As much as 6%
to 8% of patients with low back pain may have piriformis-related pain.
Deep palpation of the gluteal area as well as utilization of various
provocative maneuvers may help in the clinical diagnosis. One such test
involves eliciting pain with passive adduction and internal rotation of the
hip when it is flexed at 90 degrees.

Anatomy: The piriformis muscle serves to abduct and externally rotate


the femur. The flat muscle extends from the anterior aspect of the S2–S4
vertebrae, courses through the pelvis, exiting through the greater sciatic
foramen, and inserts medially at the greater trochanter. The muscle is
covered by the gluteus maximus (Fig. 57.1). Deeper to it, the sacral
plexus and part of the superior gemellus muscle are respectively located.
The inferior gluteal nerve and artery, the sciatic nerve, and the pudendal
nerves and artery typically pass distally the piriformis muscle. Although
less common, the sciatic nerve can divide proximal to the piriformis
muscle with a division travelling through the piriformis muscle and a
portion exiting distally. The intramuscular portion of the nerve may be
prone to irritation, leading to “sciatica”-type pain, or posterior lower
extremity radicular pain that may be described by some patients as
buttock pain.

Patient position: Prone with pillows placed for patient comfort.

Probe: Broadband low-frequency curved array transducer.

Probe position: Short axis/transverse.


Needle: 22G or 25G 3.5-inch or 5-inch spinal or 22G stimulating needle.
Drugs: Local anesthetic with or without corticosteroid (i.e., 2 mL of 1%
lidocaine and methylprednisolone 20 to 40 mg). Alternatively, 50 units
of botulinum toxin can be injected for more lasting results.

Technique: A sterile prep and drape per usual standard fashion including
a sterile ultrasound cover is performed. After the liberal application of
ultrasound gel, the posterior superior iliac spine is identified and the
transducer is then placed in a transverse orientation. While maintaining
an axial position, the transducer is moved inferiorly with the lateral
sacrum coming into view medially and the posterior inferior iliac spine
on the lateral side of the image. Further caudal scanning will result in the
disappearance of the ilium image as the transducer moves to the level of
the greater sciatic notch. At this point, the operator will see the gluteus
maximus superficially and the piriformis muscle deep to it. An assistant
can facilitate the identification of the piriformis muscle by taking the
patient’s leg, flexed at the knee, and internally and externally rotating the
thigh, observing the piriformis muscle moving in a typical pattern. At
this point, care should be taken to identify the underlying sacral plexus
or intramuscular portion of the sciatic nerve to prevent traversing it.
After the application of local anesthetic at the skin for superficial
anesthesia, a spinal or stimulating needle is then inserted using an inline
approach, passing first through the skin, subcutaneous tissue, and
gluteus maximus and entering the sheath of the piriformis muscle (Fig.
57.2). Electrical stimulation should elicit femur internal rotation and
abduction. The desired drug is then deposited just under the sheath of the
piriformis muscle under direct visualization.
Tips

1. Proper identification of the sacral plexus and sciatic nerve will


prevent deposition of local anesthesia onto them with resultant motor
and sensory deficits.
2. Intramuscular administration of botulinum toxin can be performed in
a similar fashion, as mentioned previously, by advancing the needle
into the belly of the piriformis muscle. A nerve stimulator may be
utilized to prevent the injection of toxin near the sciatic nerve.
Suggested Readings
Byrd JWT. Piriformis syndrome. Oper Tech Sports Med. 2005;13:71–79.
Hallin R. Sciatic pain and the piriformis muscle. Postgrad Med. 1983;74:69–72.
Peng PWH, Narouze S. Ultrasound-guided interventional procedures in pain medicine: a review of
anatomy, sonoanatomy, and procedures. Part I: nonaxial structures. Reg Anesth Pain Med.
2009;34:458–474.
Smith J, Hurdle MF, Locketz AJ, et al. Ultrasound-guided piriformis injection: technique
description and verification. Arch Phys Med Rehabil. 2006;87:1664–1667.
58
Ultrasound-Guided
Subacromial Bursa
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: The subacromial bursa injection is a


commonly performed procedure in orthopedic, musculoskeletal, and
pain practices. Surface landmarks provide inadequate appreciation of
anatomy, and the rate of inaccurately placed injections is between 13%
and 71%.1–3 A surface anatomy-based (“blind”) subacromial injection is
often targeted for the region just above the supraspinatus tendon.
Unfortunately, the needle often enters superior aspect of the
supraspinatus tendon, and the injectate must then rupture through the
rotator cuff capsule with subsequent overflow of medications into the
superior aspect of the rotator cuff and then spreading into bursa via a
ruptured rotator cuff capsule.4 Therefore, ultrasound guidance is highly
recommended to more accurately target the bursal space.

Anatomy: The subacromial bursa is a very narrow space that is noted just
above the supraspinatus muscle. It is a self-contained thin-walled
structure extending from the anterolateral aspect of the shoulder
(subdeltoid) to the subacromial space and the supraspinatus area. This
bursa is the largest in the human body and may accumulate up to 15 to
20 mL of fluid. However, the subacromial space is only a potential
space, normally represented by two membranous surfaces separated by a
thin film of lubricating fluid.

Transducer: High-frequency 25-to 50-mm linear transducer.

Transducer orientation: The transducer is placed in a position as in


Figure 58.1. The notch of the transducer should be pointed toward the
patient’s ear. Slight modifications in the angle may be needed based on
the patient’s anatomy and position.

Needle: 25G 2-inch needle.

Local anesthetic: 2 to 3 mL of 1% lidocaine and 40 mg of triamcinolone.

Patient position: The sitting position is preferred, which allows the


weight of the arm to pull the humeral head down. One can also have an
assistant gently pull the arm down to open the supraspinatus outlet. This
is rarely necessary.

Technique: A “double-syringe setup” is preferable where one 3 mL


syringe is filled with saline and the other 5 mL syringe filled with the
local anesthetic and steroid solution.
Injection of saline produces a “hydrodissection” of tissues and
verification of the needle tip position before instillation of corticosteroid
solution is attempted.
After a sterile skin preparation, the transducer is placed over the distal
acromion of the shoulder with the notch pointing toward the ear. Then
transducer is sliding over the distal acromion onto the shoulder. This
maneuver is necessary to obtain the “bird’s beak” view of the tendon
supraspinatus. The supraspinatus tendon will be seen emerging from
beneath the acromion and extending downward to the greater tuberosity.
Panning the transducer slightly anterior, no more than 0.5 cm, often
provides a better view of the bursa (Fig. 58.2). Normally, it will appear
as a thin anechoic fluid layer, which lies just above the supraspinatus
tendon. With slight adjustments to the transducer, one can also bring into
view the thin double hyperechoic lines of the capsule. In the case of
bursitis, an increase in the amount of fluid is seen without difficulties.

The needle is inserted in an in-plane view into the bursa (Fig. 58.3).
Once the needle enters the bursa, gentle hydrodilation is performed
while the needle tip is turned away from the rotator cuff. This will
protect the upper fibers of the supraspinatus tendon. An injection of a
local anesthetic and corticosteroid solution is done under ultrasound
observation to ensure the injectate is placed appropriately into the bursa
and not into the rotator cuff tendon.
References
1. Eustace J, Brophy D, Gibney R, et al. Comparison of the accuracy
of steroid placement with clinical outcome in patients with shoulder
syndromes. Ann Rheum Dis. 1997;56:59–63.
2. Esenyel C, Esenyel M, Yeslitepe R, et al. The correlation between
the accuracy of steroid injections and subsequent shoulder pain and
function in subacromial impingement syndrome. Acta Orthop
Traumatol Turc. 2003;37:41–45.
3. Naredo E, Cabero F, Beneyto P, et al. A randomized comparative
study of short term response to blind injection versus sonographic-
guided injection of local corticosteroids in patients with painful
shoulder. J Rheumatol. 2004;31:308–314.
4. Rutten MJ, Maresch BJ, Jager GJ, et al. Injection of the
subacromial-subdeltoid bursa: blind or ultrasound-guided? Acta
Orthopaedica. 2007;78:254–257.
59
Ultrasound-Guided
Bicipital Tendon
Sheath Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Bicipital tendinosis is a common clinical


problem. The biceps tendon is susceptible to repeated compression loads
with impingement mechanics.1 Furthermore, anywhere a tendon curves
around bone, there is increased compression forces that make a tendon
susceptible to degenerative changes. Because of the susceptibility to
tendinosis, and rupture of the biceps tendon, extra caution should be
taken to avoid an intratendinous injection with steroids. Ultrasound-
guided injections provide an accurate way to target the peritendinous
sheath of the biceps tendon. In a study on 30 patients, Hashiuchi et al.2
documented that 40% of blind injections that were directed to the biceps
tendon sheath resulted in injections inside the tendon, or away from the
target. Ultrasound-guided injections has been shown to provide a
statistically significant better outcome when compared to blind
injections.3 Of note, there is an ascending branch of the anterior
circumflex artery that typically runs laterally to the biceps tendon. Prior
to injection, it is helpful to apply the color Doppler and identify this
regional blood vessels so that one can avoid injury to the vessel or intra-
arterial injection. Some prefer to perform the injection routinely on the
medial aspect of the tendon in order to avoid this vessel (Fig. 59.1).
Anatomy: The long head of the biceps tendon originates as a bifurcated
attachment at the glenoid labrum. The tendon lies over the humeral head
anteriorly and runs between the greater tuberosity and lesser tuberosity
within the bicipital groove. The tendon is stabilized within the groove by
the transverse humeral ligament, which includes some extensions of the
subscapularis muscle (Figs. 59.1 and 59.2). The tendon sheath of the
long head communicates proximally with the glenohumeral joint
capsule. This communication with the glenohumeral joint is important
diagnostically because effusion within the joint can fill the tendon sheath
following a full-thickness rotator cuff tear. The color Doppler can often
help differentiate a local biceps synovitis from an effusion from the
glenohumeral joint.

Transducer: 25-to 50-mm high-frequency linear transducer.

Transducer orientation: Transverse to the biceps tendon (see Fig. 59.2).

Needle: 27G 1.5-inch, 25-g 0.5-inch Local anesthetic: 2 to 3 mL of


1% lidocaine or other suitable local anesthetic with 20 to 40 mg of
triamcinolone.
Patient position: Sitting or supine. In a supine position, the hand is
supinated and the humerus is externally rotated. By having the patient
flex the elbow to 90 degrees, an assistant can then begin to slowly
externally rotate the arm until the image of the bicipital groove is
optimized for visualization. This is an ideal position for injection.

Technique: The transducer is placed in the transverse position for a cross


sectional view of the tendon and tendon sheath (see Fig. 59.2). The
needle is inserted in an out-of-plane fashion. The needle tip can be easily
visualized and directed to a position medial to the tendon. Caution
should be taken to avoid contact with the tendon. A biceps tendon sheath
with no effusion only offers approximately 2 mm clearance to place the
needle within the sheath. An injection of local anesthetic and steroid
solution is then performed. If the injection is done laterally to the
tendon, one should exercise caution to avoid the circumflex humeral
artery, which can easily be identified using color Doppler. For this
reason, it is advisable to inject at the medial aspect of the tendon sheath.
References
1. Singaraju VM, Kang RW, Yanke AB, et al. Biceps tendinitis in
chronic rotator cuff tears: a histologic perspective. J Shoulder
Elbow Surg. 2008;17:898–904.
2. Hashiuchi T, Sakurai G, Morimoto M, et al. Accuracy of the biceps
tendon sheath injection: ultrasound-guided or unguided injection? A
randomized controlled trial. J Shoulder Elbow Surg. 2011;20:1069–
1073.
3. Zhang J, Ebraheim N, Lause GE. Ultrasound-guided injection for
the biceps brachii tendinitis: results and experience. Ultrasound
Med Biol. 2011;37:729–733.
60
Ultrasound-Guided
Glenohumeral Joint
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications The accuracy of blind glenohumeral


injections has been the subject of many studies.1–3 Ultrasound guidance
significantly improves the accuracy of the intra-articular injections. In
addition, ultrasound-guided glenohumeral injections can be less time-
consuming and more successful on the first attempt compared to
fluoroscopic guidance. More importantly, misplaced injections can result
in tendon injections (which can lead to a tendon weakening), skin
depigmentation, soft-tissue damage, and dissatisfaction. Several methods
of intra-articular injections have been described such as an anterior,
anterior rotator cuff interval, superior rotator cuff interval, and posterior
approach.4–6 If the injection is planned for the purpose of an arthrogram,
extra caution should be exercised to avoid the tissues of interest, such as
the tendon of the subscapularis. For this purpose, a rotator cuff interval
or posterior approach is typically preferred. Regardless of the approach,
it is important to avoid traversing the cartilaginous glenoid labrum.
Typically, musculoskeletal practitioners prefer a posterior approach,
although a modified rotator cuff interval approach can be used as an
alternative. Both approaches are described as follows.

Transducer: A linear high-frequency broad-band transducer (anterior


approach) or a curvilinear low-frequency transducer (posterior
approach).

Needle: 2 in to 3.5 in 22-gauge to 25-gauge spinal or echogenic needle.

Local Anesthetic: 5 mL of 1% ropivacaine, procaine, or other suitable


local anesthetic with 40 mg of triamcinolone.
Posterior approach Anatomy: The posterior approach is considered the
safest because no neurovascular structures are lying at the needle path. The
tendon infraspinatus extends over the glenohumeral joint to the posterior greater
tuberosity. A needle directed to the glenohumeral joint from the posterior
approach will traverse the tendon or muscle infraspinatus. The target for the
intra-articular posterior glenohumeral injection is a region overlying the humeral
head just distal to the glenoid labrum as noted by the asterisk in Figure 60.1.
One should avoid traversing the glenoid labrum for the intra-articular injection.
The bevel of the needle should be directed downward as contact is made over the
humeral head.

Patient Position: The patient is placed in the sitting position with the arm
across the lap (slight internal rotation) or alternatively in the scarf
position with the hand on the opposite shoulder (Fig. 60.2).
Transducer Orientation: A transducer is placed inferiorly and parallel to
the spine of scapula.

Technique: To visualize the posterior glenohumeral joint, position the


transducer just below and parallel to the spine of scapula. Slide the
transducer side to side until visualization of the humeral head and
glenoid labrum comes into view. The glenoid labrum will be seen as a
slightly hyperechoic, triangular-shaped structure between the glenoid
labrum and humeral head.
After skin anesthesia is performed either medially or laterally to the
transducer, affix a 25G spinal or 22G echogenic needle to the 3-mL
syringe filled with buffered lidocaine. The buffered lidocaine will be
injected in small aliquots intermittently as the needle is directed toward
the joint for the purpose of local anesthesia and navigation. The needle is
directed in an in-plane view toward the target, which is the posterior
humeral head just lateral to the labrum (Fig. 60.3). Again, avoid
traversing the labrum with a needle. Just prior to entering the
glenohumeral capsule, the bevel should be rotated toward the humeral
head to prevent damage of the hyaline cartilage. For larger shoulders, a
steeper approach angle may be required. In such a case, echogenic
needles with or without imaging enhancement will be beneficial. During
the injection, monitor the injectate to ensure that it is distending the joint
capsule and not flowing extra-articularly or expanding within the rotator
cuff tendon.

Modified rotator cuff interval (anterior) approach Anatomy: The rotator


cuff interval is a triangular space between the subscapularis and supraspinatus
tendons and the base of the coracoid process, covered by the rotator interval
capsule, whose main component is the coracohumeral ligament (CHL) (Fig.
60.4). The rotator cuff interval contains the long head of the biceps tendon
(LBT) and the superior glenohumeral ligament (SGHL).
This approach should be considered for obese patients when the
posterior access is deemed to be particularly difficult. In addition, there
is no risk of the labrum damage. The target is located next to the biceps
tendon at the superior aspect of the shoulder at the rotator cuff interval
where the needle traverses through the capsule and SGHL between the
subscapularis and biceps tendon.

Transducer Orientation: A linear transducer is placed over the shoulder in


the position as noted in Figure 60.5. The transducer notch is directed
toward the sternum. Adjust the transducer position to maximize
visualization of the biceps tendon.

Needle: 25G 2-inch or 27G 1.5-inch needle.

Local Anesthetic: 5 mL of local anesthetic with 40 mg of triamcinolone.

Patient Position: Patient is sitting or recumbent with palm supinated and


arm in neutral position or slight external rotation.

Injection Technique: The ultrasound image for this technique is obtained


by placing the transducer over the anterosuperior shoulder pointing
toward the sternum. Small adjustments of the transducer may be needed
to bring the biceps tendon into view. In such a way, the rotator cuff
interval view is optimized as discussed previously. The target is centered
in the middle of the viewing field. A local anesthetic is injected
subcutaneously above the transducer for the purpose of skin anesthesia
at the site corresponding to the center of the transducer. The needle is
directed in an out-of-plane view under the transducer and should be
watched carefully for the needle tip to appear in the visual field. The
needle is aimed to the region medial to the biceps tendon (Fig. 60.6).
The needle is advanced through the capsule and positioned adjacent to
the biceps tendon. One can inject saline or local anesthetic to and watch
for confirmation that the flow is entering the intra-articular space. Once
this is confirmed, the injection can be completed.
References
1. Balint P, Kane D, Hunter J, et al. Ultrasound guided versus
conventional joint and soft tissue fluid aspiration in rheumatology
practice: a pilot study. J Rheumatol. 2002;29:2209–2213.
2. Rutten M, Collins J, Maresch B, et al. Glenohumeral joint injection:
a comparative study of ultrasound and fluoroscopically guided
techniques before MR arthrography. Eur Radiol. 2009;19:722–730.
3. Demirhan M, Sonmez M, Kahraman S, et al. Accuracy of anterior
glenohumeral injections: a cadaver study. Arch Orthop Trauma
Surg. 2010;130:297–300.
4. Cicak N, Matasovic T, Bajraktarevic T. Ultrasonographic guidance
of needle placement for shoulder arthrography. J Ultrasound Med.
1992;11:135–137.
5. Valls R, Melloni P. Sonographic guidance of needle position for MR
arthrography of the shoulder. AJR Am J Roentgenol. 1997;169:845–
847.
6. Zwar R, Read J, Noakes J. Sonographically guided glenohumeral
joint injection. AJR Am J Roentgenol. 2004;183:48–50.
61
Ultrasound-Guided
Supraspinatus
Tendon Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background: Tears of the rotator cuff typically begin in the


supraspinatus tendon.1 A corticosteroid injection into the tendon may
cause a potential weakening of the tendon, and therefore, these
procedures are deemed inappropriate.
As new injection technologies and applications emerge such as platelet-
rich plasma, bone marrow aspirate, growth factor supplements, as well
as cell-and gene-modified cell therapies, musculoskeletal practitioners
will need to develop injection skills to deliver these new therapeutic
interventions via injection.2 Advances in both ultrasound technology as
well as emerging substances that augment tissue proliferation and
enhance healing are in its infancy. The use of percutaneous tendon
decalcification procedures (Fig. 61.1) will also require skills of needle
guidance into the rotator cuff tendons.3
Anatomy: The supraspinatus muscle originates in the supraspinatus fossa
extending under the acromion to attach in a triangular footprint on the
greater tuberosity. The muscle has a long tendinous portion in the
anterior half of the muscle, which typically inserts in the anterior-most
area of the highest impression on the greater tuberosity (Figs. 61.2 and
61.3).
The footprint of the muscle is triangular in shape and is smaller than
previously believed because a portion of it is actually belongs to the
tendon infraspinatus. The supraspinatus tendon is reinforced by fibers of
the coracohumeral ligament and fibers from adjacent tendons. Therefore,
lesions in the area of the supraspinatus tendon may involve one or more
of these elements, and the evolution of a tear in this area may be
determined by the location of the initial lesion. Microscopically, in the
region of the supraspinatus and infraspinatus tendons, the cuff is
composed of five layers defined by the attachments and orientations of
the fibrous elements in each of these layers.
Transducer: A broadband 25-to 50-mm linier transducer.
Transducer orientation: The transducer is placed over the lateral
acromion and directed so that the transducer is pointed at the patient’s
ear. Sliding the transducer laterally will obtain the bird’s beak view of the
tendon supraspinatus.

Needle: 25G 1.5-inch spinal or 22G echogenic needle. Larger bore


needles are often used for percutaneous decalcification procedures.
Local anesthetic: 1% lidocaine or other local anesthetic is used to
anesthetize the needle track. No local anesthetic injection is injected into
the tendon.
Patient position: Patient is in a seated position with the arm in a
modified Crass position shown in Figures 61.1 through 61.4.
Technique: After a sterile preparation of the skin and transducer head,
the transducer head is optimized into the position previously described
and the bird’s beak view of the tendon will become visible (see Fig.
61.1). The needle is inserted and directed toward the tendon
supraspinatus. The needle tip can be directed to the tendon periosteal
insertions at the footplate on the greater tuberosity as seen in Figure 61.4
as indicated for the treatment of insertion tendinosis and insertion site
tendon tears.
References
1. Codman EA. The Shoulder: Rupture of the Supraspinatus Tendon
and Other Lesions In or About the Subacromial Bursa. Boston:
Thomas Todd Co., 1934.
2. Nixon AJ, Watts AE, Schnabel LV. Cell-and gene-based approaches
to tendon regeneration. J Shoulder Elbow Surg. 2012;21:278–294.
3. Serafini G, Sconfienza L, Lacelli F, et al. Rotator cuff calcific
tendonitis: short-term and 10-year outcomes after two-needle us-
guided percutaneous treatment—nonrandomized controlled trial.
Radiology. 2009;252:157–164.
62
Ultrasound-Guided
Acromioclavicular
Joint Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Acromioclavicular (AC) joint


degeneration occurs early in life.1 The benefit of a local corticosteroid
injection is controversial. Symptom relief, however, has been reported
up to a year, and on occasion, pain relief can be even longer.2
Periarticular corticosteroid injections can lead to local complications,
including changes in skin overlying the joint. Ultrasound guidance may
improve the procedural accuracy. In a study on 80 cadavers, Borbas et
al.3 reported that 90% of injections were intra-articular compared to
70% when palpatory methods were used. In another large cadaveric
study, AC joint injections guided by ultrasound had 95% success versus
72% in the palpation group, but there was no significant difference in the
accuracy related to experience of the operators.4 It is not uncommon for
patients presenting with shoulder pain to have several sources of
nociception. It is important to recognize the contribution of the AC joint
in the overall presentation of patients with shoulder pain. The AC joint
may be only one of multiple targeted injections that may be necessary in
a patient presenting with shoulder pain. The location of pain arising
from the AC joint may be diverse and radiate into the deltoid region and,
on occasion, into the radial side of the forearm similar to problems
stemming from the rotator cuff and subdeltoid bursa.
Anatomy: The AC joint is formed by the articulation of the distal end of
the clavicle and the acromion process of the scapula (Fig. 62.1). The AC
joint is the junction of two flat bones that commonly degenerates early
in most individuals. The joint can be easily palpated in slender
individuals. Small osteophytes are commonly encountered. Bony
exostosis from the inferior AC joint is a common source of impingement
of the supraspinatus tendon. The AC joint can be narrowed or shielded
by overhanging osteophytes, and an injection may be technically
challenging.
Transducer: A broadband 25-mm linear transducer.
Transducer orientation: Long axis to the clavicle and acromion.

Needle: 27G 1.5-inch needle or 25G 1.5-inch needle.


Local anesthetic: 0.5 mL of local anesthetic with 20 to 40 mg of
triamcinolone.
Patient position: The patient is sitting. Typically, an external arm
rotation can assist in opening the joint. One can also have an assistant
gently traction the arm down, which also can open the joint.
Technique: The AC joint is visualized by placing a linear transducer
longitudinally along the clavicle and sliding the transducer head over the
joint (Fig. 62.2). The hyperechoic periosteum of the distal clavicle and
acromion can be easily visualized, distinguishing the joint margin. A
small hyperechoic fibrocartilaginous disk may be visualized within the
joint space. The joint capsule is visualized as a semilunar hyperechoic
line extending from the clavicle to the acromion above the joint. A
needle is typically advanced in an out-of-plane view. The needle tip
should be seen as a small hyperechoic dot within the joint. By gently
moving the needle, its tip can be seen flashing on and off, helping to
verify position. To ensure intra-articular injection, 0.1 to 0.2 mL of
normal saline that is shaken to create microbubbles can be injected,
which will act as an intra-articular contrast to confirm correct placement
and injection within the AC joint. Once the needle tip is within the joint,
the intra-articular space injection of 1 mL of a local anesthetic and
corticosteroid solution can be performed.
Alternatively, the injection may be performed utilizing an in-plane
technique, a small ultrasound transducer, and an acoustic offset method
(i.e., application of copious amount of sterile gel and placing the needle
through the gel, skin, and joint capsule).
References
1. Buttaci CJ, Stitik TP, Yonclas PP, et al. Osteoarthritis of the
acromioclavicular joint: a review of anatomy, biomechanics,
diagnosis, and treatment. Am J Phys Med Rehabil. 2004;83:791–
797.
2. Hossain S, Jacobs LG, Hashmi R. The long-term effectiveness of
steroid injections in primary acromioclavicular joint arthritis: a
five-year prospective study. J Shoulder Elbow Surg. 2008;17:535–
538.
3. Borbas P, Kraus T, Clement H, et al. The influence of ultrasound
guidance in the rate of success of acromioclavicular joint injection:
an experimental study on human cadavers. J Shoulder Elbow Surg.
2012:1–4.
4. Sabeti-Aschraf M, Lemmerhofer B, Lang S, et al. Ultrasound
guidance improves the accuracy of the acromioclavicular joint
infiltration: a prospective randomized study. Knee Surg Sports
Traumatol Arthrosc. 2011;19:292–295.
63
Ultrasound-Guided
Sternoclavicular Joint
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: The sternoclavicular (SC) joint is the


diarthrodial articulation between the axial and appendicular skeletons. It
is subject to the same disease processes that occur in other joints,
including degenerative arthritis, rheumatoid arthritis, infection, and
subluxation. Most of these conditions present with swelling of the joint,
which may be associated with pain and/or tenderness. A few studies
have been published addressing therapeutic injections of the SC joint.
Indications for injections of the SC joint at present are done empirically
for reasons similar to other joint injections. SC joint pain is relatively
uncommon compared to other sources of shoulder/pectoral girdle pain.
Anatomy: The SC joint is formed by the articulation of the proximal end
of the clavicle with the clavicular fossa in the superior lateral aspect of
the sternum (Fig. 63.1). The proximal end of the clavicle is slightly
elevated, and thus can easily be palpated at the SC joint. With scapular
retraction, the sternal side of the joint becomes more easily palpated.
The SC joint is rather shallow, and caution should be taken not to
overpenetrate the joint to prevent traversing the joint and entering
underlying structures. The SC joint is an atypical synovial joint; like the
acromioclavicular joint, the articular surfaces are covered with
fibrocartilage. The medial end of the clavicle articulates with the
manubrium and first costal cartilage. The capsule is thickened anteriorly
and posteriorly to form the SC ligaments.

Transducer: A broadband 25-mm linier transducer.


Transducer orientation: Center the middle of the transducer head
longitudinally over the joint with the probe longitudinal to joint, angled
at 45 degrees to midline (Fig. 63.2).
Needle: 27G 1.5-inch or 25G 1.5-inch needle.
Local anesthetic: 0.5 mL of 1% lidocaine or other suitable local
anesthetic with 20 mg of triamcinolone.
Patient position: Supine.
Technique: After a sterile preparation, the center of the transducer is
placed within the middle of the joint and the transducer head in-line with
the clavicle. The joint will appear as a notch with the clavicle end
projecting more anterior than the sternum. The SC joint has a small thin
articular capsule that can be distended in the case of effusion (Fig. 63.2).
The joint space is lined up to the center of the screen. The needle is
placed in an out-of-plane technique. The needle tip will appear as a
small hyperechoic dot in the field of vision. The needle tip is directed
into the center of the joint. To confirm intra-articular placement of the
needle, the needle can be rotated back and forth approximately 20 to 30
degrees to reflect the ultrasound beam off the tip of the needle bevel.
Another option is to inject a small amount of normal saline that is first
shaken to make microbubbles within the saline, which will act as a
contrast solution that will be clearly visible under ultrasound. Once the
needle tip is within the joint confirming intra-articular placement, 1 mL
of local anesthetic and corticosteroid is injected into the joint.
Bibliography
Klauser A, Tagliafico A, Allen G, et al. Clinical indications for musculoskeletal ultrasound: a
Delphi-based consensus paper of the European society of musculoskeletal radiology. Eur
Radiol. 2012;22:1140–1148.
Peterson C, Hodler J. Evidence-based radiology (part 2): is there sufficient research to support the
use of therapeutic injections into the peripheral joints? Skeletal Radiol. 2010;39:11–18.
64
Ultrasound-Guided
Intra-articular Elbow
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Lateral and medial epicondylitis is the


most common reason for chronic elbow pain. Other causes, such as joint
synovitis or arthritis, are also possible. These patients may benefit from
an intra-articular corticosteroid injection. A simplistic method for
performing an intra-articular injection of the elbow is to place the needle
between the radial head and the capitulum, which is easily visualized
under ultrasound. The use of musculoskeletal ultrasound can be utilized
to detect joint synovitis when applying color Doppler as well as to detect
fluid, septic effusion, etc.
Anatomy: The joint capsule invests the entire elbow joint. Inferiorly, it is
attached to the humeral shaft just above the coronoid and radial fossae,
to the anterior aspect of the coronoid process and to the annular
ligament. Because of this anatomical relationship, intra-articular
injections of the elbow joint can be accomplished by injecting the radial
capitulum articulation, which will then communicate to the rest of the
intra-articular space of the elbow. This is the preferable approach for
intra-articular injections of the elbow (Fig. 64.1).
Transducer: Small 25-mm linear transducer.
Transducer orientation: Longitudinal in a line with the radius.

Needle: 25G to 27G, 1-inch to 1.5-inch Local anesthetic: 2 mL of


local anesthetic with 40 mg of triamcinolone.
Patient position: Sitting or lateral decubitus with the elbow bent at 90
degrees.
Technique: After a sterile preparation, the transducer is placed over the
radiocapitellar articulation and the transducer head is kept along the long
axis to the radius. The joint should be lined up to the middle of the
transducer. The transducer is placed over the superior lateral aspect of
the joint in order to have access to the lateral joint for the injection (see
Fig. 64.1). The injection is performed in an out-of-plane approach. After
the intra-articular position is confirmed, a mixture of local anesthetic and
corticosteroid is injected. Typically, 10 to 20 mg of triamcinolone is
utilized.
Lateral epicondyle injection: extensor carpi radialis brevis Background and
Indications: The most common tendon involved in lateral epicondylitis is the
extensor carpi radialis brevis (ECRB) tendon at its origin on the lateral
epicondyles. Lateral epicondylitis is typically an angiofibroblastic tendinosis,
which is why it can become recalcitrant to treatment. There are emerging
technologies of ultrasound-guided intratendon injections for managing this
condition.
It is not uncommon for the extensor digitorum communis to contribute
to lateral epicondylitis. Ultrasound findings of tendinopathy may include
tendon thickening, hypoechogenicity, fibular disruption, and
intratendinous calcifications as well as adjacent bone irregularity. In
addition to the injection techniques, some practitioners perform a
microtendonotomy with the needle under ultrasound guidance to
enhance healing.
Anatomy: The common extensor tendon is a flat tendon that originates
from the anterolateral surface of the lateral epicondyle. It receives
contributions from four superficial extensor muscles, including the
ECRB, the extensor digitorum communis (EDC), the extensor digiti
minimi (EDM), and the extensor carpi ulnaris (ECU). ECRB makes the
most deep articular fibers, whereas the EDC contributes to more of the
superficial portion of the common extensor tendon. Note the upper
(supracondylar), mid (upper epicondyle), and lower (lower epicondyle)
relationship of the extensor carpi radialis longus (ECRL), ECRB, and
EDC tendons (Fig. 64.2). These are common targets for injection
therapies.
Transducer: Hockey-stick or 25-mm linear transducer.
Transducer Orientation: Transducer is placed over the radial capitulum
joint and slid up or more proximal until the epicondyle and the tendon
attachments of the ECRB come into full view. Then the distal portion of
the transducer is rotated toward the thumb of the patient.

Needle: 27G 1.5-inch or 25G 1.5-inch needle.


Local Anesthetic: 1% lidocaine.
Patient Position: Patient is sitting with the hands folded in the lap and
with the elbow bent. Alternatively, the patient can be supine with the
arm abducted and the forearm flexed in the prone position.
Technique: Following a sterile preparation of the skin, the transducer is
placed over the radius and then panned upward toward the epicondyle
orienting the transducer toward the thumb of the patient. This will bring
into view the epicondyles and the extensor tendon attachments.
The needle is inserted and directed in an in-plane view toward the lateral
epicondyle (see Fig. 64.2). If platelet-rich plasma (PRP) is administered,
an infiltration of the tendon attachments with local anesthetic (lidocaine)
should be done prior to the PRP injection.
Bibliography
Bianchi S, Martinoli C. Chapter 8: elbow. In: Ultrasound of the Musculoskeletal System. Berlin,
Germany: Springer-Verlag;2007:524–525.
Connell D, Burke F, Coombes P, et al. Sonographic examination of lateral epicondylitis. AJR Am J
Roentgenol. 2001;176:1763–1777.
Fairbank SR, Corelett RJ. The role of the extensor digitorum communis muscle in lateral
epicondylitis. J Hand Surg Br. 2002;27:405–409.
Ferrara M, Marcelis S. Ultrasound of the elbow. J Belge Radiol. 1997;80:122–123.
Greenbaum B, Itamura J, Vangsness CT, et al. Extensor carpi radialis brevis. An anatomical
analysis of its origin. J Bone Joint Surg Br. 1999;81:926–929.
Housner JA, Jacobson JA, Misko R. Sonographically guided percutaneous needle tenotomy for the
treatment of chronic tendinosis. J Ultrasound Med. 2009;28:1187–1192.
Kapoor S. Pain management of chronic lateral epicondylitis: emerging new therapeutic options.
Pain Med. 2012;13:848.
Laurence LP, McShane JM. Lateral epicondylitis of the elbow: US findings. Radiology.
2005;237:230–234.
Lin J, Jacobson JA, Fessell DP, et al. An illustrated tutorial of musculoskeletal sonography. II.
Upper extremity. AJR Am J Roentgenol. 2000;175:1071–1079.
Nirschl RP. Muscle and Tendon Trauma: Tennis Elbow. Philadelphia, PA: WB Saunders; 1985.
Poltawski L, Ali S, Jayaram V, Watson T. Reliability of sonographic assessment of tendinopathy in
tennis elbow. Skeletal Radiol. 2012;41:83–89.
Regan W, Wold LE, Coonrad R, et al. Microscopic histopathology of chronic refractory lateral
epicondylitis. Am J Sports Med. 1992;20:746–749.
65
Ultrasound-Guided
Wrist Injections
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Wrist ultrasonography not only provides a


means to evaluate anatomical structures of a joint and to facilitate
injections but also helps to detect an active hyperemia associated with
synovitis. Therefore, it may be a useful tool to monitor disease activity.
That is why ultrasound is becoming more popular in rheumatologic
practice. Ultrasound guidance improves accuracy of the radiocarpal
injections. Ultrasound-guided contrast injection in radiocarpal magnetic
resonance (MR) arthrograms has been shown to be a cost-effective
method that spares unnecessary radiation exposure. It is also useful in
evaluating fluid collection at the radiocarpal joint.
Anatomy: There are three proximal carpal bones that have important
joint function; they are the scaphoid, the lunate, and the triquetrum. The
capsule communicates through the whole proximal joint as seen in
Figure 65.1. The distal end of the radius is concave with articular facets
contributing to the radiocarpal joint. This proximal concave surface of
the joint is contiguous with the triangular fibrocartilage complex, which
extends from the ulnar styloid to the distal radius. The distal surface of
the joint is the articular surfaces of the scaphoid, the lunate, and the
triquetrum. The capsule of the radiocarpal joint is attached to the distal
radius and ulna and extends to the proximal carpal row. This joint is
reinforced by extrinsic carpal ligaments.
Transducer: Hockey-stick or 25-mm linear transducer.
Transducer orientation: The transducer is placed in the long axis over
the radiocarpal joint. In this view, a slight angulation of the joint surface
is noted (Fig. 65.2), which is important anatomically because a needle
guided into the joint will need to have a slight angulation to be properly
aligned to the articular surface. The lateral joint space of the
radioscaphoid joint below the base of the thumb has a fat pad that, if
inadvertently injected, may lead to an atrophy causing cosmetic
deformity. Therefore, the joint is entered from the dorsal aspect of the
wrist.
Needle: 25G or 27G 1.5-inch needle.
Local anesthetic: 2 mL of 1% lidocaine or other suitable local anesthetic
with 10 to 20 mg of triamcinolone.
Patient position: Sitting with the arm placed on the hand-table or supine
with pronated hand.
Technique: After a sterile preparation, the transducer is placed
longitudinally over the wrist overlying the radiocarpal joint (Fig. 65.3).
The transducer is adjusted by scanning side to side in order to center the
radiocarpal joint in the middle of the viewing field. Before the injection,
a careful palpation of the extensor tendons overlying the wrist should be
performed to plan the location of the needle insertion and to avoid
traversing the tendons. Note the approximate 40-degree angle that is
required to enter the radiocarpal joint. Although an in-plane injection
can be carried out, an out-of-plane technique is less cumbersome. This
provides a quicker and more reliable approach into the joint. The needle
is inserted at an angle of approximately 30 to 40 degrees into the joint
and under the transducer head (see Figs. 65.2 and 65.3). To confirm the
location of the needle tip, two techniques can be utilized: a rotation of
the needle or injecting saline. Once the needle position is confirmed to
be within the joint space, the corticosteroid solution can then be injected.
Suggested Readings
Bianchi S, Martinoli C. Chapter 10: wrist. In: Ultrasound of the Musculoskeletal System. Medical
Radiology. Berlin, Germany: Springer-Verlag;2007.
Cunnington J, Marshall N, Hide G, et al. A randomized, double-blind, controlled study of
ultrasound-guided corticosteroid injection into the joint of patients with inflammatory arthritis.
Arthritis Rheum. 2010;62:1862–1869.
Koga H, Engebretsen L, Brinchmann JE, et al. Mesenchymal stem cell-based therapy for cartilage
repair: a review. Knee Surg Sports Traumatol Arthrosc. 2009;17:1128–1197.
Lohman M, Vasenius J, Nieminen O. Ultrasound guidance for puncture and injection in the
radiocarpal joint. Acta Radiol. 2007;48:744–747.
66
Ultrasound-Guided
Ulnar Triquetral
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Ultrasonography has proven valuable in


evaluating inflammatory arthropathies, joint effusion, tendinopathy, and
the ligaments of the wrist as well as the triangular fibrocartilage
complex, (TFCC). Detailed techniques of ultrasound examinations of the
intrinsic and extrinsic ligaments and the TFCC at the ulnar triquetral
joint has been described.1 The ulnar triquetral region of the wrist is both
anatomically and biomechanically complex, and pain overlying this
region should be carefully evaluated to elicit a specific diagnosis. For
example, an excessively long ulnar styloid can impinge upon TFCC.2
Anatomically, the tip of the ulnar styloid is covered by a meniscus
homologue. Injuries are a common cause of ulnar-sided wrist pain and
can frequently lead to functional disability. Before considering injections
in this region of the wrist, one should consider carefully a target tissue or
diagnosis and the most appropriate treatment.

Anatomy: The intrinsic ligaments are situated entirely within the carpus,
between the carpal bones. The ligaments are divided into two major
categories, namely, the intraosseous ligaments, which lie between the
carpal bones, and the intrinsic capsular ligaments attaching to the carpal
bones. In addition, there are two other important intraosseous ligaments
proximally, which is the scapholunate ligament (SLL) and the
lunotriquetral ligament (LTL) (Fig. 66.1). Note the shape of the distal
end of the radius and ulnar and its formation of an articular surface for
the for the radioscaphoid, radiolunate, and ulnar-triquetrum articulation.

Transducer: Hockey-stick or 25-mm linear transducer.

Transducer orientation: Transducer placed longitudinally along the


ulnar-triquetrum articulation in a similar fashion to previous
radioscaphoid.

Needle: 27G 1.5-inch or 30G 1-inch needle.

Local anesthetic: 1 mL of 1% lidocaine or other suitable local anesthetic


with 10 to 20 mg of triamcinolone.

Patient position: Sitting with the arm placed on the hand-table or supine
with pronated hand.
Technique: After sterile preparation, the transducer is placed
longitudinally over the dorsal aspect of the ulnar-triquetrum articulation.
A slight wrist flexion makes the surface of the dorsal wrist flatten and
maintain good contact with the transducer. The transducer is held in a
position where the ulnar-triquetrum joint is centered in the visual field. A
rounded extensor carpi ulnaris (ECU) tendon at the dorsal lateral ulna
can be easily palpated. Panning the transducer distally will expose the
most dorsal aspect of the joint (Fig. 66.2). The needle is inserted in an
out-of-plane technique just to the radial side of the tendon that extends
above the ulnar styloid. The needle tip will appear as a hyperechoic dot
under the transducer. Once the needle tip is confirmed to be within the
intra-articular space, the injection is performed.
References
1. Taljanovic M, Goldberg M, Sheppard J, et al. US of the intrinsic
and extrinsic wrist ligaments and triangular fibrocartilage complex
—normal anatomy and imaging technique. Radiographics.
2011;31:e44.
2. Zahiri H, Zahiri C, Ravari F. The tip of an excessively long ulnar
styloid can impinge upon the triangular fibrocartilage complex
(TFCC) against the triquetrum. Ulnar styloid impingement
syndrome. Int Orthop. 2010;34:1233–1237.
67
Ultrasound-Guided
Carpal Tunnel
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Carpal tunnel syndrome (CTS) is the most


frequent entrapment mononeuropathy due to the compression of the
median nerve at the wrist. The clinical examination, consisting of a
history, physical examination, and provocative tests, has been considered
sufficient for CTS diagnosis. Ultrasound has been extensively studied as
an alternative tool to diagnose CTS.1 Ultrasonography is operator-
dependent, but shows high reproducibility after adequate training. It is
not an alternative diagnostic tool to electrodiagnostic tests and vice versa
in CTS, but they are complementary; one provides anatomic information
of the nerve and its surrounding structures, whereas the other provides
information on the level of the lesion and the function of the nerve fibers
with the largest diameters.2
Cochrane data shows that local corticosteroid injections for CTS provide
symptom relief for 1 month.3 Symptom relief beyond 1 month compared
to placebo has not been demonstrated. Further research is required to
determine the length of benefit of a local corticosteroid injection and the
benefit for mild and moderate CTS. The corticosteroid injection should
be only a component of the comprehensive management of CTS.
Anatomy: The contents of the carpal tunnel include two sets of four
tendons (the flexor digitorum superficialis and the flexor digitorum
profundus), one single tendon (the flexor pollicis longus), and one nerve
(the median nerve) (Fig. 67.1).

The median nerve is most superficial, found just beneath the flexor
retinaculum. The location of the median nerve can vary, and it can be
found more medial; thus, ultrasound identification and guided carpal
tunnel injections around the median nerve places ultrasound-guided
injections at a significant advantage to blinded and other methods. The
median nerve is also quite mobile and moves with finger flexion and
extension, which can be noted on dynamic ultrasound examination. The
tendons within the carpal tunnel may play an important role in median
nerve compression and the production of symptoms in CTS.
Management of the tendinopathy and considerations on the finger flexor
tendon may be important considerations in the management of CTS (see
Fig. 67.1).

Transducer: Hockey-stick or broadband 25-mm linear transducer.

Transducer orientation: Transducer placed transversely over the wrist.

Needle: 27G 1.5-inch or 25G 1.5-inch needle.

Local anesthetic: 2 mL of 1% lidocaine or other suitable local anesthetic


with 2 to 40 mg of triamcinolone.

Patient position: Recumbent with supinated hand.

Technique: With the hand and fingers relaxed, the transducer is placed
transversely over the wrist (Fig. 67.2). The median nerve is identified.
Color Doppler is applied to localize the ulnar artery. In 20% of the
population, the median artery is seen. In such case, the median nerve is
often bifurcated. The needle is inserted from the ulnar side in an in-plane
view and directed under in a shallow angle toward the median nerve (see
Fig. 67.2). An injection of local anesthetic and a corticosteroid solution
can be performed initially adjacent to the palmaris longus tendon. The
needle is then pulled back and placed underneath the nerve with the
bevel up toward the nerve allowing the solution to infiltrate underneath
and hydrodissection to occur. Caution must be taken to avoid contact
with the tendons as the needle is passed toward the median nerve. To
avoid the inconvenience of numbness in the median distribution
following the injection, the normal saline can be mixed with a
corticosteroid solution.
References
1. Beekman R, Visser LH. Sonography in the diagnosis of carpal
tunnel syndrome: a critical review of the literature. Muscle Nerve.
2003;27:26–33.
2. Mondelli M, Filippou G, Galo A, et al. Diagnostic utility of
ultrasonography versus nerve conduction studies in mild carpal
tunnel syndrome. Arthritis Rheum. 2008;59:357–366.
3. Marshall N, Tardif G, Ashworth N. Local corticosteroid injection
for carpal tunnel syndrome. Cochrane Database Syst Rev. 2002;
(4):CD001554.
68
Ultrasound-Guided
Intraarticular Hip
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: The hip is one of the deepest joints, and
ultrasound-guided injections may be challenging. Nevertheless,
ultrasound has an advantage over fluoroscopy in performing
intraarticular hip injections. The need for intraarticular access may be for
the purpose of arthrograms, corticosteroid injections, viscoelastic
supplementation and, in the future, cellular and other biologic therapies.
Intraarticular hip injections may also be necessary to rule out hip pain
from other potential sources of pain such as the spine, the sacroiliac
joint, and other soft tissues. Injections based on surface anatomy are
inaccurate in 20% to 50% of cases, and needles dangerously pass within
4 to 5 mm of the femoral nerve.1
The accuracy of ultrasound-guided hip injections has been evaluated.
Pourbagher et al.2 performed 30 sonographic-guided hyaluronidate
injections with contrast-enhanced computed tomography (CT)
documentation and demonstrated accuracy in all injections. Smith et al.3
reported ultrasound-guided arthrography with an average procedure time
of 112 seconds and a 97% rate of accuracy.

Anatomy: The hip joint is a ball and socket joint consisting of the
femoral head and acetabulum. The acetabulum is surrounded by the
triangular fibrocartilaginous labrum, which is clearly visible on
ultrasound exam (Fig. 68.1). The femoral head is a rounded structure
that is two-thirds of a sphere covered by hyaline cartilage except for the
region at the fovea. The anterior synovial recess extending to the neck of
femur is the most accessible target for ultrasound-guided injections (Fig.
68.2). The neurovascular bundle of the femoral vein, nerve, and artery
lies medially to the hip joint. The articular capsule is a strong and dense
fibrous structure that thickens closer to the femoral head and tapers as it
extends over the femoral neck and attaches to the intertrochanteric
region. The iliofemoral ligament forms the superior band and an inferior
band, which together is called the Y ligament Bigelow. This ligament
reinforces the anterior aspect of the capsule.4
Transducer: 3 to 5 MHz curvilinear transducer. A 7.5 to 15 MHz linear
transducer may be used in thinner patients.

Transducer orientation: Longitudinal to the femoral head and neck.

Needle: 22G 3.5-inch spinal or 25G 3.5-inch spinal.

Local anesthetic: 1% Lidocaine with or without corticosteroid.

Technique: The patient is placed in a supine position with a pillow under


the knee to slightly relax the anterior capsule. The neurovascular bundle
is identified on ultrasound and the skin is marked. The skin and the
ultrasound transducer are then sterile prepped. The transducer is placed
transversely just inferior to the inguinal crease. The curved surface of
the femoral head is identified. At this point, the transducer is slowly
rotated counterclockwise while keeping the head of the femur in view.
By adjusting or sliding the transducer position further medial, the head
of the femur, the femoral neck, and the anterior synovial recess are seen
as in Figures 68.2 and 68.3.
The injection target is the anterior synovial recess close to the osseous
junction between the head and the neck (see Fig. 68.3). Note the
hyperechoic iliofemoral ligament and hip capsule as well as the
overlying hypoechoic articular cartilage over the femoral head. In
addition, color Doppler is activated to rule out the circumflex artery that
may be located at the needle trajectory. A block needle is inserted at the
distal end of the ultrasound transducer in an in-plane fashion, and it is
advanced into the anterior lateral recess. The needle is traversed through
the articular capsule. An injectant should be seen as spreading
proximally along the bone surface into the joint filling the articular
capsule and lifting it slightly.
References
1. Leopold SS, Battista V, Oliverio JA. Safety and efficacy of
intraarticular hip injection using anatomic landmarks. Clin Orthop
Relat Res. 2001;391:192–197.
2. Pourbagher MA, Ozalay M, Pourbagher A. Accuracy and outcome
of sonographically guided intraarticular sodium hyaluronate
injections in patients with osteoarthritis of the hip. J Ultrasound
Med. 2005;24:1391–1395.
3. Smith J, Hurdle MF, Weingarten TN. Accuracy of sonographically
guided intraarticular injections in the native adult hip. J Ultrasound
Med. 2009;28:329–335.
4. Wagner FV, Negrão JR, Campos J et al. Capsular Ligaments of the
Hip: Anatomic, Histologic, and Positional Study in Cadaveric
Specimens with MR Arthrography. Radiology. 2012;263(1):189–
198.
Bibliography
Battaglia M, Guaraldi F, Vannini F, et al. Platelet-rich plasma (PRP) intraarticular ultrasound-
guided injections as a possible treatment for hip osteoarthritis: a pilot study. Clin Exp
Rheumatol. 2011;29:754.
Carson B, Wong A. Ultrasonographic guidance for injections of local steroids in the native hip. J
Ultrasound Med. 1999;18:159–160.
Choudur HN, Ellins ML. Ultrasound-guided gadolinium joint injections for magnetic resonance
arthrography. J Clin Ultrasound. 2011;39:6–11.
Galle J, Bader A, Hepp P, et al. Mesenchymal stem cells in cartilage repair: state of the art and
methods to monitor cell growth, differentiation and cartilage regeneration. Curr Med Chem.
2010;17:2274–2291.
Koga H, Engebretsen L, Brinchmann JE, et al. Mesenchymal stem cell-based therapy for cartilage
repair: a review. Knee Surg Sports Traumatol Arthrosc. 2009;17:1128–1197.
Masala S, Fiori R, Bartolucci DA, et al. Diagnostic and therapeutic joint injections. Semin
Intervent Radiol. 2010;27:160–171.
Migliore A, Tormenta S, Massafra U, et al. Repeated ultrasound-guided intraarticular injections of
40 mg of Hyalgan may be useful in symptomatic relief of hip osteoarthritis. Osteoarthritis
Cartilage. 2005;13:1126–1127.
Nöth U, Steinert AF, Tuan RS. Technology insight: adult mesenchymal stem cells for osteoarthritis
therapy. Nat Clin Pract Rheumatol. 2008;4:371–380.
Sánchez M, Guadilla J, Fiz N, et al. Ultrasound-guided platelet-rich plasma injections for the
treatment of osteoarthritis of the hip. Rheumatology (Oxford). 2012;51:144–150.
69
Ultrasound-Guided
Patellar Tendon
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Tendinosis is characterized histologically


by tissue degeneration with failed reparative response and an absence of
inflammatory cells.1,2 In the patellar tendon, it is referred to as jumpers
knee.3,4 Tendinosis is associated with a relative expansion of tendinous
tissue, longitudinal and collagen fibers, and a loss of clear demarcation
between adjacent collagen bundles. In normal tendons, stainable ground
substance (extracellular matrix) is absent and vasculature is not
conspicuous. Tenocytes are also generally inconspicuous, and fibroblasts
and myofibroblasts are absent. This is in stark contrast to patients with
symptomatic tendinopathy where there is obvious discontinuity and
disorganized collagen fibers with associated increased the amount of
mucoid ground substance. Ultrasonography can assess the thickness of
tendons and their echogenicity.
Patellar tendinopathy can be recalcitrant to treatment.5 Because of the
tensile loads placed on the patellar tendon, corticosteroids injections
within the tendon or tendon attachments are contraindicated. However,
new therapies are emerging, including sclerotherapy,6 prolotherapy,7
platelet-rich plasma (PRP) injections,8–10 and the use of growth factors11
and other cells12,13 to promote healing.
Anatomy: The patella functions as a classic sesamoid bone improving
mechanical leverage for the quadriceps mechanism.14 The patellar
tendon extends from the inferior patellar pole to the tibial tubercle and is
arranged in fascicles contributing to the course fibular appearance on
ultrasound (Figs. 69.1 and 69.2). The anterior fascicles are typically
longer than the posterior. The anterior attachments are longer because
they are more proximal to the patella and attach more distally to the tibia
than the corresponding posterior fascicles (see Fig. 69.1). The distal
attachment of the patellar tendon at the tibial tubercle is crescent-shaped,
and thus, the lateral fibers are more proximal than distal (see Fig. 69.1).
The posterior aspect consists of both an articular zone and a nonarticular
zone, which are devoid of tendon attachments and covered by a synovial
fold of tissue. During flexion of the knee, the patellar tendon length is
under the pull of the quadriceps tendon. For the same amount of
elongation, the shorter fascicles of the posterior ridge strain more than
the longer anterior fascicles. The shorter posterior attachments are more
vulnerable to develop overuse damage and tendinopathy. This can be
seen on a physical examination where the posterior aspect of the patellar
tendon attachments are usually more sensitive on a physical examination
and thus should be a target for injection. In order to palpate these fibers,
it requires the practitioner to lift the inferior pole of the patella up in
order to gain access for palpation (see Fig. 69.2).
Transducer: 7.5 to 15 MHz linear transducer.
Transducer orientation: Transducer placed in plane with the patellar
tendon as seen in Figure 69.3.

Needle: 27G 1.5-inch needle.

Local anesthetic: Lidocaine.

Technique: After a sterile preparation, the transducer is placed sagittal in


the plane with the patellar tendon. The transducer should be in position
so that the anterior pole of the patella is clearly visualized and the fibular
pattern of the patellar tendon is also well visualized. A 1% buffered
lidocaine solution is injected approximately 0.5 cm below the transducer,
and the needle is inserted in a shallow angle and moved under the
transducer head until the needle is well visualized. The needle can then
be directed toward the inferior patellar pole (see Fig. 69.3). If PRP is to
be injected, the tenoperiosteal insertions of the patella tendon must first
be infiltrated with lidocaine to reduce pain associated with a PRP
injection. The needle can be left in place for 1 to 2 minutes, allowing for
the anesthetic to take effect. The lidocaine syringe can then be
exchanged for the injectate or PRP for the procedure. Multiple small
aliquots, usually 0.2 to 0.3 mL, are injected at multiple locations
throughout the tendon with a focus at the regions of visible tendinopathy
and posterior attachments.
Knee: intra-articular injection/aspiration technique (suprapatellar)
Background and Indications: Corticosteroids have been used for the purpose of
intra-articular injections for osteoarthritis since hydrocortisone was introduced in
1951. Corticosteroids have had a long history of use for chronic knee pain.15
Randomized clinical trials and meta-analyses have concluded that corticosteroids
provide short-term benefit for chronic knee pain.16 Viscoelastic supplementation
injections have gained widespread acceptance in knee osteoarthritis.17
The accuracy of injection and misplaced injectate during intra-articular
injections contributes to variable clinical outcomes.18 Misplaced
corticosteroid injections at the knee may result in postinjection pain,
crystal synovitis, hemarthrosis, joint sepsis, and steroid articular
cartilage atrophy, as well as systemic effects, such as fluid retention or
exacerbation of hypertension or diabetes mellitus.19,20 Berkoff et al.21 in
a literature review noted that ultrasound guidance of knee injections
resulted in better accuracy than did anatomical guidance (95.8% versus
77.8%, P < .001). Large crystalline steroids injected into the joint
potentially may have mechanical erosive effects with weight-bearing
activity immediately following a corticosteroid injection; therefore, it
has been recommended by some to limit weight-bearing activity for at
least a day following a steroid injection.22 Not all physicians are in
agreement with this.5

Anatomy: The most important anatomical structures for an ultrasound-


guided intra-articular knee injection involves the anterior compartment
of the knee. The patella has a slight convex anterior surface. The
quadriceps tendon is attached to the anterior one-third to half of the
patella. The suprapatellar fat pad is located immediately under the
quadriceps tendon insertion (Fig. 69.4). More proximally, the prefemoral
fat pad is attached to the distal surface of the femur. A hypoechoic
elongated S-shaped structure situated between inferior surface of the
quadriceps tendon and the two fat pads is the suprapatellar synovial
recess, the common collector of the joint effusion and the target for
ultrasound-guided injections.
Transducer: 7.5 to 15 MHz linear transducer.

Transducer Orientation: The transducer is placed transversely over the


quadriceps tendon above the superior pole of the patella and can be slid
slightly medially or laterally (Fig. 69.5).

Needle: 25G 1.5-inch to 2-inch needle for intra-articular injections or a


22G 1.5-inch needle for aspiration procedures.

Local Anesthetic: Ropivacaine, procaine.

Position: Supine with pillow underneath flexed 30-degree knee.


Technique: After sterile skin and transducer preparation, the patient is
placed in supine position as noted previously. The ultrasound transducer
is placed transversely over the quadriceps tendon above the superior
pole of the patella and can be slid slightly medial to visualize the
hyperechoic joint effusion if a joint aspiration is to be performed. If
there is no knee effusion and the purpose of the procedure is for an intra-
articular injection, then the linear transducer is placed over the superior
pole of the patella in the transverse plane as shown in Figure 69.5. If an
aspiration is desired, having the knee completely extended pushes the
synovial fluid under the patella providing a means to clearly visualize
the synovial fluid below the patella, which will also extend beyond the
superior aspect of the patella. It is helpful to apply ample amount of gel
during this procedure, which allows for better visualization of the lateral
margin of the patella under the lateral aspect of the transducer head. It is
also helpful to apply only light pressure; otherwise, the fluid will be
displaced from the synovial recess. After skin anesthesia, the needle is
advanced in an in-plane fashion. The needle is directed under the patella
(see Fig. 69.5).
References
1. Khan K, Bonar F, Desmond P, et al. Patellar tendinosis (jumper’s
knee): findings at histopathologic examination, US, and MR
imaging. Victorian Institute of Sport Tendon Study Group.
Radiology. 1996;200:821–827.
2. Khan KM, Cook JL, Bonar F, et al. Histopathology of common
tendinopathies. Update and implications for clinical management.
Sports Med. 1999;27:393–408.
3. Ferretti A, Puddu G, Mariani PP, Neri M. The natural history of
jumper’s knee. Patellar or quadriceps tendonitis. Int Orthop.
1985;8:239–242.
4. Kannus P. Tendons—a source of major concern in competitive and
recreational athletes. Scand J Med Sci Sports. 1997;7:53–54.
5. Zhang W, Nuki G, Moskowitz R, et al. OARSI recommendations
for the management of hip and knee osteoarthritis: part III: changes
in evidence following systematic cumulative update of research
published through January 2009. Osteoarthritis Cartilage.
2010;18:476–499.
6. Hoksrud A, Bahr R. Ultrasoundguided sclerosing treatment in
patients with patellar tendinopathy (jumper’s knee). 44-month
followup. Am J Sports Med. 2011;39:2377–2380.
7. Ryan M, Wong A, Rabago D, et al. Ultrasoundguided injections of
hyperosmolar dextrose for overuse patellar tendinopathy: a pilot
study. Br J Sports Med. 2011;45:972–977.
8. Filardo G, Kon E, Della Villa S, et al. Use of platelet-rich plasma
for the treatment of refractory jumper’s knee. Int Orthop.
2010;34:909–915.
9. Kon E, Filardo G, Delcogliano M, et al. Platelet-rich plasma: new
clinical application: a pilot study for treatment of jumper’s knee.
Injury. 2009;40:598–603.
10. Volpi P, Marioni L, Bait C, et al. Treatment of chronic patellar
tendinosis with buffered platelet-rich plasma: a preliminary study.
Med Sport. 2007;60:595–603.
11. Anitua E, Andia I, Sanchez M, et al. Autologous preparations rich
in growth factors promote proliferation and induce VEGF and HGF
production by human tendon cells in culture. J Orthop Res.
2005;23:281–286.
12. Pascual-Garrido C, Rolón A, Makino A. Treatment of chronic
patellar tendinopathy with autologous bone marrow stem cells: a 5-
year-followup. Stem Cells Int. 2012:953510.
13. James SL, Ali K, Pocock C, et al. Ultrasound guided dry needling
and autologous blood injection for patellar tendinosis. Br J Sports
Med. 2007;41:518–521.
14. Basso O, Johnson D, Amis A. The anatomy of the patellar tendon.
Knee Surg Sports Traumatol Arthrosc. 2001;9:2–5.
15. Neustadt DH. Intra-articular injections for osteoarthritis of the knee.
Cleve Clin J Med. 2006;73:897–898, 901–904, 906–911.
16. Bellamy N, Campbell J, Robinson V, et al. Intra-articular
corticosteroid for treatment of osteoarthritis of the knee. Cochrane
Database Syst Rev. 2006;2:CD005328.
17. Bannuru R, Natov N, Obadan I, et al. Therapeutic trajectory of
hyaluronic acid versus corticosteroids in the treatment of knee
osteoarthritis: a systematic review and meta-analysis. Arthritis
Rheum. 2009;61:1704–1711.
18. Jackson DW, Evans NA, Thomas BM. Accuracy of needle
placement into the intra-articular space of the knee. J Bone Joint
Surg Am. 2002;84A:1522–1527.
19. Jones A, Regan M, Ledingham J, et al. Importance of placement of
intra-articular steroid injections. BMJ. 1993;307:1329–1330.
20. McGarry J, Daruwalla Z. The efficacy, accuracy and complications
of corticosteroid injections of the knee joint. Knee Surg Sports
Traumatol Arthrosc. 2011;19:1649–1654.
21. Berkoff D, Miller L, Block J. Clinical utility of ultrasound guidance
for intra-articular knee injections: A review. Clin Interv Aging.
2012;7:89–95.
22. Charalambous C, Paschalides C, Sadiq S, et al. Weight bearing
following intra-articular steroid injection of the knee: survey of
current practice and review of the available evidence. Rheumatol
Int. 2002;22:185–187.
70
Ultrasound-Guided
Intra-articular Ankle
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: The accuracy of ultrasound-guided


talotibial joint injections has been evaluated and shown in one study to
be 100% (20 out of 20) versus 85% (17 out of 20) for nonguided
injections.1 The accuracy rate for ultrasound-guided sinus tarsi injections
was 90% (18 out of 20) versus 35% (7 out of 20) for nonguided
injections in a cadaver study utilizing methylene blue for confirmation.1
The improved accuracy rate for intra-articular injections has made
ultrasound increasingly popular for both diagnostic and injection
procedures. In addition, ultrasound has been shown to have a higher
sensitivity than examinations and x-rays in detecting synovitis and joint
erosions and appears to be similar or even superior to other imaging
techniques such as magnetic resonance imaging (MRI) for this
purpose.2,3 In the case of gouty arthritis, ultrasound was found to be
more effective at detecting bone erosions in comparison to x-rays in
symptomatic as well as asymptomatic joints.4,5 Intra-articular injections
of hyaluronic acid (HA) for ankle osteoarthritis have also been
evaluated. Sun et al.6 evaluated 46 patients with unilateral ankle arthritis
who were receiving three weekly HA injections and reported pain relief
and improved function. Evidence is still weak with regard to the use of
HA in ankle osteoarthritis. Part of the problem is the complexity of the
weight-bearing stress and biomechanics of the ankle joint.

Anatomy: The ankle mortise joint is one of the more stable joints in the
body. The talar dome is wider at the anterior aspect and narrower at the
posterior aspect. Because of this anatomical relationship, when the ankle
is dorsiflexed, it wedges the talus between the tibia and lateral malleolus,
tightening the ligaments and stabilizing the joint. When the ankle is
plantar flexed, the more narrow portion of the talus lies within the mortis
and the joint is more mobile and less stable. Functionally, this provides a
stable ankle mortise during the midstance phase of the gait and continues
the stabilization of the ankle joint as the heel is lifted and one toes off in
the propulsion phase of the gait. In longitudinal scanning of the joint, a
triangular-shaped space is identified by the hyperechoic cortex of the
distal tibia and superior talar dome (Fig. 70.1). The articular capsule as
well as the hyperechoic fat pad can be easily visualized under ultrasound
in this anterior joint space.

The talar dome is covered by a hyaline cartilage, which is seen as a


hypoechoic layer over the hyperechoic bone. A small amount of fluid
will also be detected within the joint in normal individuals. Looking at
the ankle joint in the transverse plane, the talar dome appears as a
horizontal plateau that is seen as a hyperechoic transverse line from the
cortex of the bone of the talus and the hypoechoic hyaline cartilage
overlying the bone. The talar dome appears to have squared corners that
are an important anatomical feature that needs to be visualized for the
intra-articular injection as described as follows.

Transducer: 7.5 to 15 MHz linear transducer.

Transducer orientation Method 1 for the Ankle Joint Aspiration: Linear


transducer is held in the longitudinal axis (Fig. 70.2).

Method 2 for the Intra-articular Injection: The linear transducer is held


transversely over the joint surface (Fig. 70.3).
Needle: 27G 1.5-inch needle for intra-articular injection and 22G 1.5-
inch needle for aspiration procedure.

Local anesthetic: Ropivacaine, procaine.

Position: Patient is placed supine with the knee flexed to 90 degrees and
the foot flat on the table.

Techniques Ankle Joint Aspiration: After a sterile preparation, the


transducer is initially placed transversely over the ankle mortise joint in
order to locate and mark on the skin the dorsal pedis artery. Once the
dorsal pedis artery is marked, the transducer is then placed in a
longitudinal position over the joint. The transducer can then be slid
either medial or lateral to the neurovascular bundle. A local anesthetic is
injected over the skin just inferior to the ultrasound transducer. A 22G
needle can then be directed toward the hyperechoic V-shaped structure
created by the distal tibia and the talar dome as noted in Figure 70.2.
With a joint effusion, the particular capsule will fill and lift or displace
the periarticular fat pad, making it easier to enter the articular capsule
and perform the aspiration. If the procedure is done for the purpose of
culture and sensitivity, lidocaine should not be injected in the joint
because of its bacteriostatic properties.

Ankle Intra-articular Injection: After a sterile preparation, the linear


transducer is placed transversely over the joint. The talar dome and
ankle mortise joint are brought into view, as seen in Figure 70.3. The
transducer is moved in a medial direction until the medial talotibial joint
comes into view. In this position, a local anesthetic can be infiltrated
over the skin just medial to the transducer head and the needle is
directed in an in-plane position under the transducer head and directed
toward the ankle mortise. The ankle mortise joint will be seen as a thin
hypoechoic line, as seen in Figure 70.3.
The needle is then directed in plane under the transducer head and
directed into the joint.
References
1. Wisniewski S, Smith J, Patterson D, et al. Ultrasound-guided versus
nonguided tibiotalar joint and sinus tarsi injections: a cadaveric
study. PM R. 2010;2:277–281.
2. Filippucci E, Iagnocco A, Meenagh G, et al. Ultrasound imaging for
the rheumatologist VII. Ultrasound imaging in rheumatoid arthritis.
Clin Exp Rheumatol. 2007;25:5–10.
3. Micu MC, Nestorova R, Petranova T, et al. Ultrasound of the ankle
and foot in rheumatology. Med Ultrason. 2012;14:34–41.
4. Wright S, Filippucci E, McVeigh C, et al. High-resolution
ultrasonography of the first metatarsal phalangeal joint in gout: a
controlled study. Ann Rheum Dis. 2007;66:859–864.
5. Thiele R, Schlesinger N. Diagnosis of gout by ultrasound.
Rheumatology. 2007;46:1116–1121.
6. Sun S, Hsu C, Sun H, et al. The effect of three weekly intra-
articular injections of hyaluronate on pain, function, and balance in
patients with unilateral ankle arthritis. J Bone Joint Surg Am.
2011;93:1720–1726.
71
Ultrasound-Guided
Tarsal Tunnel
Injection
MICHAEL N. BROWN AND MICHAEL GOFELD

Background and indications: Tarsal tunnel syndrome is an entrapment


neuropathy that involves compression of the posterior tibial nerve or its
branches within the fibro-osseous tunnel that lies beneath the flexor
retinaculum. Symptoms may involve pain at the medial aspect of the
ankle extending into the medial aspect of the foot, the plantar aspect of
the foot, and the toes. Patients may experience paresthesias as well as
burning pain. Periodically, the patient’s pain can even radiate proximally
along the medial calf. Depending on the compression site and the
specific nerve branch involved, clinical symptoms may be more
localized to the medial plantar aspect of the heel. The cause is often
biomechanical faults and intrinsic foot structure that leads to
overpronation and increased tension on the posterior tibial nerve within
the tarsal tunnel. Branches of the posterior tibial nerve (medial and
lateral plantar nerves) can be entrapped by synovial cysts, bone and
other joint abnormalities, tumors, tenosynovitis, or vascular anomalies.
Ultrasound provides an excellent image guidance technique to place
local anesthetic within the tarsal tunnel and around the posterior tibial
nerve or its branches. On physical examination, a positive Tinel sign can
often be elicited over the posterior tibial nerve under the flexor
retinaculum. A triple compression stress test (TCST) is a physical
examination test that has been described where the ankle is placed in full
plantar flexion and the foot inverted and constant digital pressure
applied over the posterior tibial nerve. This will replicate paresthesias
and pain and has been shown to have a sensitivity of 85.9% and a
specificity of 100% for tarsal tunnel syndrome diagnosis.

Anatomy: The tarsal tunnel is a fibro-osseous tunnel that extends from


the posterior medial aspect of the ankle to the medial aspect of the
plantar region of the foot. It is divided into an upper (tibiotalar)
compartment and a lower (talocalcaneal) compartment. The lower tarsal
tunnel (Fig. 71.1) is covered by the flexor retinaculum (created by the
fusion of the superficial and deep aponeuroses of the leg) and the
abductor hallucis muscle with its fascia. Its osseous floor is formed by
the posteromedial aspect of the talus, the inferomedial aspect of the
navicular bone, and the medial aspects of the sustentaculum tali and
calcaneus. Typically, the posterior tibial nerve is found within the upper
tarsal tunnel, which trifurcates into its terminal branches just above the
medial malleolus. These branches are the medial and lateral plantar
nerves and the medial calcaneal nerve, (Fig. 71.2). The medial calcaneal
nerve can also branch from the lateral plantar nerve, and there are many
other anatomical variations.
The medial and lateral plantar neurovascular bundles are separated by
the interfascicular septum. Within the tarsal tunnel lie three important
tendons. Starting from the anterior position just behind the medial
malleolus lies the tibialis posterior tendon, next is the flexor digitorum
longus, and, finally, the flexor hallucis longus in the most posterior
position. These muscles can be remembered by the simple mnemonic
Tom (Tibialis posterior), Dick (flexor Digitorum longus) and Harry
(flexor Hallucis longus) listed from anterior to posterior. Typically, one
artery (the posterior tibial artery) and two veins on either side of the
artery can be seen within the tarsal tunnel. The posterior tibialis tendon
and the flexor digitorum longus demonstrate a typical hyperechoic
fibular structure consistent with the tendon, whereas the flexor hallucis
longus within the tunnel will appear much larger and will have the
appearance of muscle. This is because a significant muscular portion of
the flexor hallucis longus (FHL) lies within the tarsal tunnel. The FHL is
located posterior to the nerve. The nerve will often be found lying just
posterior to the vein as seen in Figure 71.1.

Transducer: 7.5 to 15 MHz linear transducer.

Transducer orientation: Transducer is placed along the plane of the


flexor retinaculum as seen in Figure 71.3.

Needle: 27G 1.5-inch needle.

Local anesthetic: Lidocaine.

Position: Patient can be placed either supine with the knee flexed and the
hip externally rotated to expose the medial aspect of the foot, or in a
lateral decubitus exposing the medial side of the foot.

Technique: The transducer is placed in an oblique position with the notch


directed toward the medial malleolus. The other end of the transducer is
directed toward the heel. The pulsation of the posterior tibial artery is
used for the initial orientation. Two veins will be on either side of the
artery. Two tendons are seen anterior to these three vessels (i.e., the
posterior tibialis tendon and the flexor digitorum longus). Just posterior
or medial to the vein, as noted in Figures 71.1 and 71.2, the nerve can be
visualized. The transducer can be slightly adjusted to eliminate
anisotropy and to achieve a typical honeycomb appearance of the nerve.
If the nerve has already branched, at least two nerves will be visualized:
the medial and lateral plantar nerves. The lateral plantar nerve is the
most posterior. A needle can be directed toward the nerve within the
tarsal tunnel either by in the in-plane or the out-of-plane approach. The
in-plane approach provides a means of coming behind the vascular
structures if the nerve is found posterior to the vein. An out-of-plane
approach is satisfactory if the nerve is found just posterior to the
vascular structures next to the vein. The needle tip can be directed to the
nerve and the tarsal tunnel and perineural regions around the nerve
where it can be infiltrated easily with a local anesthetic.

Bibliography Abouelela AA, Zohiery AK. The triple compression


stress test for diagnosis of tarsal tunnel syndrome. Foot (Edinb).
2012;22:146–149.
Delfaut E, Demondion X, Bieganski A, et al. Imaging of foot and ankle nerve entrapment
syndromes: from well-demonstrated to unfamiliar sites. Radiographics. 2003;23:613–623.
Govsa F, Blige I, Ozer MA. Variations in the origin of the medial and inferior calcaneal nerves.
Arch Orthop Trauma Surg. 2006;126:6–14.
Horwitz M. Normal anatomy and variations of the peripheral nerves of the leg and foot. Arch
Surg. 1938;36:626.
Lam S. Tarsal tunnel syndrome. J Bone Joint Surg Br. 1967;49:87–92.
Lau J, Daniels T. Tarsal tunnel syndrome: a review of the literature. Foot Ankle Int. 1999;20:201–
209.
Louisia S, Masquelet A. The medial and the inferior calcaneal nerves: an anatomic study. Surg
Radiol Anat. 1999;21:169–173.
Sarrafian S. Cross sectional and topographic anatomy. In: Sarrafian SK, ed. Anatomy of the Foot
and Ankle. 2nd ed. Philadelphia, PA: Lippincott Williams & Wilkins;1993:433–437.
72
Ultrasound-Guided
Peripheral Nerve
Stimulation

MICHAEL GOFELD

Background and indications: Peripheral nerve stimulation (PNS) was


the first attempted method for neuromodulation in chronic neuropathic
pain. The PNS effect is based on the gate control theory, wherein a
stimulation of thick A-beta fibers blocks the nociceptive input of spinal
neurons conducted via small A-delta and C fibers. Even before
development of the gate control theory, Sweet and Weptik experimented
with PNS by inserting stimulating electrodes onto the infraorbital nerve
and eliciting a paresthesia. Notably, they performed experiments on each
other. In the ensuing decades, multiple case series and observational
clinical studies were published reporting variable but overall significant
effectiveness of this method. Mononeuropathies, stump neuroma pain,
and the complex regional pain syndrome have been tagged as
indications. Because PNS is based on surgical implantation and the
achievement of paresthesia as sine qua non for pain relief, randomized
control studies are deemed impossible. This notable drawback is typical
not only for PNS but also for other complex health interventions in
which no pharmacotherapy-like research is possible.
Given the necessity to surgically explore the target nerve, often
accompanied by a need to perform external neurolysis and transposition,
PNS had been belonged to the neurosurgical domain until it became
apparent that ultrasound may help in localizing peripheral nerves and
neuromas. Ultrasonography may also facilitate preoperative localization,
planning of the access, and guiding dissection. Development of
percutaneous methods has been delayed because of regulatory
difficulties and insufficient interest amongst neurosurgeons and pain
physicians.
PNS can be arbitrary divided into two methods. The first method is a
stimulation of a definite nerve by either surgical or percutaneous lead
placement onto the nerve. The author prefers a surgical technique
because it is easier to suture the lead next to the nerve. The second
method can be best described as a stimulation of a painful skin region
that may or may not be outlined within cutaneous distribution of a
specific nerve. This method received the name peripheral field nerve
stimulation (PFNS).

Anatomy
1. PNS is performed according to the anatomical pathway of the target
nerve.
2. PFNS is performed based on pain distribution and only the lead
position is defined by the anatomical site of the pain.

Patient position: Variable and related to the specific target.

Probe: Broadband high-frequency linear transducer, low-frequency


curvilinear probe or a small (25-mm) linear probe for intraoperative
ultrasound.
Equipment: Either percutaneous or surgical (paddle) standard spinal
cord stimulation leads. In the past, custom-made leads (e.g., spiral,
sandwich-type) were used, but they are not commercially available.
Special PNS leads are currently under development.

Technique
Peripheral Nerve Stimulation

Percutaneous (cylindrical) lead placement technique: This is essentially


the same technique that is used for peripheral nerve continuous catheter
placements described elsewhere in the text. It is unclear which nerve-
lead configuration—parallel or perpendicular—should provide better
and more consistent stimulation. However, given peripheral nerves’
mobility and inability to anchor the lead in soft tissue, lead migration is
likely to happen. When this occurs, stimulation will be lost. A purely
percutaneous approach is most suitable for a PNS trial before
implantation. Even during a percutaneous trial, the limb must be
completely immobilized.

Paddle (surgical) lead placement technique: Paddle placement is a


neurosurgical approach that requires a cut down and mobilization of the
nerve. Ultrasound is an invaluable technique in all stages of implantation
and for postoperative management. A proximal segment of the intact
nerve of either upper or lower extremity is usually suited for the PNS
implantation. This allows avoiding both unnecessary excessive
exploration of a damaged nerve and crossing moving joints with lead
extensions. For the upper limb, typically the midarm is used for ulnar,
median, and radial nerve stimulation. Surgical access and stimulation of
the nerves of the lower limb are more complicated and less sustainable.
The femoral nerve is accessed below the inguinal crease. The saphenous
nerve can be found in the subsartorial space. The sciatic nerve and its
branches can be approached via different sites, such as the posterior
thigh, the popliteal, distal leg, and the proximal tarsal canal.
Planning of the pulse generator implantation is also challenging and
influenced by patient habitus, preferences, and local anatomy.
A preoperative ultrasound is performed in both the short (Fig. 72.1) and
long axis (Fig. 72.2) and the incision location is identified. It is helpful
to place the ultrasound probe according to the future skin incision and
learn what tissues need to be dissected or retracted in order to reach the
nerve. Color Doppler is an invaluable tool to assess and avoid vascular
structures during dissection. Use of ultrasound may help to minimize the
incision and to estimate the required depth of the wound.

During surgical access, an intraoperative ultrasound is useful for further


guidance (Fig. 72.3). Even a superficial nerve may be difficult to locate
in an open wound due to bleeding and tissue distortion caused by
retractors. The nerve may appear similar to a blood vessel, tendon, or
band of the adipose tissue.
Ideally, a paddle lead is placed under the nerve. Gentle and limited
mobilization is performed lest postoperative neurapraxia may occur. The
nerve sheath is minimally dissected to avoid excessive bleeding and to
provide a scaffold for the lead placement (Fig. 72.4). Arranging the lead
along the nerve into the sheath is usually sufficient to prevent lateral
dislodgement (Fig. 72.5). A nonabsorbable suture is placed at the neck
of the lead to prevent longitudinal migration. Alternatively, a specially
designed lead with the attached mesh sheet can be positioned under the
nerve, and the mesh is sutured to the underlying fascia (Fig. 72.6).
Placement of the lead under the nerve has an additional advantage. It
allows for follow-up ultrasound examinations in the postoperative period
to assess the healing process and complications (e.g., hematoma,
infection). It may also be useful in subsequent assessments of the lead
position and for guiding maintenance programming (Fig. 72.7).
Peripheral Field Nerve Stimulation: In PFNS, the role of ultrasound is
to control the depth of the lead placement. Inadvertent deep placement
into the muscle will result in painful muscle spasms. Conversely, a
superficial insertion may lead to painful stimulation and skin erosion.
Ideally, the electrode is inserted within the most superficial fascial plane.
At this level, the subcutaneous branches of peripheral nerves subdivide
before they end up as the cutaneous receptors.
The ultrasound transducer is positioned according to the chosen pathway
of the subcutaneous lead. The skin is anesthetized immediately next to
the short side of the probe at the operator side. Occasionally, the whole
future needle track is anesthetized to allow for the painless insertion of a
large-bore introducer and lead. In this case, no paresthesia can be
elicited until local anesthesia is worn off. A large-bore introducer (e.g.,
14G Tuohy needle or Angiocath) is placed in an in-plane approach with
the bevel facing up into the superficial facial plane and its progress is
carefully followed with ultrasound. It is important to ensure the needle
remains in the desired plane without penetrating into the underlying
muscles. The stylet is removed. Additional local anesthetic or normal
saline may be injected to expand the compartment during the needle
placement. A desired electrode is inserted through the needle. By
keeping the lead stable and applying gentle pressure, the needle is
removed over the lead. The final lead position is verified with a
longitudinal scan (Fig. 72.8). It is also useful to perform a short-axis
scan to ensure the lead is completely deployed in the fascial plane. The
lead is seen as a bright hyperechoic dot with hypoechoic shadow. If the
lead placement is temporary, for a stimulation trial, it should be secured
by either a suture or sterile acrylic glue. A transparent adhesive is
applied. If the lead was placed as a permanent implant, it is desirable to
perform a skin incision and pocket preparation first and then
implantation of the lead through the incision is performed. The lead is
usually secured to the superficial fascia with a suture and coiled in the
wound one to two times. After that, the rest of the tunneling and
connection with a pulse generator is performed similarly to techniques
implemented in the spinal cord stimulation.

Tips
1. The PNS trial is performed similarly to continuous nerve block
catheter placement. Limb immobilization is required for the trial
period.
2. Perioperative ultrasound is useful in planning surgical access and
localization of the target nerve for a paddle lead placement.
3. Postoperative ultrasound is a valuable imaging tool for the
assessment of the implanted lead position and facilitation of the
electronic programming.
4. Ultrasound is a useful imaging tool for PFNS that assists in the
precise identification of the superficial fascial plane and helps to
guide lead placement.

Suggested Readings
Burgher AH, Huntoon MA, Turley TW, et al. Subcutaneous peripheral nerve stimulation with
inter-lead stimulation for axial neck and low back pain: case series and review of the literature.
Neuromodulation. 2012;15:100–106.
Eisenberg E, Waisbrod H, Gerbershagen HU. Long-term peripheral nerve stimulation for painful
nerve injuries. Clin J Pain. 2004;20:143–146.
Hassenbusch SJ, Stanton-Hicks M, Schoppa D, et al. Long-term results of peripheral nerve
stimulation for reflex sympathetic dystrophy. J Neurosurg. 1996;84:415–423.
Huntoon MA, Burgher AH. Ultrasound-guided permanent implantation of peripheral nerve
stimulation (PNS) system for neuropathic pain of the extremities: original cases and outcomes.
Pain Med. 2009;10:1369–1377.
Huntoon MA, Hoelzer BC, Burgher AH, et al. Feasibility of ultrasound-guided percutaneous
placement of peripheral nerve stimulation electrodes and anchoring during simulated
movement: part two, upper extremity. Reg Anesth Pain Med. 2008;33:558–565.
Huntoon MA, Huntoon EA, Obray JB, et al. Feasibility of ultrasound-guided percutaneous
placement of peripheral nerve stimulation electrodes in a cadaver model: part one, lower
extremity. Reg Anesth Pain Med. 2008;33:551–557.
Skaribas I, Aló K. Ultrasound imaging and occipital nerve stimulation. Neuromodulation.
2009;13:126–130.
Slavin K, ed. Peripheral Nerve Stimulation (Progress in Neurological Surgery). Vol. 24. Basel,
Germany: S. Karger AG; 2011.
Sweet WH. Control of pain by direct electrical stimulation of peripheral nerves. Clin Neurosurg.
1976;23:103–111.
73
Ultrasound-Guided
Intrathecal Pump
Management
HADI S. MOTEN AND BRIAN DURKIN

Introduction
Since the discovery of dorsal horn mu receptors, the goal of the spinal
administration of opioids has been to minimize the amount of opioid
needed while still adequately controlling a patient’s pain. In 1991, the
U.S. Food and Drug Administration (FDA) approved the use of
programmable pumps for drug infusion. Then, in 1995, the FDA
approved the use of intrathecal morphine. This lead to a dramatic
increase in the use of intrathecal pumps (ITPs) to control pain.1
Currently, ITPs are used not only to control pain, but also to aid those
plagued by spasticity. With the increased use of ITPs, as well as an
increased typical duration of use, pain physicians find themselves faced
with the need to frequently access and refill ITPs. Accessing the pump’s
reservoir can become difficult due to a patient’s difficult anatomy,
complications from the underlying disease pathology, or pump
malposition.2 Traditionally, fluoroscopy has been utilized to aid in pump
refills, but recently, ultrasound-guided refills have become more
common. Ultrasound-guided ITP comes in two distinct varieties: sound-
assisted, where ultrasound is used to locate and mark the appropriate
access point,3 and real-time ultrasound-guided, where ultrasound is
continuously used not only to identify the location, but also to monitor
the administration of medication.4
Refill protocol including images
Once the intrathecal pump has been successfully implanted, telemetry is
utilized to interrogate the pump and identify the type of medication,
concentration, and dosage. A frequent follow-up is initially required to
identify appropriate dosage, assess pain control, and evaluate the patient
for signs of potential complications. The pump is interrogated to
ascertain the expected volume of medication in the reservoir. This is
compared to the actual amount remaining in the reservoir by engaging
the reservoir with a Huber needle and withdrawing all remaining
medication.
Contents of a typical ITP refill kit include
• Sterile drapes
• Pump template
• Huber needle
• Syringes
• Micropore filters
• Tubing
• Stop cocks
(Differences may exist between manufacturers and pump models.)
Refill process

Setup: The manufacturer’s kit is opened and prepared in a sterile


manner.

Skin Examination: Look for any gross abnormalities, rashes, or signs of


underlying pathology.
The pump is palpated, and its position is approximated (Fig. 73.1).
Computer Telemetry: The pump is interrogated, checked for
malfunction, and the volume of medication expected to remain is noted.

Preparation of Site: The area is prepared with pathogen appropriate


solution. A wide prep is utilized to account for any area that may come
in contact with the probe and allow for various approaches to the pump
location. Often, the identified area is carefully prepped using sterile
drapes (Fig. 73.2).

Probe Preparation: The ultrasound probe is prepped in a sterile manner,


and a sterile lubricant is applied to the skin.

Pump Is Identified: The ultrasound probe is placed over the lubricated


anticipated pump location. The pump reservoir, access port, and
surrounding tissue are identified. The surrounding tissue is evaluated for
signs of obvious pathology.

Standard Ultrasound Image: Basic landmarks when viewing a pump via


ultrasound (Fig. 73.3).

Difficult Access: Demonstrates the difficulty in accessing the refill port


in an ill patient that has a seroma over the refill port location.

Local Lidocaine Infiltration: The path of the Huber needle is


anesthetized with lidocaine to minimize discomfort (Fig. 73.4).
Accessing Port: In the image shown, the Huber needle is advanced along
the same path as the lidocaine infiltration. The out-of-plane approach is
visualized (Fig. 73.5).

Aspiration: Once the needle is engaged, the contents are aspirated and
this volume is compared to the pump’s anticipated volume (Fig. 73.6).
Infusion: Fresh infusate should be delivered via a microfilter in small
increments, allowing for frequent aspiration to ensure that the needle is
located within the lumen. Upon completion, the tubing is clamped off
using the clip to ensure that the infusate does not leak into the
surrounding tissue.

Infusion: Utilizing color Doppler, the infusion entering the appropriate


area will be identified. The color column indicates proper flow within
the anticipated area of the pump reservoir (Fig. 73.7).
Pocket Fill: A color Doppler indicating flow of infusate in areas not
expected to be within the pump’s reservoir. Flow around the anticipated
reservoir location would indicate a pocket fill (Fig. 73.8 and 73.9).

Infusion: A color Doppler indicated the spread of infusate within the


expected boundaries of the pump reservoir, which indicates a successful
pump refill.

Subcutaneous Infusion: A Doppler indicating the flow of infusate into


the subcutaneous tissue. Infusate is injected superior to the pump access
point, and the flow is not identified within the pump reservoir (Fig.
73.10).

Programming: Once completed, the pump is programmed with the


appropriate medication, concentration, dose, and alarm dates.
Risks
Accessing the intrathecal pump should only be attempted by properly
trained and skilled physicians. Any time the pump is accessed, there is
the potential for infection, introduction of a foreign body into the
underlying tissue or pump, and damage to the pump or catheter. Risks
associated with intrathecal pumps can be categorized as shown in Table
73.12,5:
Best practices
Many of the issues noted previously occur despite our best efforts. The
following is a list of practices that can help minimize untoward
complications:
• The physician with the assistance of a nurse should review and
confirm the correct medication and concentration before filling the
pump.
• Care should be taken to abide by the manufacturers guidelines and
avoid overfilling the pump.
• Ultrasound can be used as described previously to identify the
underlying anatomy as well as identify that the needle entered the
pump’s lumen correctly.
• The pump should be interrogated after filling is complete and the
medication, concentration, dosage, and alarm date should be printed
and checked against the patient’s chart.
• The patient’s skin should be evaluated before each fill, and the patient
should be questioned regarding skin ailments. If a rash develops, the
sterilization solution should be altered.
• When the pump is implanted, care should be taken to place the
catheter below the pump in order to minimize the likelihood of
catheter damage during refills.
• Care should be taken to aspirate every 5 mL to ensure proper needle
location and avoid subcutaneous filling.
• A proper sterile technique should be utilized along with stopcocks and
micropore filters to minimize the risk of contamination.

References
1. Wallace, M Yaksh TL. Long term spinal analgesic delivery: a
review of preclinical and clinical literature. Reg Anesth Pain Med.
2000;25:117–157.
2. Panchal SJ, Gonzales J. Intrathecal pumps. Tech Reg Anesth Pain
Manag. 2000;4(3):137–142.
3. Greher M, Eichenberger U, Gustorff B. Sonographic localization of
an implanted infusion pump injection port: another useful
application of ultrasound in pain medicine. Anesthesiology.
2005;102:243.
4. Hurdle MF, Locketz AJ, Smith J. A technique for ultrasound-guided
intrathecal drug-delivery system refills. Am J Phys Med Rehabil.
2007;86:250–251.
5. Gofeld M, McQueen C. Ultrasound-guided intrathecal pump access
and prevention of the pocket fill. Pain Med. 2011;12:607–611.
Index

A
Accessory axillary artery, 179, 180f
Accidental dural puncture (ADP), 339
Acetabulum, 540, 541f
Acromioclavicular joint injection, 520, 521f, 522, 522f ADARPEF. See
Association des Anesthesistes Reanimateurs Pediatriques d’Expression
Francaise (ADARPEF) Adductor canal
in infrapatellar nerve block, 222
in saphenous nerve block, 213, 215f
Adhesive capsulitis, 158
Adjuvants, 16–17, 17f
ADP. See Accidental dural puncture (ADP) Airway management equipment, 46f
Alcock canal, 455–456, 456f, 457f Alveolar artery, in mandibular nerve block,
425, 426f, 427
Ambulatory continuous peripheral nerve block pump, 132, 132f Amputees,
residual limb pain in, 494
Analgesia, perineural
adjuvants and, 16–17, 17f
complications of, 15
indications for, 14
pharmacology of, 14–17, 15f, 17t technique, 16–17
Anatomy
in acromioclavicular joint injection, 520, 521f in ankle block, 269
in anterior sciatic nerve block, 248, 249f as artifact, 73, 74f, 75, 75f, 76f, 77,
77f, 78f, 79, 79f, 80f in axillary nerve block, 176, 177f, 178f selective,
163, 164f, 165f in bicipital tendon sheath injection, 507, 508f, 509f in
carpal tunnel injection, 537, 538f
in caudal epidural anesthesia in pediatric patients, 388–389, 388f in caudal
epidural steroid injection, 331–332
in celiac ganglion block, 441, 442f
in cervical medial branch block, 474–475
in cervical nerve root injections, 481
in cervical sympathetic block, 433, 434f, 435, 435f, 436f in elbow/forearm
blocks, 181–190
in femoral nerve block, 202, 203f
in genitofemoral nerve block, 451
in glenohumeral joint injection, 510, 511f, 513, 513f in ilioinguinal and
iliohypogastric nerve blocks, 274, 276, 277
in infraclavicular block, 168, 169f, 170f, 171
in infrapatellar nerve block, 222–223
in interscalene block, 144f–147f, 147–148
with posterior approach, 136, 137f
in intra-articular ankle injection, 550, 551f in intra-articular elbow injection,
526, 527f, 528, 529f in intra-articular facet joint injection, 486, 487f, 488f
in intra-articular hip injection, 540, 541f, 542f in labor epidural, 339
in lateral femoral cutaneous nerve block, 209, 210f, 211f in lateral sciatic
block, 258, 259f, 262
in lateral thoracic paravertebral block, 316, 317f in lumbar facet medial
branch block, 486, 487f, 488f in lumbar plexus block, 194, 196–197
in lumbar transforaminal steroid injection, 484
in mandibular nerve block, 425, 426f
in maxillary nerve block, 423, 424f
neuraxial, 463, 464f–473f, 467, 472
in obturator nerve block, 231, 232f
paravertebral, 290–303
in paravertebral block, 306–307
in patellar tendon injection, 544, 545f, 546f in pectoral block, 322–323, 323f,
324f peripheral nerve, 32, 33f, 34, 35f, 36, 37f, 38f in piriformis injection,
501, 502f
in popliteal block, 258, 259f, 262
in pudendal nerve block, 455–456, 455f, 456f, 457f, 458f in quadratus
lumborum block, 283–286, 283f–286f in sacroiliac joint injection, 498,
499f in saphenous nerve block, 213–217
in serratus anterior block, 322–323, 323f, 324f in stellate ganglion block, 428,
429f, 430f in sternoclavicular joint injection, 523, 524f in subacromial
bursa injection, 504
in subgluteal sciatic block, 252, 253f, 254
in superior hypogastric plexus block, 446, 447f in supraclavicular block, 152,
153f, 154, 154f, 155f in suprascapular block, 159, 159f
in supraspinatus tendon injection, 516, 517f, 518f in tarsal tunnel injection,
554–555, 555f, 556f in third occipital nerve block, 474–475
in transversus abdominis block, 278, 279f in ulnar triquetral injection, 534,
535f variations in, 77, 78f, 79, 79f, 150, 150f wrist, 185, 188, 535f
in wrist injections, 531, 532f
Aneurysm, 83, 83f
Angle of isonation, 50, 59f
Animal model, pigs as, 114–115
Ankle block
anatomy in, 269
indications for, 269
technique for, 269–272
Ankle injection, 550–552, 551f, 552f Anterior longitudinal ligament, in labor
epidural, 339, 343f Aorta, in celiac ganglion block, 441, 442, 442f, 443f
Apical four-chamber view, 3f, 4, 5f Arterial dissection, 79
Arteries, 55, 75, 75f, 76f Arthroscopy, shoulder, 158
Articular process, 463, 464f, 466f, 467, 469f in intra-articular facet joint
injection, 486, 487f, 488, 488f, 489f in lumbar facet medial branch block,
486, 487f, 488, 488f, 489f in lumbar transforaminal epidural steroid
injection, 484
Artifacts
anatomic, 73, 74f, 75, 75f, 76f, 77, 77f, 78f, 79, 79f, 80f bony, 330
imaging, 60, 60f, 61, 62
Association des Anesthesistes Reanimateurs Pediatriques d’Expression
Francaise (ADARPEF), 349–350
Atherosclerosis, 79, 81f
Attenuation, 50, 51f
Axillary artery
accessory, 179, 180f
in axillary block, 179, 180f
in brachial plexus block in pediatric patients, 366
Axillary block
anatomy in, 176, 177f, 178f median nerve in, 177
musculocutaneous nerve in, 176, 179f
perivascular technique for, 177
technique for, 176–177, 179f
veins in, 179
Axillary nerve block
selective
anatomy in, 163, 164f, 165f posterior humeral circumflex artery in, 163,
164f, 166
technique, 166
suprascapular nerve block in, 158–159, 159f, 160f, 161

B
Biceps femoris, in anterior sciatic nerve block, 248, 249f Bicipital tendon sheath
injection, 507, 508f, 509, 509f Bladder tumor, transurethral resection of,
226, 229
Blood flow, 55, 56f
Blunt needles, intraneural injections with, 124
Bone shielding, 330
Botulinum toxin, 163
Botulinum toxin injection, for spasticity, 493–495, 495f Brachial artery
in axillary block, 177
in elbow/forearm blocks, 181, 182f
Brachial plexus, 35f, 37f, 39f in infraclavicular block, 168, 169f, 170f in
interscalene block, 143, 144f–147f, 147
with posterior approach, 136–137, 137f, 138f in supraclavicular block, 152,
153f, 154f, 155f, 156
Brachial plexus block, in pediatric patients
axillary approach for, 358–360, 359f, 360f infraclavicular approach for, 363–
364, 366, 366f, 367f interscalene approach for, 361–362, 362f, 363f
lateral infraclavicular block in, 367, 368f supraclavicular approach for,
363–364, 365f, 366
Brachial plexus variation, 77, 78f, 79f Brightness, 54
Bupivacaine, 18–19, 19t, 47
Buprenorphine, 16, 18

C
Carotid artery
in cervical sympathetic block, 433, 437f in stellate ganglion block, 428, 429f,
430, 431f Carpal tunnel injection, 537–539, 538f
Catheter insertion, 47, 49
abdominal, 128, 130f
advantages of, 128
ambulatory, 132, 132f
benefits of ultrasound in, 130–132
in brachial plexus block, in pediatric patients, 366
complications with, 128
epidural, in pediatric patients
caudal approach for, 393–394, 393f–398f, 396–397, 399
direct intervertebral approach for, 400–401, 400f, 401f, 402f in femoral
nerve block, 207
history of, 128
in infraclavicular block, 172
with medial approach, 174
in interscalene block, 149–150
with posterior approach, 141–142, 142f
lower extremity, 128, 130f
in lumbar plexus block, 201
nerve stimulation in, 130–131
in paravertebral block, 308
in pediatric patients, 132, 354–355
in popliteal block, 263
in sciatic block, 133f
in subgluteal sciatic block, 256
in supraclavicular block, 156
thoracic, 128, 129f
upper extremity, 128, 129f
Cauda equina, 467
Caudal, 13f
Caudal epidural anesthesia, in pediatric patients, 348–349
anatomy in, 388–389, 388f
indications for, 388
technique for, 392
Caudal epidural steroid injection
landmark palpation technique for, 331
sonofluoroscopic-guided
anatomy in, 331–332
procedure for, 333–334, 333f–337f, 337
ultrasound-only, 330–331
Celiac artery, in celiac ganglion block, 441, 442, 442f, 443f Celiac ganglion
block
anatomy in, 441, 442f
endoscopic, 442, 443f
indications for, 441
transabdominal, 444–445, 444f
Certification, in 3D/4D imaging, 70
Cervical axial transducer orientation, 473f Cervical block. See Interscalene block
Cervical ganglia, 428, 429f
Cervical medial branch block
anatomy in, 474–475
indications for, 474
technique for, 477, 478f, 479f, 480f Cervical nerve root injections, 481–482,
482f, 483f Cervical paramedian transducer orientation, 472f Cervical
sympathetic block. See also Stellate ganglion block anatomy in, 433,
434f, 435, 435f, 436f anterior approach for, 436, 438
fluoroscopic guidance in, 440
indications for, 433
lateral approach for, 435, 436, 438
technique for, 438, 439f, 440
Cervical sympathetic trunk, 433, 437f
Chassaignac tubercle, 428, 433, 481
Children. See Pediatric patients
Chloroprocaine, 47
Civco eTrax, 107, 108f
Clonidine, 15, 16, 18
Common peroneal nerve, 258, 261f, 265f, 266
Compound beam imaging, 55f, 57f Contrast, 54, 54f
Coracohumeral ligament, 513, 513f
Coracoid block, 367, 368f
Coronal, 13f
Corticosteroids, 16
Costotransverse ligament
in paravertebral anatomy, 303
in paravertebral block, 306, 307, 310
in thoracic paravertebral blockade, 290, 292, 293f Coulomb’s law, 22–23, 23f
Cranial, 13f
Cremaster muscle, in genitofemoral nerve, 451
Cross-sectional taxonomy, 13f
Curved arrays, 55, 57f

D
Deep circumflex artery, in ilioinguinal and iliohypogastric nerve blocks, 276
Deep peroneal nerve, in ankle block, 269, 270f Deep venous thrombosis (DVT),
79, 82f, 83f Deltopectoral groove, 366
Depth, penetration, 54, 55f
Dermatomes, 10, 11f
Development, of spinal cord, 386
Dexamethasone, 16–17, 18
Diabetic neuropathy, 86f
Diffraction, 50, 52f, 55f, 60
Direction, of beam, 50, 53f
Distal, 13f
“Donut” sign, 124–125
Dorsal nerve of clitoris, 455, 457f
Dorsal nerve of penis, 455, 457f
Dural sac, in caudal epidural steroid injection, 332
Dynamic range compression, 54, 54f

E
Echogenic needles, 100, 101f–104f, 103–105
Elbow/forearm blocks
anatomy in, 181–190
complications with, 191
indications for, 181
technique for, 191
Elbow injection, 526, 527f, 528f, 529f, 530, 530f Electrical stimulation. See also
Peripheral nerve stimulation (PNS) in anterior scalene block, 251
catheter insertion and, 130–131
in femoral nerve block, 205
for nerve location, 115–117
in parasacral block, 239, 246, 247f
in pediatric patients, 352
in subgluteal sciatic block, 256
thresholds, 43–44
ultrasound-guidance v., in research, 111–112
Electromagnetic needle guidance, 105, 106f, 107
Electromyography (EMG), 494–495
Elevation compounding, 67
EMG. See Electromyography (EMG)
Endoneurium, 33f, 34
Endoscopic celiac ganglion block, 442, 443f Endoscopic probe, 441
Endothoracic fascia, 306–307, 309, 316
Epidural anesthesia
labor
accidental dural puncture in, 339
anatomy in, 339
labor stages in, 339
neuraxial technique for, 341
in obese patients, 344
real-time neuraxial technique for, 344
in scoliosis patients, 345
in spinal instrumentation patients, 345
technique for, 340, 341f–344f, 344
in pediatric patients, 348–349
anatomy in, 388–389, 388f
indications for, 388
technique for, 392
Epidural catheters, in pediatric patients
caudal approach for, 393–394, 393f–398f, 396–397, 399
direct intervertebral approach for, 400–401, 400f, 401f, 402f Epidural space
in caudal epidural steroid injection, 332
identification of, 70
in neuraxial anatomy, 463
Epidural steroid injection, lumbar transforaminal, 484–485, 485f Epineurium,
32, 33f, 34, 35f Equipment, 45, 46f, 47, 47f, 48f, 49
Esophagus, in cervical sympathetic block, 433
Examination, two-dimensional ultrasound, neuraxial anatomy in, 463, 464f–
473f, 467, 472

F
Facet joint. See also Zygapophyseal joints in labor epidural, 339, 342f, 344f in
lumbar facet medial branch block, 486, 487f in neuraxial anatomy, 464f,
465f, 469f in third occipital nerve block, 474–475, 476f Fascia, 58
Fascia iliaca, 202, 205, 207
Fascicles, 32, 33f, 34
FAST exam, 6–7, 7f
Femoral artery
in anterior scalene block, 248, 249f
in anterior sciatic block, 251
in femoral cutaneous nerve block, 209, 210f, 211
in femoral nerve block, 202, 203f
in genitofemoral nerve block, 452, 453f
in saphenous nerve block, 213, 214f, 215f, 217, 218f, 220, 221
Femoral nerve, 202, 203f
in anterior scalene block, 248, 249f, 250f in genitofemoral nerve block, 453f
in lumbar plexus block, 370
in saphenous nerve block, 213, 214f
Femoral nerve block
anatomy in, 202, 203f
catheter insertion in, 207
distal, in pediatric patients, 375, 376f, 377f indications for, 202
intraneural injection in, 205, 207, 208f lateral femoral cutaneous block with,
209
in pediatric patients, 373–375, 374f
Femoral vein
in anterior sciatic nerve block, 248, 249f in obturator nerve block, 229f
Focal zone, 54, 55f
Focused assessment with sonography (FOCUS), 1–4, 2f–6f Forearm. See
Elbow/forearm blocks Forearm anatomy, 184

4D-imaging
acquisition in, 63–64, 64f, 65f challenges for, in neuraxial anesthesia, 69–70
defined, 64
display of, 64
multiplanar reconstruction in, 64, 65f, 66, 66f real-time rendering in, 68
of spine, 67
Frame rate, 69
Fraunhofer zone, 54, 55f
Frequency, 50, 51f, 56, 58f Fresnel zone, 54, 55f

G
Gain, 54, 60
Gel, 70
Genitofemoral nerve, 195
in femoral nerve block, 203f
in lateral femoral cutaneous nerve block, 210f in lumbar plexus block, 194,
195
Genitofemoral nerve block
anatomy in, 451
indications for, 451
technique for, 452–455, 452f, 453f Ghosts, 58, 59f
Glenohumeral joint injection
indications for, 510
modified rotator cuff interval approach for, 513–514, 513f, 514f, 515f
posterior approach for, 510, 511f, 512
Gluteus maximus
in piriformis injection, 501, 502f
in pudendal nerve block, 456, 456f, 457f, 459f GPS needle guidance, 107–
108, 108f, 109f H
Hip injection, intra-articular, 540, 541f, 542–543, 542f Hunter’s canal. See
Adductor canal Hyperechoic appearance, 36, 58, 59f
Hypoechoic appearance, 36, 58, 59f, 61

I
Iliac crest, 464f
Iliofemoral ligament, 540
Iliohypogastric nerve
in femoral nerve block, 203f
in lateral femoral cutaneous nerve block, 210f in lumbar plexus block, 194,
195
in quadratus lumborum block, 285f
Iliohypogastric nerve block
anatomy in, 274, 276, 277
indications for, 274
in pediatric patients, 409–411, 411f, 412f technique for, 275–277
Ilioinguinal nerve
in femoral nerve block, 203f
in lateral femoral cutaneous nerve block, 210f in lumbar plexus block, 194,
195
in quadratus lumborum block, 285f
Ilioinguinal nerve block
anatomy in, 274, 276, 277
indications for, 274
in pediatric patients, 409–411, 411f, 412f technique for, 275–277
Iliopsoas, in obturator nerve block, 231, 232f Image quality, 50, 51f–62f, 54–56,
58, 60–62, 69
Impedance, 23–29, 24f–29f Indications, 1
Infants. See Pediatric patients
Inferior gluteal artery, in pudendal nerve block, 456
Inferior rectal nerve, 455, 457f
Inferior thyroid artery, in cervical sympathetic block, 433, 435f Inferior vena
cava view, 3f, 4, 6f Infraclavicular block
anatomy in, 168, 169f, 170f, 171
catheter insertion in, 172
in coracoid process, 172f
in deltopectoral groove, 172f
“lateral and sagittal” technique for, 174
medial approach for, 172f, 173–174, 173f technique for, 171–174, 172f, 173f
Infraclavicular fossa, 366
Infraclavicular plexus, 168, 169f, 171
Infrapatellar nerve (IPN)
neuromas, 223
safe zone for arthroscopy with, 222
in saphenous nerve block, 217, 218f (See also Saphenous nerve block)
Infrapatellar nerve (IPN) block
anatomy in, 222–223
indications for, 222, 223
Injection pressure, 117–119
Innervation, 10
Interactive virtual reality simulators, 97, 98f Intercostal blockade, in pediatric
patients, 411–412
anterior, 412–413, 413f, 414f posterior, 413–414, 415f
Intercostal muscles
in lateral thoracic paravertebral block, 316, 317f, 319f in pectoral block, 324f
in serratus anterior block, 324f
Intercostal space anatomy, 317f
Interlaminar space, 467
Interligamentous place, 455
Internal pudendal artery, in pudendal nerve block, 455, 456f, 457f Interscalene
block
anatomy in, 144f–147f, 147–148
catheter insertion in, 149–150
direct plexus injection in, 125, 126
local anesthetic in, 148
periplexus injection in, 125, 126
with posterior approach, 136–137, 138f–142f technique for, 148–150
Interscalene space, in interscalene block, 143, 147
Interspinous ligament
in labor epidural, 339
in neuraxial anatomy, 463
Intra-articular ankle injection, 550–552, 551f, 552f Intra-articular elbow
injection, 526, 527f, 528f, 529f, 530, 530f Intra-articular facet joint injection
anatomy in, 486, 487f, 488f indications for, 486
technique for, 488, 489f, 490, 490f, 491f, 492f Intra-articular hip injection,
540, 541f, 542–543, 542f Intrafascicular injection, 117–118
Intraneural injection, 33f
with blunt needles, 124
in femoral nerve block, 205, 207, 208f
incidence of, 126
intrafascicular, 117–118
needle-to-nerve proximity and, 114
nerve stimulation thresholds and, 43
in popliteal block, 126
swelling in, 124, 125f
Intrathecal pump (ITP), 565–569, 566f–571f, 571–572, 572t Intrathecal space, in
neuraxial anatomy, 463
IPN. See Infrapatellar nerve (IPN) Ischial spine, in pudendal block, 456, 456f,
457f, 459f Ischiorectal fossa, in pudendal nerve block, 455
ITP. See Intrathecal pump (ITP)

K
Knee arthroscopy, 222, 223
Knee injection, 544, 545f–548f, 547–548

L
Laminae, 463, 467, 470f, 471f, 472
LAST. See Local anesthetic system toxicity (LAST) Lateral, 13f
Lateral arcuate ligament, 284
Lateral epicondyle injection, 526, 527f, 528f, 529f, 530, 530f Lateral femoral
cutaneous nerve
anatomy, 209, 210f
in femoral nerve block, 203f
in lumbar plexus block, 370
Lateral femoral cutaneous nerve block
anatomy in, 209, 210f, 211f indications for, 209
technique for, 209, 210f, 211, 211f Lateral infraclavicular block, 367, 368f
Lateral pectoral nerve (LPN), 322, 323f, 327
Lateral thoracic paramedian transducer orientation, 467f Lidocaine, 47
Life-Tech EchoBright needle, 102f, 103, 104
Ligamentum flavum
in labor epidural, 339, 340, 341f, 342f, 343f in neuraxial anatomy, 463, 472
Linear arrays, 55, 57f
Liposomal bupivacaine, 18–19, 19t
Local anesthetics (LAs), 14–17, 15f, 17t in pediatric patients, 352–355
Local anesthetic system toxicity (LAST), 355
Longus colli muscle
in cervical sympathetic block, 433, 434f, 435, 437f, 438, 440
in stellate ganglion block, 428, 430f, 431f, 432
LPN. See Lateral pectoral nerve (LPN) LTL. See Lunotriquetral ligament (LTL)
Lumbar facet medial branch block
anatomy in, 486, 487f, 488f indications, 486
technique for, 488, 489f, 490, 490f, 491f, 492f Lumbar plexus block
anatomy in, 194, 196–197
continuous, 201
indications for, 194
lateral insertion approach for, 196–198
medial approach for, 198, 199–200
in pediatric patients, 370–371, 372f, 373
sagittal approach for, 198, 200f
technique for, 194, 196–201
Lumbar transducer orientation, 466f
Lumbar transforaminal epidural steroid injection, 484–485, 485f Lumbar
transverse transducer orientation, 465f Lung, in supraclavicular block, 152,
153f, 156
Lunotriquetral ligament (LTL), 534
Lymph nodes, as artifact, 75, 76f

M
Magnetic coupling needle guidance, 107, 107f Mandibular nerve, 425, 426f
Mandibular nerve block
anatomy in, 425, 426f
indications for, 425
technique for, 425, 426f, 427, 427f Mannequins, 98, 98f, 99f Maxillary artery,
in mandibular nerve block, 427, 427f Maxillary nerve, 423, 424f
Maxillary nerve block
anatomy in, 423, 424f
indications for, 423
technique for, 423, 424f
Maximal intensity projection, 66–67
Mechanically steered arrays, 63, 64f
Mechanical needle guidance, 105, 106f
Medial, 13f
Medial branch nerve, 486. See also Lumbar facet medial branch block Medial
pectoral nerve (MPN), 322, 323f
Medial thoracic paramedian transducer orientation, 469f Median nerve
in axillary block, 177
in carpal tunnel injection, 537, 538f
in elbow/forearm block, 181, 182
Median nerve block. See Elbow/forearm blocks Mepivacaine, 47
Mesoneurium, of sciatic nerve, 260
Midazolam, 16
Minimal intensity projection, 67
Minimum stimulating current (MSC), 115–117
Mirror artifacts, 61, 61f
MPN. See Medial pectoral nerve (MPN) MSC. See Minimum stimulating current
(MSC) Multiplanar reconstruction (MPR), 64, 65f, 66, 66f, 68, 69
Muscle, as artifact, 75, 77f
Musculocutaneous nerve, in axillary block, 176, 179f N
NCS. See Nerve conduction studies (NCSs) Needle(s), 40
blunt, intraneural injections with, 124
echogenic, 100, 101f–104f, 103–105
in pediatric patients, 352
visibility, 69
Needle guidance
clinical data on, 109–110
combined electromagnetic, 105, 106f, 107
GPS, 107–108, 108f, 109f history of, 111
magnetic coupling, 107, 107f
mechanical, 105, 106f
real-time 4D, 68–69
simulator for, 93
software in, 110
Needle-to-nerve proximity
electrical stimulation in, 115–117
importance of, 114
injection pressure and, 117–119
ultrasound and, 119–120
Nerve conduction studies (NCSs), 494–495
Nerve depolarization, 22
Nerve impedance, 28–29, 29f
Neuraxial anatomy, 463, 464f–473f, 467, 472
Neuropathology, 84, 85f
Neurostimulation. See Electrical stimulation O
Obese patients
distal parasacral block in, 244, 246, 247f femoral nerve block in, 205
interscalene block in, 143
labor epidural in, 344
lateral paravertebral block in, 316
lumbar plexus block in, 198
obturator block in, 226, 231
speclation with, 50
Oblique muscles
in ilioinguinal and iliohypogastric nerve blocks, 275–276
in quadratus lumborum block, 285f
in transversus abdominis block, 278, 280f Obturator externus, in obturator
nerve block, 226, 227f, 228f Obturator internus
in parasacral block, 236, 245f
in pudendal block, 455
Obturator nerve
in “3-in-1” block, 205
anatomy, 226, 227f
in anterior scalene block, 248, 250f
in lateral femoral cutaneous nerve block, 210f in lumbar plexus block, 370
Obturator nerve block
assessment of, 229
distal
anatomy in, 231, 232f
indications for, 231
technique for, 234–235
indications for, 226
technique for, 227–228
Omohyoid artery, in supraclavicular block, 153f Opacity, 68
Ossification, in pediatric patients, 386

P
Pajunk Sonoplex needles, 103, 103f, 104, 104f Parasacral block
distal
assessment of, 245
indications for, 244
in-plane technique for, 246
sciatic nerve in, 244–246
technique for, 244–245
proximal
cephalad approaches for, 240
indications for, 236
technique for, 237–243
Parasacral parallel shift, 240
Parasternal long-axis view, 2, 3f
Parasternal short-axis view, 2, 3f, 4, 4f Paraumbilical block, in pediatric patients,
407–409, 409f, 410f Paravertebral anatomy, 290–303, 306–307
Paravertebral blocks
lateral thoracic
anatomy in, 316, 317f
indications for, 316
technique for, 317–318, 318f–321f medial
anatomy in, 306–307, 306f
blind approach for, 305
catheter insertion in, 308
complications with, 31
indications for, 305
loss-of-resistance technique for, 306
pressure-measurement technique for, 306
sagittal approach for, 307–308, 308f–311f transversal approach for, 308–
309, 311–314, 311f, 312f in pediatric patients, 415–417, 417f, 418f
Paresthesia, 111
Patellar tendon injection, 544, 545f–548f, 547–548
Patient position, 1–2, 2f
Pectineus, in obturator nerve block, 227f, 228f, 229f, 230f, 231, 232f Pectoral
block
indications for, 322
technique for, 325, 327, 327f–329f Pediatric patients
adult versus, 350
anatomy in, 348
anterior sciatic block in, 379–380, 383
brachial plexus block in
axillary approach for, 358–360, 359f, 360f infraclavicular approach for,
363–364, 366, 366f, 367f interscalene approach for, 361–362, 362f,
363f lateral infraclavicular block in, 367, 368f supraclavicular
approach for, 363–364, 365f, 366
catheters in, 132, 354–355
caudal epidural anesthesia in, 348–349
anatomy in, 388–389, 388f
indications for, 388
technique for, 392
dosing guidelines for, 352–355
epidural catheters in
caudal approach, 393–394, 393f–398f, 396–397, 399
direct intervertebral approach for, 400–401, 400f, 401f, 402f equipment for,
350–352
femoral nerve block in, 373–375, 374f
distal, 375, 376f, 377f history of anesthesia in, 348
iliohypogastric block in, 409–411, 411f, 412f ilioinguinal nerve block in, 409–
411, 411f, 412f intercostal blockade in, 411–412
anterior, 412–413, 413f, 414f posterior, 413–414, 415f
lumbar plexus block in, 370–371, 372f, 373
midfemoral sciatic nerve block in, 383
nerve stimulation in, 352
ossification in, 386
paraumbilical block in, 407–409, 409f, 410f paravertebral block in, 415–417,
417f, 418f popliteal block in, 383–384
probes for, 351–352
rectus sheath block in, 407–409, 409f, 410f sciatic nerve block in, 375, 378,
379f, 380f, 381f, 382f skeleton in, 386
spinal cord development in, 386
toxicity in, 353–355
transversus abdominis plane block in, 404–406, 406f, 407f ultrasound in, 351–
352, 386–387
vertebral column in, 386
Pediatric Regional Anesthesia Network (PRAN), 350
Pedicles, 463
Perineal nerve, 455
Perineurium, 32, 33f, 34, 35f Peripheral field nerve stimulation, 562–563, 563f
Peripheral nerve anatomy, 32, 33f, 34, 35f, 36, 37f, 38f Peripheral nerve
stimulation (PNS), 558–560, 559f–563f, 562–563
Peritoneum, in transversus abdominis plane block, 278
Peroneal nerve
in ankle block, 271
in popliteal block, 258, 259f
Pharmacology, 14–17, 15f, 17t Phased arrays, 55, 57f, 64, 64f Physiatry, 493
Pigs, as animal model, 114–115
Piriformis, in parasacral block, 236, 239, 240, 245f Piriformis injection, 501–
503, 502f, 503f Planes, 13f
Plexus sheath, in supraclavicular block, 152
Plexus trunks, 34, 35f, 36, 37f, 38f, 39f, 40f PNS. See Peripheral nerve
stimulation (PNS) Popliteal artery, 258, 259f, 261f Popliteal block
anatomy in, 258, 259f, 262
complications with, 268
indications for, 258
intraneural injection in, 126
in pediatric patients, 383–384
posterior approach for, 263–264
supine/lateral approach for, 265–267
Popliteal vein, 258, 259f, 261f Posterior approach, interscalene block with, 136–
137, 138f–142f Posterior border of ischium (PBI), in parasacral block, 244,
245f, 246
Posterior humeral circumflex artery (PHCA), in axillary nerve block, 163, 164f,
166
Posterior renal fascia, 284
Posterior sacral plate, in caudal epidural steroid injection, 334, 334f Posterior
tibial artery, in ankle block, 269, 271f Posterior tibial nerve, in ankle block,
269, 271f PRAN. See Pediatric Regional Anesthesia Network (PRAN)
Pressure, injection, 117–119
Prevertebral fascia, in stellate ganglion block, 428, 430f, 431f, 432
Probe depth, 54
Probe movements, 53f
Proximal, 13f
Pseudoaneurysm, 83, 84f
Psoas compartment, 371
Psoas compartment block, 370
Psoas major, 194, 196
Pterygoid plate
in mandibular nerve block, 425, 427, 427f in maxillary nerve block, 423, 424f
Pudendal canal, 455, 456, 456f, 457f Pudendal nerve, 455, 456f, 457f, 459f
Pudendal nerve block
anatomy in, 455–456, 455f, 456f, 457f, 458f fluoroscopy in, 461
indications for, 455
technique for, 460–461
transgluteal approach for, 461
Pudendal neuralgia, 455
Pulsed mode, 56, 58f

Q
Quadratus femoris, in femoral nerve block, 206f Quadratus lumborum, 283–284,
284f, 285f Quadratus lumborum block
anatomy in, 283–286, 283f–286f indications for, 283
technique for, 286–288, 287f, 288f Quadratus lumborum space, 284f

R
Real-time ultrasound-guided neuraxial labor epidural, 344
Rectus sheath (RS) block, in pediatric patients, 407–409, 409f, 410f Recurrent
laryngeal nerve
in cervical sympathetic block, 433
in stellate ganglion block, 428
Reflection, 50, 51f, 58, 59f Reflectors, in echogenic needles, 100, 101f, 102f,
103f Refraction, 50, 51f, 60
Rendering, 66–67, 68
Residual limb pain, in amputees, 494
Reverberation, 60, 61
Rib tubercle, 467
Ropivacaine, 47
Rotator cuff interval, in glenohumeral joint injection, 513–514, 513f, 514f, 515f
Rotator cuff pathology, 158

S
Sacral ala, 464f
Sacral cornua, in caudal epidural steroid injection, 333, 333f, 334, 335f, 336f
Sacral groove, 464f
Sacral hiatus
in caudal epidural anesthesia in pediatric patients, 388–389, 390f, 391f in
caudal epidural steroid injection, 335f Sacral hiatus apex, in caudal
epidural steroid injection, 331
Sacral plexus, in parasacral block, 236, 239
Sacrococcygeal ligament
in caudal epidural anesthesia in pediatric patients, 389, 390f, 391f in caudal
epidural steroid injection, 331, 333, 333f, 334
Sacroiliac joint, 498, 499f
Sacroiliac joint injection, 498–500, 499f, 500f Sacrospinous ligament, 455, 456,
457f, 459f Sacrotuberous ligament, 455, 456f, 457f Sagittal, 13f
Sagittal paravertebral space, 299f–302f, 303
Saphenous nerve, in ankle block, 269, 270f, 272f Saphenous nerve block. See
also Infrapatellar nerve (IPN) block anatomy in, 213–217
in ankle, 224
indications for, 213
midthigh approach for, 220
perivenous approach for, 220
technique for, 220
Saphenous vein
in ankle block, 272f
in saphenous nerve block, 213, 214f
Sartorius muscle
in femoral nerve block, 205
in lateral femoral cutaneous nerve block, 209, 210f in saphenous nerve block,
217
“Saw” sign, in labor epidural, 340
Scalenes
in interscalene block, 144f, 145f, 146f, 147, 147f in supraclavicular block, 152,
155f
Scan and mark approach, 463
Scapholunate ligament (SLL), 534
Scattering, 50, 52f
Schwannoma, 85f
Sciatic block
anterior
anatomy in, 248, 249f
indications for, 248
in pediatric patients, 379–380, 383
technique for, 248–251
catheter insertion in, 133f
lateral
anatomy in, 258, 259f, 262
complications with, 268
indications for, 258
posterior popliteal approach for, 263–264
technique for, 265–267
midfemoral, in pediatric patients, 383
in pediatric patients, 375, 378, 379f, 380f, 381f, 382f subgluteal
anatomy in, 252, 253f, 254f indications for, 252
technique for, 253–256
Sciatic nerve, 38f
mesoneurium of, 260
in parasacral block, 236, 244–246
in piriformis injection, 501, 502f
in popliteal/lateral sciatic bock, 258–260, 266, 267
in pudendal block, 456f, 457f, 459f Sclerotomes, 10, 12f
Scoliosis, in labor epidural, 345
SCTL. See Superior costotransverse ligament (SCTL) Semimembranosus, in
anterior scalene block, 248, 249f Semitendinosus, in anterior scalene block,
248, 249f Serratus anterior block
anatomy in, 322–323, 323f, 324f indications for, 322
technique for, 325, 327, 327f–329f SGHL. See Superior glenohumeral
ligament (SGHL) Shadowing, 61, 61f, 62
Shoulder arthroscopy, 158
Side lobe artifact, 60, 60f
Signal strength, 50
Simulator
high-fidelity, 88
interactive virtual reality, 97, 98f
mannequins, 98, 98f, 99f method, 88–91, 89f–92f for needle guidance, 93
Skeleton, in pediatric patients, 386
Skin preparation, 45
SLL. See Scapholunate ligament (SLL) Smoothing, 68
Sodium channels, 14, 15f
Software, needle enhancement, for needle guidance, 110
Soma AxoTrack, 107, 107f
Sonographic principles, 50, 51f–62f, 54–56, 58, 60–62
SonoSim Ultrasound Training Solution, 97, 98f Spasticity, botulinum injection
for, 493–495, 495f Speclation, 52f, 60
Spermatic cord, in genitofemoral nerve block, 451, 452, 453f Spinal cord
development, in pediatric patients, 386
Spinal instrumentation, in labor epidural, 345
Spinal roots, 34, 35f, 36, 37f, 38f, 39f, 40f Spine, 3D/4D imaging of, 67
Spinous process
in labor epidural, 340, 341f, 342f, 343f in neuraxial anatomy, 463, 471f, 472
Stellate ganglion block. See also Cervical sympathetic block anatomy in, 428,
429f, 430f indications for, 428
Sterile procedure, 45, 47, 48f, 49
Sternoclavicular joint injection, 523–524, 524f, 525f Sternocleidomastoid, in
interscalene block, 143, 146f Stimulation thresholds, 43–44
Subacromial bursa injection, 504–505, 505f, 506f Subclavian artery
in interscalene block, 143, 145f
in supraclavicular block, 152, 153f, 155f Subclavian vein, in supraclavicular
block, 152
Subcostal four-chamber view, 3f, 4, 5f Subepineurial injection, 34
Sulcus ulnaris, 181
Superior costotransverse ligament (SCTL)
in paravertebral anatomy, 303
in paravertebral block, 306, 307, 310
in thoracic paravertebral blockade, 290, 292, 293f Superior glenohumeral
ligament (SGHL), 513, 513f Superior hypogastric plexus, 446, 447f
Superior hypogastric plexus block
anatomy in, 446, 447f
indications for, 446
technique for, 447, 448f, 449f, 450f Supraclavicular block
anatomy in, 152, 153f, 154, 154f, 155f indications for, 152
technique for, 154, 156
Suprascapular block
anatomy in, 159, 159f
in axillary nerve block, 158–159, 159f, 160f, 161
complications of, 161
indications for, 158–159
technique for, 161
Suprascapular nerve, 158
Supraspinatus tendon, in subacromial bursa injection, 505, 506f Supraspinatus
tendon injection, 516, 517f, 518f, 519, 519f Supraspinous ligament
in labor epidural, 339
in neuraxial anatomy, 463
Surface rendering, 66

T
TAP. See Transversus abdominis plane block (TAP) Tarsal tunnel injection, 554–
555, 555f, 556f, 557, 557f Tendons, as artifact, 73, 74f
TFCC. See Triangular fibrocartilage complex (TFCC) Third occipital nerve
block
anatomy in, 474–475
indications for, 474
technique for, 475, 476f, 477f Thoracic axial transducer orientation, 470f, 471f
Thoracic paramedian transducer orientation, 468f, 469f

3D imaging
acquisition in, 63–64, 64f, 65f challenges for, in neuraxial anesthesia, 69–70
display of, 64
multiplanar reconstruction in, 64, 65f, 66, 66f planes in, 65f
rendering techniques in, 66–67
single volume acquisition in, 67–68
of spine, 67
“3-in-1” block
femoral nerve block and, 205
lateral femoral cutaneous block in, 209
Threshold, 67
Thyroid gland, in cervical sympathetic block, 433, 434f, 435f, 437f Thyroid
pathology, 84, 86f, 87f Tibial nerve
in ankle block, 269, 271f
in popliteal block, 258, 259f, 261f, 263, 265f, 266
Time gain compensation, 60, 61, 61f, 62
Tissue harmonic imaging, 56
Tissue type, signal strength at, 50, 51f Tourniquet pain, 163
Training
in 3D/4D imaging, 70
simulator, 88–91, 89f–92f, 93, 94f–99f, 96–98
Transabdominal celiac ganglion block, 444–445, 444f Transurethral resection of
bladder tumor (TURT), 226, 229
Transverse, 13f
Transverse cervical artery, in interscalene block, 145f Transverse paravertebral
space, 292, 293f–297f Transverse process
in labor epidural, 339, 342f, 344f in neuraxial anatomy, 463, 465f, 468f, 472
Transversus abdominis plane block (TAP)
anatomy in, 278, 279f
indications for, 278
in pediatric patients, 404–406, 406f, 407f technique for, 279–281
Triangular fibrocartilage complex (TFCC), 534, 536f TURT. See Transurethral
resection of bladder tumor (TURT) Two-dimensional ultrasound
examination, neuraxial anatomy in, 463, 464f–473f, 467, 472
U
Ulnar nerve, in elbow/forearm blocks, 181, 182, 183
Ulnar nerve block. See Elbow/forearm blocks Ulnar triquetral injection, 534,
535f, 536, 536f UltraSonix GPS, 108, 109f

V
Variation, anatomic, 77, 78f, 79, 79f, 150, 150f Vascular sonopathology, 79, 80f,
81f Vasculature, as artifact, 75, 75f, 76f Vastus medialis, in saphenous nerve
block, 213, 214f, 216f Veins, 55, 75, 75f, 76f Velocity gates, 55
Vertebrae
in neuraxial anatomy, 466f, 467, 468f, 469f, 472
in thoracic paravertebral space, 290, 291f Vertebral arch, 463
Vertebral artery, in interscalene block, 146f Vertebral body, 470f
in labor epidural, 339, 341f, 342f in neuraxial anatomy, 463, 466f
Vertebral column, in pediatric patients, ultrasound considerations with, 386
Vertebral foramen, 463, 467
Visible Humans, 88–91, 89f–92f Volume contrast imaging, 67
Volume rendering, 67

W
Wavelength, 50, 51f
Whiplash syndrome, 474
Wrist anatomy, 185, 188, 535f
Wrist injections, 531, 532f, 533, 533f. See also Carpal tunnel injection Y
Y ligament Bigelow, 540

Z
Zygapophyseal joint(s). See also Facet joint syndrome, 486
in third occipital nerve block, 474
Zygoma, in maxillary nerve block, 423, 424f

You might also like