Sunteți pe pagina 1din 15

Anesthesia for elective eye surgery

Authors
Alvaro A Macias, MD
Joseph Bayes, MD
Kathryn E McGoldrick, MD, FCAI(Hon)
Section Editor
Jeffrey H Silverstein, MD
Deputy Editor
Nancy A Nussmeier, MD, FAHA
Disclosures: Alvaro A Macias, MD Nothing to disclose. Joseph Bayes, MD Nothing to disclose. Kathryn E McGoldrick,
MD, FCAI(Hon) Nothing to disclose. Jeffrey H Silverstein, MDGrant/Research/Clinical Trial Support: Covidien
(delirium/POCD); CasMed (delirium/POCD); Hospira (delirium/POCD [Dexmedetomidine]); Foresight (delirium/POCD
[Cerebral oximeter and probes]); BIS (delirium/POCD [Processed EEGT Monitors and probes]). Nancy A Nussmeier, MD,
FAHA Employee of UpToDate, Inc.
Contributor disclosures are reviewed for conflicts of interest by the editorial group. When found, these are addressed by
vetting through a multi-level review process, and through requirements for references to be provided to support the content.
Appropriately referenced content is required of all authors and must conform to UpToDate standards of evidence.
Conflict of interest policy

All topics are updated as new evidence becomes available and our peer review process is complete.
Literature review current through: Feb 2015. | This topic last updated: Jan 09, 2015.
INTRODUCTION The goals of anesthetic care during elective eye surgery are pain-free surgery,
facilitation of the surgical procedure, rapid recovery, and minimization of risks associated with surgery and
anesthesia.
This topic reviews the techniques for providing analgesia, sedation, or anesthesia during cataract,
glaucoma, and vitreoretinal surgery. Other aspects of surgical management of these conditions are
discussed separately. (See "Cataract in adults", section on 'Treatment' and "Open-angle glaucoma:
Treatment", section on 'Types of therapy' and "Angle-closure glaucoma", section on
'Treatment' and "Retinal detachment", section on 'Treatment'.)
ANESTHESIA FOR CATARACT SURGERY Cataract surgery is one of the most common procedures
requiring anesthetic care [1].
Preoperative consultation The preoperative medical evaluation is reviewed in detail elsewhere,
including preoperative testing, evaluation of comorbid conditions, and perioperative decisions regarding
chronically administered medications. (See "Cataract in adults", section on 'Preoperative medical
evaluation'.)
Consultation for anesthetic management emphasizes the following additional considerations [2]:
Anesthetic considerations Most cataract surgery is performed with a topical or regional anesthetic
technique, combined with monitored anesthesia care (MAC). The preanesthetic consultation includes:
Assessment of communication and cooperation skills Communication and cooperation skills
are essential (ie, patients ability to understand, communicate, and cooperate with commands to
avoid all movement) if topical or regional anesthesia with MAC is planned.
Focused medical history Performance of a focused medical history, including use of any
anticoagulant or antithrombotic therapies, previous eye operations, and overall suitability for sameday surgery [3].

Focused physical examination Detection of abnormalities on the physical examination,


including any conditions that interfere with the ability to lie supine comfortably (eg, congestive heart
failure, severe chronic obstructive pulmonary disease, severe back pain, or claustrophobia) if topical
or regional anesthesia with MAC is planned. A standard assessment of the airway is always
performed.
Ophthalmic considerations
Anticoagulant and antithrombotic therapies: implications for bleeding For cataract surgical
patients having a high risk of clotting and embolic complications due to cardiac or vascular
pathology, administration of therapeutic doses of aspirin and warfarin are continued throughout the
perioperative period [4,5]. One large retrospective study noted no higher incidence of sightthreatening bleeding complications after regional anesthesia (ie, eye block) in patients taking
aspirin, warfarin, or clopidogrel up until the time of cataract surgery, compared with those who
discontinued the therapy [6].
However, limited data are available regarding the risk of bleeding during cataract surgery in patients
receiving dual antiplatelet therapy (eg, aspirin plus clopidogrel). Dual antiplatelet therapy is often
used after placement of a drug-eluting stent, and fatal stent clotting may develop if antiplatelet
medications are prematurely stopped. In these cases, we suggest delaying eye surgery, if possible,
until after the minimum period recommended for daily administration of dual antiplatelet therapy
[5,7]. (See "Cataract in adults", section on 'Antithrombotic agents' and "Antiplatelet therapy after
coronary artery stenting".)
Axial eye length: implications for globe puncture during regional block The axial length of
the eye (distance from the cornea to the retina) is measured by ultrasound before a cataract
operation to determine the proper intraocular lens size to be implanted. It has been noted that
patients with long eyes (axial length >25 mm) have an increased risk of needle injury during a
retrobulbar (intraconal) block, usually due to penetration of the posterior pole of the globe [8-10]. A
history of myopia in childhood or the presence of globe-enveloping intraorbital hardware such as a
scleral buckle are indications that the eye may be longer than average [10]. (See 'Retrobulbar block
(intraconal block)' below.)
Also, patients with an abnormal outpouching of the eye (staphyloma) (figure 1), usually located in
the posterior portion of the globe, are at increased risk for puncture by a retrobulbar needle [9,10].
In such patients, retrobulbar block is usually avoided in favor of a peribulbar (extraconal) or subTenon block, topical anesthesia, or general anesthesia. (See'Regional anesthesia' below.)
Prior intraocular gas bubble: implications for expansion by nitrous oxide Patients with a
previously injected intraocular gas bubble cannot receive nitrous oxide until an ophthalmologists
examination has been performed to document complete absorption of the gas bubble. There are
multiple case reports describing blindness after nitrous oxide administration within a few months of
gas bubble injection [11-13]. (See 'Anesthesia for vitreoretinal surgery' below.)
Other aspects of the surgical treatment of cataracts are presented separately. (See "Cataract in adults",
section on 'Treatment' and "Cataract in children", section on 'Management'.)
Monitoring and sedation
Needle insertion for regional anesthesia The following considerations apply during needle insertion
for a regional block:

American Society of Anesthesiologists (ASA) standard monitors (including capnography for


moderate or deep sedation) should be used during performance of the regional block, as well as
during surgery [14]. The anesthesiologist must be vigilant for evidence of the oculocardiac reflex
(resulting in bradycardia or asystole) or accidental injection of local anesthetic into a blood vessel
(resulting in systemic toxicity) or into the central nervous system (resulting in brainstem anesthesia)
[2]. Although rare, these situations might necessitate airway support or other emergent intervention.
(See 'Oculocardiac reflex manifestations' below and 'Systemic complications' below.)
Supplemental oxygen is frequently administered at the time of initial sedation and during the block
to reduce the risk of hypoxemia during sedation for the regional block.
Sedatives (eg, midazolam) and/or opioids (eg, remifentanil) with a short duration of action are
administered immediately before the regional block to reduce or eliminate the pain of needle
insertion and injection of the local anesthetic. The goal of sedation is to minimize anxiety while
providing the maximum degree of safety. An alternative technique is administration of small doses
of propofol (eg, 10 to 20 mg increments) until the patient loses consciousness for approximately two
minutes while the block is performed.
The anesthesiologist should avoid prolonged sedation, since subsequent performance of the
surgical procedure requires an awake and cooperative patient. (See'Surgical procedure' below.)
Surgical procedure During monitored anesthesia care (MAC) for the surgical procedure, the following
considerations apply:
Sedation is minimized to reduce the risk of side effects (eg, restlessness, confusion,
unresponsiveness, or airway obstruction) that may jeopardize the patients ability to cooperate
during surgery. Patient cooperation is necessary throughout the procedure to avoid any head
movement. Even minor movement, which is highly magnified under a microscope, may result in eye
injury.
If the patient experiences eye pain, the ophthalmologist is alerted to request administration of
additional local anesthetic. Heavy sedation should not be used as a substitute for inadequate
analgesia [15].
Compressed room air (21 percent oxygen concentration) is delivered under the surgical drapes. If
supplemental oxygen is required during the procedure, anair/oxygen mixture (with an oxygen
concentration <30 percent) is used. Minimizing oxygen concentration reduces the risk of fire during
surgery with use of a heat source (eg, electrocautery) [16].
The gas (room air or an air/oxygen mixture <30 percent oxygen) is delivered under the surgical
drapes at a rate of 10 L/minute. Otherwise, completely covering the face with drapes may cause
rebreathing of clinically significant amounts of carbon dioxide (CO 2). Another option is use of a
vacuum line to remove expired CO2.
Anesthetic techniques Most cataract operations are performed with MAC and a topical or regional
anesthetic technique [2]. In the United States, topical analgesia is the most common technique, followed
by peribulbar (extraconal) block or, less commonly, retrobulbar (intraconal) block [17]. Sub-Tenon blocks
are most popular in the United Kingdom [18]. Typically, general anesthesia is reserved for adults who are
unable to communicate, cooperate, or remain stationery during eye surgery, and for children [2].
The technical considerations, benefits, and drawbacks of each of these anesthetic techniques are
described below.

Topical anesthesia
General principles Topical anesthesia is a good choice for short, uncomplicated procedures in
which the surgeon does not require complete akinesia (eg, cataract surgery with
phacoemulsification through a small corneal incision), as well as in the following situations:
Fully anticoagulated patients, due to concerns regarding the risk of bleeding during needle
insertion to perform a regional technique.
Patients with monocular vision (ie, unilateral blindness), due to concerns regarding the risk of
vision loss in the remaining eye because of perforation or penetration of the globe during
performance of a regional eye block.
Technique The anesthesiologist or the ophthalmologist applies local anesthetic eye drops or
gels on the cornea and conjunctiva (eg, lidocaine 2% jelly,proparacaine 0.5% solution,
or tetracaine 0.75% solution) [19]. Anesthetic gels produce greater levels of drug in the anterior
chamber than equal doses of drops and may provide superior surface analgesia. However, gels may
form a barrier to bactericidal agents; thus, they are administered after antiseptic solutions.
Benefits Topical analgesia may be used as the sole anesthetic technique. The risk of
complications may be lower than with other anesthetic techniques since no needle is used. Vision is
quickly regained in the postoperative period. (See 'Retrobulbar block (intraconal block)' below
and 'Peribulbar block (extraconal block)'below and 'Sub-Tenon block' below.)
Drawbacks Topical analgesia cannot provide akinesia of the eye. Since the patient must
voluntarily remain motionless for the entire procedure, appropriate patient selection is important.
Anxious patients with a low pain threshold would fare better with a regional anesthetic technique or
general anesthesia.
Risks of topical anesthesia are minimal, but very rarely, an allergic reaction or infection may occur.
Regional anesthesia Profound anesthesia and akinesia of the eye are provided with a successful
regional anesthesia technique (ie, eye block).
Commonly used agents The local anesthetics that are commonly used include:
Lidocaine in a concentration not higher than 2%, due to concerns of myotoxicity with higher
concentrations
Bupivacaine 0.75%
Typically, a combination of lidocaine and bupivacaine is used. Some clinicians use ropivacaine 0.75%;
reported results include less pain on injection compared with other local anesthetics as well as excellent
intraoperative akinesia and postoperative pain control [20,21].
The enzyme hyaluronidase as an ancillary agent is usually used with local anesthetics. Concentrations
between 1 and 7.5 units/mL are most commonly used, but concentrations as low as 0.75 units/mL may be
effective [22]. The addition of hyaluronidase:
Increases tissue permeability of the local anesthetic
Promotes dispersion of the local anesthetic
Reduces the increase in orbital pressure associated with the injected volume
Enhances the quality of the block
Reduces the risk of injury to the extraocular muscles

Epinephrine is rarely required to prolong the duration of anesthesia if a combination


of lidocaine and bupivacaine is used, since block duration will usually exceed one hour, which is adequate
for most procedures.
Peribulbar block (extraconal block)
General principles Peribulbar (extraconal) block was developed as a safer alternative to the
retrobulbar (intraconal) block for providing anesthesia and akinesia of the eye [23]. The needle is
placed less deeply and at a different angle compared with placement for retrobulbar block.
Theoretically, these modifications make the peribulbar block less likely to result in perforation of the
globe posteriorly, injury to the optic nerve, or injection into the central nervous system with resultant
brainstem anesthesia [23,24].
Also, peribulbar block produces more reliable akinesia of the orbicularis oculi, the muscle that closes
the eyelids, compared to retrobulbar block [2,25]. This is due to the larger volume and distribution of
local anesthetic injected (see 'Retrobulbar block (intraconal block)' below). Specifically, the patient
cannot open the eyelid because cranial nerve III is blocked. Although the patient may be able to
close the eyelid because cranial nerve VII is usually not completely blocked, the lid closure force is
typically weak [25].
For all of these reasons, the peribulbar technique is more popular, but the retrobulbar block has not
been entirely replaced [26]. (See 'Retrobulbar block (intraconal block)' below.)
Technique Typically, between 4 to 8 mL of local anesthetic solution is used. A single injection in
the inferior lateral quadrant of the orbit is usually sufficient [27,28] (figure 2).
A needle 1 inch long is recommended to reduce the risk of injuring structures deep in the orbit. We
use an Atkinson needle, which has a relatively blunt tip compared with the traditional needle used for
intramuscular injections. The Atkinson needle may enhance the operators ability to identify scleral
tissue if it is encountered before perforation of the globe. However, some clinicians use sharper
needles to minimize the pain of insertion and to theoretically limit the amount of damage to the globe
if inadvertent perforation does occur.
Benefits The likelihood of inadvertent perforation of the globe or brainstem is theoretically less
than with a retrobulbar (intraconal) block [23,24]. Also, peribulbar block produces more reliable
akinesia of the orbicularis oculi, the muscle that closes the eyelids, compared to retrobulbar block
[25]. Thus, a supplemental facial nerve block is rarely necessary. The success rate of the peribulbar
(extraconal) block in producing anesthesia and akinesia of the eye is reported to be at least 84
percent, but may be higher with experience [27,29].
Drawbacks At least five minutes are required for the block to take effect, which is slightly longer
than with a retrobulbar (intraconal) block. Conjunctival chemosis is more common after peribulbar
block, compared to retrobulbar block [26]. Globe injury is possible, although one study reported no
instances of globe penetration or perforation in a series of 2000 peribulbar injections [24].
Retrobulbar block (intraconal block)
General principles Retrobulbar (intraconal) block produces profound analgesia and takes effect
rapidly, in less than five minutes. This block was previously the predominant technique, but it is used
less frequently now since theoretically safer techniques have been developed. (See 'Topical
anesthesia' above and 'Peribulbar block (extraconal block)' above and 'Sub-Tenon block' below.)

Technique Approximately 5 mL of local anesthetic is placed inside the muscular cone formed by
the four recti muscles. Usually, a 1.25 inch needle or shorter is used to reduce the risk of
complications (figure 3).
Benefits Retrobulbar block has a faster onset than peribulbar block and is associated with less
chemosis (ie, swelling or edema of the conjunctiva). In experienced hands, it has a success rate of
>85 percent [30].
Drawbacks Although controversial, retrobulbar (intraconal) block is thought to have a higher risk
of serious complications (eg, globe perforation and brainstem anesthesia) than the peribulbar
(extraconal) block technique [2,10,23,24]. (See 'Peribulbar block (extraconal block)' above
and 'Anesthetic complications' below.)
Globe perforation The incidence of globe perforation as a result of retrobulbar block varies
depending on many factors, including the skill of the operator and patient anatomy [31]. In two
case series of patients receiving retrobulbar blocks, the incidence of globe penetration or
perforation was 0.03 to 0.08 percent [8,32].
Patients with long axial eye length (>25 mm) have an increased risk of needle injury to the globe
during performance of a retrobulbar block. One report notes a high incidence of inadvertent globe
perforation in 1 out of 140 patients (0.7 percent) having an axial length 26 mm [33]. Thus, other
regional techniques (eg, peribulbar [extraconal] block or sub-Tenon block), topical anesthesia, or
general anesthesia are preferred in these patients. (See 'Ophthalmic considerations' above.)
If retrobulbar (intraconal block) is indicated to produce profound akinesia despite a long axial eye
length, the approach for the needle insertion may be modified by introducing the needle less deeply
and changing its angulation to reduce the risk of striking and injuring the posterior portion of the
elongated globe.
Detection and management of globe perforation are discussed below. (See 'Globe or optic
nerve perforation' below.)
Brainstem anesthesia The incidence of apparent central spread of local anesthetic into the
brainstem was reported to be 16 cases in a series of 6000 patients receiving retrobulbar blocks
[34]. Eight of these patients (0.13 percent) developed respiratory arrest.
Absence of eyelid akinesia Although akinesia of the orbit is profound, retrobulbar block
may leave the orbicularis oculi muscles fully functional. Eyelid squeezing could cause
extrusion of the intraocular contents during critical periods of certain procedures (eg, during
corneal transplantation). Thus, a facial nerve block is often necessary to prevent squeezing of
the eyelid [2,35].
Eyelid hematoma Eyelid hematoma is more common after retrobulbar block compared to
peribulbar block [26].
Sub-Tenon block
General principles The sub-Tenon block employs a blunt cannula, rather than a needle, to
induce regional anesthesia of the eye [2,36,37]. It is used in the United Kingdom and some other
countries more often than in the United States. Onset of the block is rapid, but the extent of akinesia
is variable and is proportional to the volume of local anesthetic injected [37,38].
Absolute contraindications to the sub-Tenon block include infection or a prior scleral buckle, while
relative contraindications include prior retinal or glaucoma surgery.

Technique A blunt cannula is inserted through a small incision in the conjunctiva and Tenon
capsule, also known as the episcleral membrane, with subsequent deposition of approximately 5 mL
of local anesthetics (figure 4). The local anesthetic reaches the posterior part of the globe, even with
an anterior injection. Approximately five minutes is required for the block to take effect [2,39].
Shorter (12 mm), more flexible plastic cannulae or ultrashort (6 mm) cannulae may be preferable to
longer, more rigid metallic cannulae, although they are associated with a higher incidence of
conjunctival hemorrhage and chemosis. Also, a newer minimally invasive technique for sub-Tenon
block without incision has been developed [40].
Benefits The success rate for sub-Tenon block for producing anesthesia of the eye is reported to
be >97 percent [38,41].
Because a blunt cannula (not a needle) is used for the block, globe perforation and other serious
complications such as brainstem anesthesia can occur but are rare [18,42,43]. In one series of 6000
sub-Tenon blocks, there were no serious block-related complications [18].
In very myopic patients with an elongated axial eye length, there is a reduced risk of posterior pole
perforation since the technique avoids needle placement in the posterior orbit.
Also, since major hemorrhage is rare with the sub-Tenon block, it may be a good choice for the
anticoagulated patient at risk for retrobulbar hemorrhage [18].
Drawbacks Chemosis and minor subconjunctival hemorrhage occur with a higher frequency
compared with needle blocks [43]. Chemosis is unlikely if small volumes are injected via a long
cannula.
General anesthesia A small percentage of adult patients undergoing cataract surgery require general
anesthesia. These include patients who are unable to communicate and those who cannot cooperate due
to neurocognitive dysfunction, severe anxiety, or claustrophobia. General anesthesia is also considered in
patients who are unable to lie supine comfortably (eg, patients with severe and symptomatic congestive
heart failure, chronic obstructive pulmonary disease, or back pain) or unable to remain motionless (eg,
patients with tremor disorders such as Parkinson disease, severe anxiety, or claustrophobia).
Furthermore, most children require general anesthesia because they are not able to reliably remain
motionless during eye surgery.
During general anesthesia for eye surgery, a deep plane of anesthesia is maintained to avoid
laryngospasm, coughing, or other movement. Endotracheal intubation is often used, particularly in
patients at risk for aspiration during general anesthesia. If an endotracheal tube is used, a
nondepolarizing neuromuscular blocking agent (eg,rocuronium or vecuronium) is administered and
titrated according to monitoring of muscle relaxation with a peripheral nerve stimulator.
However, in patients without specific aspiration risk, the laryngeal mask airway (LMA) has been used with
increasing frequency during cataract surgery [44,45]. Advantages of the LMA include less risk of
increasing IOP during insertion and removal due to less straining and coughing [45,46]. As with
endotracheally intubated patients, a deep plane of anesthesia is maintained during eye surgery to avoid
laryngospasm, coughing, or other movement. As with any facial surgery, vigilance must be maintained to
detect accidental displacement of the LMA.
ANESTHESIA FOR INTRAOCULAR GLAUCOMA SURGERY Anesthetic techniques for intraocular
glaucoma surgery (eg, trabeculectomy) are the same as those for cataract surgery. (See 'Anesthesia for
cataract surgery' above and 'Anesthetic techniques' above.)

Topical anesthesia can be the choice for patients for glaucoma surgery to avoid any transitory increase in
intraocular pressure (IOP) during injection of local anesthetic for a regional anesthetic.
Outcomes after glaucoma surgery appear to be similar regardless of the anesthetic technique used. In a
randomized study of 120 consecutive glaucoma patients undergoing combined phacotrabeculectomy with
either a regional technique (peribulbar local anesthesia) or topical anesthesia with 2% lidocaine jelly,
there were no differences in pain control and satisfaction during or shortly after the procedure, and no
differences in intraocular pressure (IOP) or the incidence of bleb leakage at follow-up after one year [47].
Similarly, anesthetic technique did not influence the success of trabeculectomy surgery during a longer
follow-up period (4.2 years) in 57 patients receiving either a regional or topical anesthetic technique [48].
However, some surgeons prefer general endotracheal anesthesia in patients undergoing trabeculectomy,
due to specific concerns regarding increased risk of damage to the optic nerve due to injections of local
anesthetic in glaucoma patients, as well as concerns regarding poor healing after administration of
subconjunctival lidocaine[49,50].
During general anesthesia for intraocular surgery, complete akinesia is necessary. Therefore, a
nondepolarizing neuromuscular blocking agent (eg, rocuronium orvecuronium) is administered and
titrated according to monitoring of muscle relaxation with a peripheral nerve stimulator, and a deep plane
of anesthesia is maintained [2].
Other aspects of the surgical treatment of glaucoma are presented separately. (See "Open-angle
glaucoma: Treatment", section on 'Laser therapy' and "Open-angle glaucoma: Treatment", section on
'Surgery' and "Angle-closure glaucoma", section on 'Laser peripheral iridotomy' and "Angle-closure
glaucoma", section on 'Other surgery'.)
ANESTHESIA FOR VITREORETINAL SURGERY Patients undergoing vitreoretinal surgery (eg, a
detached retina) usually receive a regional anesthetic block and/orgeneral anesthesia (without the use
of nitrous oxide), rather than topical anesthesia [51,52]. Topical anesthesia alone is avoided because
surgery may be quite lengthy, and it is imperative that the patient does not move [53].
If general anesthesia is used, patients should not receive nitrous oxide when injection of gas (eg, SF6 or
C3F8) is planned, or when gas was previously used to create a bubble to internally tamponade the
detached retina, unless an ophthalmologist has documented that the bubble has been completely
absorbed. Although SF6 is usually completely absorbed by 10 days, and C3F8 by six weeks, there are case
reports of blindness due to use of nitrous oxide after 25 days for SF 6 and after 41 days for C3F8[12].
Retinal detachment operations are basically extraocular, but may become intraocular if the surgeon elects
to perforate and drain subretinal fluid [2]. Hence, patients are managed in the same manner as those
having intraocular surgery (see 'Anesthesia for intraocular glaucoma surgery' above). Furthermore,
rotation of the globe with traction on the extraocular muscles during retinal detachment operations may
elecit the oculocardiac reflex. Thus, vigilance must be maintained to detect bradycardia and other
arrhythmias. (See 'Oculocardiac reflex manifestations' below.)
Other aspects of vitrectomy for detached retina and diabetic retinopathy are discussed separately.
(See "Retinal detachment", section on 'Treatment' and "Diabetic retinopathy: Prevention and treatment",
section on 'Vitrectomy'.)
ANESTHETIC COMPLICATIONS The complications of anesthesia for ophthalmic surgery can be both
vision- and life-threatening [53]. An acronym to remember the serious complications of eye surgery is

OPHTS [54]. O stands for optic nerve perforation (very rare and extremely unlikely with needles 1.25
inches or less), P for globe perforation, H for hemorrhage (eg, retrobulbar hemorrhage), T for toxic
reactions to local anesthetics (eg, injury to extraocular muscles caused by injection of these agents), and
S for systemic adverse effects (eg, spread of local anesthetic into the central nervous system or
intravascular injection with resultant cardiorespiratory depression or arrest).
Ophthalmologic complications
Globe or optic nerve perforation Globe perforation is a rare but serious complication of ocular
regional anesthesia. It is more common in myopic patients with long axial eye length (>25 mm).
(See 'Retrobulbar block (intraconal block)' above.)
Symptoms of ocular perforation are variable, ranging from intense ocular pain with abrupt loss of vision
and hypotonus, to no signs or symptoms. The clinician performing the block may have a sense of
increased resistance, particularly if a blunt needle is used.
If ocular perforation is suspected, the ophthalmologist must be notified immediately to perform
ophthalmoscopy or ultrasound in order to assess the damage. Usually the planned surgery must be
cancelled, and the patient is referred to a retinal surgeon. Occasionally, the damage can be managed with
cryosurgery, laser treatment, or mere observation. More commonly, however, proliferative
vitreoretinopathy occurs, often accompanied by retinal detachment, and this is managed with vitrectomy
and retinal reattachment surgery.
A devastating scenario occurs when the ocular perforation is undetected and an intraocular injection of
local anesthetic occurs. Less than 2 mL of solution injected into the globe can produce an ocular
explosion and permanent blindness in the affected eye.
Hemorrhage Bleeding secondary to needle-based techniques is not uncommon. Bleeding may be
superficial or deep, arterial, or venous. Superficial hemorrhage, while not vision-threatening, may produce
an unsightly circumorbital hematoma. In contrast, retrobulbar hemorrhage, when arterially based, may
cause sudden bleeding and a palpable, dramatic increase in IOP, as well as globe proptosis and upper lid
entrapment. This can jeopardize the globes vascular supply, with a potentially devastating effect on
vision. The incidence of retrobulbar hemorrhage has been reported to be 0.03 [55] to 3 percent [56].
If hemorrhage is suspected, immediate consultation with an ophthalmologist is indicated, and fundoscopic
examination, tonometric determination of IOP, ultrasound assessment, and even a lateral canthotomy or
paracentesis may be required. Continuous ECG monitoring is necessary because the oculocardiac reflex
may occur as blood extravasates from the muscle cone. (See 'Oculocardiac reflex manifestations' below.)
If the hemorrhage is mild or moderate, the decision to proceed with surgery depends on several factors,
including the amount of bleeding, the nature of the proposed ophthalmic surgery, and the patients
condition. In cases of severe hemorrhage, surgery should be cancelled.
Intramuscular injection Intramuscular injection of local anesthetics may cause injury to extraocular
muscles. This is thought to be a cause of postoperative strabismus [57].
Oculocardiac reflex manifestations Manifestations of the oculocardiac reflex can occur when
pressure is applied to extraocular muscles. These include bradycardia (a decrease of 10 to 20 percent in
the basal heart rate), junctional rhythms, hypotension, and, rarely, asystole. This reflex can occur during
injection of local anesthesia or during the surgical procedure itself.

Management includes stopping the stimulus (eg, release of traction or manipulation of the extraocular
muscles). If this is ineffective, an anticholinergic medication (eg,atropine or glycopyrrolate) is
administered.
The risk of inducing this reflex may be reduced by an effective regional anesthetic block or general
anesthesia with adequate depth.
Systemic complications Emergency equipment must be immediately available, including
resuscitation drugs and emergency airway equipment (eg, bag and mask, airways, and intubation
equipment), even though serious systemic complications due to regional or general anesthetic techniques
are rare during eye surgery.
Spread of local anesthetic into the central nervous system (ie, the brainstem) is possible, with resultant
cardiorespiratory depression or arrest requiring airway management and cardiopulmonary resuscitation.
As with all nerve blocks, accidental intravascular injection of local anesthetic may lead to systemic toxicity.
Treatment consists of IV administration of 20% lipid emulsion 1.5 mL/kg bolus followed by
0.25 mL/kg/min infusion, and supportive airway and hemodynamic management. Calcium channel
blockers, beta blockers, and local anesthetics (eg, lidocaine, procaine) should be avoided; vasopressin is
not recommended, and initial doses of epinephrine should be small (10 to 100 mcg IV). If, after 30
minutes, there is no clinical improvement, the bolus dose of 1.5 mL/kg lipid emulsion should be repeated,
and the continuous infusion should be increased to 0.5mL/kg/min. (See "Overview of peripheral nerve
blocks", section on 'Local anesthetic systemic toxicity'.)
SUMMARY AND RECOMMENDATIONS
Preanesthesia consultation for elective eye surgery with sedation and monitored anesthesia care
(MAC) includes assessment of ability to communicate, cooperate, and lie supine comfortably.
(See 'Preoperative consultation' above and 'Anesthetic considerations' above.)
Patients with a high risk of clotting and embolic complications who are
receiving aspirin, clopidogrel, or warfarin in therapeutic doses may continue these medications
before cataract surgery with minimal risk of intraocular bleeding. (See 'Preoperative
consultation' above and 'Ophthalmic considerations' above.)
The ophthalmologists measurement of axial eye length on the preoperative ultrasound is
reviewed, since long eyes (axial length >25 mm) have an increased risk of needle injury during
retrobulbar (intraconal) block. (See 'Preoperative consultation' above and 'Ophthalmic
considerations' above.)
Sedatives (eg, midazolam) and opioids (eg, remifentanil) with a short duration are administered to
reduce or eliminate pain during needle insertion and injection of local anesthetic for a regional block.
An alternative technique is administration of 10 to 20 mg increments of propofol until the patient
briefly loses consciousness while the block is performed. (See 'Needle insertion for regional
anesthesia' above.)
Supplemental oxygen is administered to reduce the risk of hypoxemia.
During the surgical procedure itself, sedation is minimized to reduce the risk of side effects that
may jeopardize the patients ability to cooperate during surgery. (See'Surgical procedure' above.)
Compressed room air or an oxygen concentration <30 percent (to minimize fire risk) is delivered at a
rate of 10 L/minute under the surgical drapes to minimize rebreathing of expired carbon dioxide.

Anesthetic techniques for cataract surgery include the topical and regional techniques, as well as
general anesthesia (see 'Anesthesia for cataract surgery' above and 'Anesthetic techniques' above):
Use of topical analgesia as the sole anesthetic technique avoids the potential complications of
both needle insertion and local anesthetic injection. A drawback is that topical analgesia will
not provide akinesia of the eye. (See 'Topical anesthesia' above.)
Regional anesthesia may be provided by peribulbar, retrobulbar, or sub-Tenon block.
Peribulbar (extraconal) regional blocks are most commonly used in the United States.
(See 'Regional anesthesia' above.)
General anesthesia is reserved for patients who cannot communicate or cooperate or who
have severe anxiety or claustrophobia, as well as some patients who are unable to lie supine
comfortably (eg, congestive heart failure, chronic obstructive pulmonary disease, or severe
back pain), and also for children. Either endotracheal intubation with a nondepolarizing
neuromuscular blocking agent or a laryngeal mask airway (LMA) may be used during cataract
surgery. Maintenance of a deep plane of anesthesia is necessary to avoid laryngospasm,
coughing, or other movement. (See 'General anesthesia' above.)
For intraocular glaucoma surgery, anesthetic techniques are the same as those for cataract
surgery. Some surgeons prefer general anesthesia for trabeculectomy; in these cases, complete
akinesia is ensured by titration of a nondepolarizing neuromuscular blocking agent according to
monitoring of muscle relaxation with a peripheral nerve stimulator and maintenance of a deep plane
of anesthesia. (See 'Anesthesia for intraocular glaucoma surgery' above.)
For vitreoretinal surgery (eg, detached retina), a regional anesthetic technique or general
anesthesia (without nitrous oxide) is preferred, rather than topical anesthesia.
Nitrous oxide is contraindicated when injection of gas (eg, SF6 or C3F8) is planned or was
previously used to create a bubble to tamponade a detached retina, unless complete bubble
absorption is documented. (See 'Anesthesia for vitreoretinal surgery' above.)
Serious eye complications of regional blocks include globe or optic nerve perforation, and
retrobulbar hemorrhage. (See 'Anesthetic complications' above.)
Rare serious systemic complications include inadvertent administration of local anesthetic into a
blood vessel or the central nervous system, with resultant cardiorespiratory depression or arrest.
(See 'Anesthetic complications' above.)
The oculocardiac reflex can occur when pressure is applied to extraocular muscles and may result
in bradycardia, junctional rhythms, hypotension, or asystole. Management includes discontinuing
manipulation of the extraocular muscles and administration of an anticholinergic medication
(eg, atropine or glycopyrrolate). (See'Anesthetic complications' above.)
Use of UpToDate is subject to the Subscription and License Agreement.
REFERENCES
1.

The Anesthesia Quality Institute. Anesthesia in the US 2014. https://ecommerce.asahq.org/p-556anesthesia-in-the-united-states-2014.aspx? (Accessed on October 22, 2014).

2.

McGoldrick KE, Gayer SI. Anesthesia for Opthalmologic Surgery. In: Clinical Anesthesia, 7th ed,
Barash PG (Ed), Lippincott Williams & Wilkins, Philadelphia 2013. p.1373.

3.

Fedorowicz Z, Lawrence D, Gutierrez P, van Zuuren EJ. Day care versus in-patient surgery for
age-related cataract. Cochrane Database Syst Rev 2011; :CD004242.

4.

Douketis JD, Berger PB, Dunn AS, et al. The perioperative management of antithrombotic
therapy: American College of Chest Physicians Evidence-Based Clinical Practice Guidelines (8th Edition).
Chest 2008; 133:299S.

5.

Fleisher LA, Fleischmann KE, Auerbach AD, et al. 2014 ACC/AHA Guideline on Perioperative
Cardiovascular Evaluation and Management of Patients Undergoing Noncardiac Surgery: Executive
Summary: A Report of the American College of Cardiology/American Heart Association Task Force on
Practice Guidelines. Circulation 2014; 130:2215.

6.

Jamula E, Anderson J, Douketis JD. Safety of continuing warfarin therapy during cataract
surgery: a systematic review and meta-analysis. Thromb Res 2009; 124:292.

7.

American Society of Anesthesiologists Committee on Standards and Practice Parameters.


Practice alert for the perioperative management of patients with coronary artery stents: a report by the
American Society of Anesthesiologists Committee on Standards and Practice Parameters. Anesthesiology
2009; 110:22.

8.

Ramsay RC, Knobloch WH. Ocular perforation following retrobulbar anesthesia for retinal
detachment surgery. Am J Ophthalmol 1978; 86:61.

9.

Edge R, Navon S. Scleral perforation during retrobulbar and peribulbar anesthesia: risk factors
and outcome in 50,000 consecutive injections. J Cataract Refract Surg 1999; 25:1237.

10.

Vohra SB, Good PA. Altered globe dimensions of axial myopia as risk factors for penetrating
ocular injury during peribulbar anaesthesia. Br J Anaesth 2000; 85:242.

11.

Yang YF, Herbert L, Rschen H, Cooling RJ. Nitrous oxide anaesthesia in the presence of
intraocular gas can cause irreversible blindness. BMJ 2002; 325:532.

12.

Seaberg RR, Freeman WR, Goldbaum MH, Manecke GR Jr. Permanent postoperative vision loss
associated with expansion of intraocular gas in the presence of a nitrous oxide-containing anesthetic.
Anesthesiology 2002; 97:1309.

13.

Lee EJ. Use of nitrous oxide causing severe visual loss 37 days after retinal surgery. Br J
Anaesth 2004; 93:464.

14.

STANDARDS FOR BASIC ANESTHETIC MONITORING. Committee of Origin:Standards and


Practice Parameters (APproved by the ASA House of Delegates on October 21, 1986, and last amended
on October 20, 2010 with an effective date of July 1, 2011 https://www.asahq.org/ForMembers/Standards-Guidelines-and-Statements.aspx (Accessed on January 05, 2015).

15.

Continuum of depth of sedation: Definition of general anesthesia and levels of


sedation/analgesia. Last updated October 21, 2009. American Society of Anesthesiologists (ASA).
Available online at: https://www.asahq.org/~/media/For%20Members/Standards%20and
%20Guidelines/2012/CONTINUUM%20OF%20DEPTH%20OF%20SEDATION%20442012.pdf (Accessed
on September 09, 2014).

16.

Apfelbaum JL, Caplan RA, Barker SJ, et al. Practice advisory for the prevention and management
of operating room fires: an updated report by the American Society of Anesthesiologists Task Force on
Operating Room Fires. Anesthesiology 2013; 118:271.

17.

Leaming DV. Practice styles and preferences of ASCRS members--2002 survey. J Cataract
Refract Surg 2003; 29:1412.

18.

Guise PA. Sub-Tenon anesthesia: a prospective study of 6,000 blocks. Anesthesiology 2003;
98:964.

19.

Eke T, Thompson JR. The National Survey of Local Anaesthesia for Ocular Surgery. II. Safety
profiles of local anaesthesia techniques. Eye (Lond) 1999; 13 ( Pt 2):196.

20.

Seidenari P, Santin G, Milani P, David A. Peribulbar and retrobulbar combined anesthesia for
vitreoretinal surgery using ropivacaine. Eur J Ophthalmol 2006; 16:295.

21.

Ozcan AA, Ozdemir N, Gnes Y, et al. Intraocular pressure, quality of block, and degree of pain
associated with ropivacaine in peribulbar block: a comparative randomized study with bupivacainelidocaine mixture. Eur J Ophthalmol 2003; 13:794.

22.

DeBroff BM, Hamilton RC, Loken RG, et al. Retrobulbar anesthesia with 7.5 vs. 0.75 IU/mL of
hyaluronidase. Can J Ophthalmol 1995; 30:262.

23.

Ripart J, Lefrant JY, Prat-Pradal D, et al. Peribulbar versus retrobulbar anesthesia for ophthalmic
surgery: an anatomical comparison of extraconal and intraconal injections. Anesthesiology 2001; 94:56.

24.

Davis DB 2nd, Mandel MR. Posterior peribulbar anesthesia: an alternative to retrobulbar


anesthesia. J Cataract Refract Surg 1986; 12:182.

25.

Hessemer V. [Peribulbar anesthesia versus retrobulbar anesthesia with facial nerve block.
Techniques, local anesthetics and additives, akinesia and sensory block, complications]. Klin Monbl
Augenheilkd 1994; 204:75.

26.

Alhassan MB, Kyari F, Ejere HO. Peribulbar versus retrobulbar anaesthesia for cataract surgery.
Cochrane Database Syst Rev 2008; :CD004083.

27.

Demirok A, Simek S, Cinal A, Yaar T. Peribulbar anesthesia: one versus two injections.
Ophthalmic Surg Lasers 1997; 28:998.

28.

Agrawal V, Athanikar NS. Single injection, low volume periocular anesthesia in 1,000 cases. J
Cataract Refract Surg 1994; 20:61.

29.

Budd J, Hardwick M, Barber K, Prosser J. A single-centre study of 1000 consecutive peribulbar


blocks. Eye (Lond) 2001; 15:464.

30.

Martin SR, Baker SS, Muenzler WS. Retrobulbar anesthesia and orbicularis akinesia. Ophthalmic
Surg 1986; 17:232.

31.

Grizzard WS, Kirk NM, Pavan PR, et al. Perforating ocular injuries caused by anesthesia
personnel. Ophthalmology 1991; 98:1011.

32.

Hamilton RC, Gimbel HV, Strunin L. Regional anaesthesia for 12,000 cataract extraction and
intraocular lens implantation procedures. Can J Anaesth 1988; 35:615.

33.

Duker JS, Belmont JB, Benson WE, et al. Inadvertent globe perforation during retrobulbar and
peribulbar anesthesia. Patient characteristics, surgical management, and visual outcome. Ophthalmology
1991; 98:519.

34.

Nicoll JM, Acharya PA, Ahlen K, et al. Central nervous system complications after 6000
retrobulbar blocks. Anesth Analg 1987; 66:1298.

35.

Sanchez-Capuchino A, Meadows D, Morgan L. Local anaesthesia for eye surgery without a facial
nerve block. Anaesthesia 1993; 48:428.

36.

Hansen EA, Mein CE, Mazzoli R. Ocular anesthesia for cataract surgery: a direct sub-Tenon's
approach. Ophthalmic Surg 1990; 21:696.

37.

Stevens JD. A new local anesthesia technique for cataract extraction by one quadrant subTenon's infiltration. Br J Ophthalmol 1992; 76:670.

38.

Roman SJ, Chong Sit DA, Boureau CM, et al. Sub-Tenon's anaesthesia: an efficient and safe
technique. Br J Ophthalmol 1997; 81:673.

39.

Lebuisson DA. Simplified and safer peribulbar anaesthesia. Eur J Implant Ref Surg 1990; 2:123.

40.

Allman KG, Theron AD, Byles DB. A new technique of incisionless minimally invasive sub-Tenon's
anaesthesia. Anaesthesia 2008; 63:782.

41.

Tsuneoka H, Ohki O, Osamu T, Kenji K. Tenons capsule anaesthesia for cataract surgery with IOL
implantation. Eur J Implant Ref Surg 1993; 5:29.

42.

Frieman BJ, Friedberg MA. Globe perforation associated with subtenon's anesthesia. Am J
Ophthalmol 2001; 131:520.

43.

Rschen H, Bremner FD, Carr C. Complications after sub-Tenon's eye block. Anesth Analg 2003;
96:273.

44.

Wainwright AC. Positive pressure ventilation and the laryngeal mask airway in ophthalmic
anaesthesia. Br J Anaesth 1995; 75:249.

45.

Lamb K, James MF, Janicki PK. The laryngeal mask airway for intraocular surgery: effects on
intraocular pressure and stress responses. Br J Anaesth 1992; 69:143.

46.

Thomson KD. The effect of the laryngeal mask airway on coughing after eye surgery under
general anesthesia. Ophthalmic Surg 1992; 23:630.

47.

Pablo LE, Ferreras A, Prez-Olivn S, et al. Comparison of the efficacy and safety of contact
versus peribulbar anaesthesia in combined eye surgery. Ophthalmologica 2009; 223:60.

48.

Geffen N, Carrillo MM, Jin Y, et al. Effect of local anesthesia on trabeculectomy success. J
Glaucoma 2008; 17:658.

49.

Eke T. Anesthesia for glaucoma surgery. Ophthalmol Clin North Am 2006; 19:245.

50.

Noureddin BN, Jeffrey M, Franks WA, Hitchings RA. Conjunctival changes after subconjunctival
lignocaine. Eye (Lond) 1993; 7 ( Pt 3):457.

51.

Ghali AM, El Btarny AM. The effect on outcome of peribulbar anaesthesia in conjunction with
general anesthesia for vitreoretinal surgery. Anaesthesia 2010; 65:249.

52.

Sohn HJ, Moon HS, Nam DH, Paik HJ. Effect of volume used in sub-Tenon's anesthesia on
efficacy and intraocular pressure in vitreoretinal surgery. Ophthalmologica 2008; 222:414.

53.

Gild WM, Posner KL, Caplan RA, Cheney FW. Eye injuries associated with anesthesia. A closed
claims analysis. Anesthesiology 1992; 76:204.

54.

Hamilton RC. A discourse on the complications of retrobulbar and peribulbar blockade. Can J
Ophthalmol 2000; 35:363.

55.

Katz J, Feldman MA, Bass EB, et al. Risks and benefits of anticoagulant and antiplatelet
medication use before cataract surgery. Ophthalmology 2003; 110:1784.

56.

Morgan CM, Schatz H, Vine AK, et al. Ocular complications associated with retrobulbar injections.
Ophthalmology 1988; 95:660.

57.

Cap H, Roth E, Johnson T, et al. Vertical strabismus after cataract surgery. Ophthalmology 1996;
103:918.

S-ar putea să vă placă și