Sunteți pe pagina 1din 19

REVIEW ARTICLE

Drugs Aging 2010; 27 (1): 21-38 1170-229X/10/0001-0021/$49.95/0

2010 Adis Data Information BV. All rights reserved.

Anaesthesia for Cataract Surgery


Emmanuel Nouvellon,1 Philippe Cuvillon,1 Jacques Ripart1,2 and Eric J. Viel1
mes, France 1 Anaesthesia Service and Pain Clinic, University Hospital Caremeau, N mes Medical School, Montpellier I University, Montpellier, France 2 Montpellier-N

Contents
Abstract. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. Anaesthesia in the Elderly Patient: Specific Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Concomitant Diseases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Continuation of Usual Treatment or Not? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Requests from the Surgeon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Anatomical Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Regional Anaesthesia (RA): Conventional Blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Complications of Needle Blocks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Retrobulbar Anaesthesia (RBA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Peribulbar Anaesthesia (PBA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 RBA versus PBA Controversy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Recent RA Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 Topical Anaesthesia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Perilimbal (Subconjunctival) Anaesthesia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Sub-Tenons Block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Needle Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.2 Surgical Approach with a Blunt Cannula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Local Anaesthetics and Adjuvant Agents for Eye Blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Controversy: Who Should Perform the Block? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Supporting Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Controversy: General Anaesthesia versus RA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7. Future of Cataract Anaesthesia. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 22 23 23 24 24 25 25 26 26 28 28 28 29 29 30 30 31 32 32 33 34 34

Abstract

Cataract surgery is the most frequent surgical procedure requiring anaesthesia in developed countries. It is performed mainly in elderly patients, who present with many coexisting diseases that induce subsequent hazards from general anaesthesia. Cataract anaesthesia is performed following various techniques of regional anaesthesia, which are detailed in this review. Needle block carries a low but real risk of complications, mainly because of needle misplacement. Correct teaching and training are mandatory to prevent complications. The main patient risk factor for inadvertent globe perforation is the presence of a myopic staphyloma. Retrobulbar block has been progressively phased out and replaced by peribulbar block, sub-Tenons block (STB) or topical anaesthesia (TA). The requirement for very deep block with total akinesia has greatly decreased with the use of phacoemulsification for cataract surgery, allowing for use of TA or low-volume STB. However,

22

Nouvellon et al.

non-akinesia techniques may give rise to impaired surgical conditions, which have the potential to result in surgical complications. A surgical approach to accessing sub-Tenons space avoids needle block, but does not totally prevent complications. When deep anaesthesia is required, low-volume STB, performed using either the needle technique or a surgical approach, appears to be the technique of choice in terms of efficacy. Increasing the anaesthetic volume provides reproducible akinesia. Various local anaesthetics may be used, depending on their availability and respective properties. The most useful adjuvant to local anaesthetic is hyaluronidase.

Ophthalmic surgery is the most frequent surgical procedure requiring anaesthesia in developed countries. Each year, cataract surgery is performed on more than 300 000 patients in the UK, about 500 000 patients in France and nearly 2 million patients in the US.[1] The vast majority of procedures are performed in a 1-day surgery setting, under regional anaesthesia (RA), also called eye block.[1-3] Eye block has historically been limited to retrobulbar anaesthesia (RBA) performed by the surgeon alone with monitored anaesthesia care or without any anaesthesiologists assistance at all. Anaesthesiologists are now increasingly becoming involved in ophthalmic RA. As surgical practice evolves following widespread use of the phacoemulsification (PKE) technique, requests from surgeons for total akinesia and lowered intraocular pressure (IOP) have decreased.[4] At the same time, the complications of conventional RBA have been extensively described, and the need for greater safety during eye block has been emphasized. This has resulted in the development of alternative techniques such as peribulbar, lowvolume sub-Tenons block (STB) and topical anaesthesia (TA), the aim of which is to improve safety, albeit at the price of imperfect akinesia. However, some surgeons still express a need for a more efficient block, providing total globe akinesia and anaesthesia of the globe, for some difficult procedures and certain patients. A greater knowledge of anatomy and of the various techniques will enable the anaesthesiologist to choose the best technique to match each situation. Whether anaesthesia influences patient outcomes has been the topic of many publications, and these cannot be extensively detailed here. The
2010 Adis Data Information BV. All rights reserved.

controversy over the choice of anaesthesia technique i.e. general anaesthesia (GA) versus RA suffers from a lack of well designed comparative studies. Moreover, non-ophthalmic complications of cataract surgery are so rare that the outcome does not significantly differ between those patients who have been operated on and those who have not. This review briefly discusses anaesthesia in elderly patients and the condition of the eye requested by the surgeon, but focuses particularly on RA, i.e. the relevant anatomy, classical (retrobulbar and peribulbar) needle block techniques together with their efficacies and complications, emerging techniques and their relative pros and cons, choice of local anaesthetics and adjuvant agents and, finally, some ongoing controversies. In addition, the debate between GA and RA is developed. Literature searches were conducted on MEDLINE and EMBASE from 2000 to 2009 using the following keywords: cataract surgery, phakoemulsification, intraocular lens replacement, eye surgery, anesthesia, regional anesthesia, general anesthesia, and elderly patient. We also searched a cumulative personal database based on monthly issues of the following journals published since 1994: Anesthesiology, Regional Anesthesia and Pain Medicine, Anesthesia & Analgesia, British Journal of Anaesthesia, Journal of Cataract & Refractive Surgery, Ophthalmology, American Journal of Ophthalmic Surgery, Eye and Archives of Ophthalmology. 1. Anaesthesia in the Elderly Patient: Specific Considerations The topic of anaesthesia in the elderly patient is sufficiently large in itself as to require a separate
Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

23

review, which can be found elsewhere (e.g. Cohendy et al.[5]). This discussion is limited to the most important points.
1.1 Concomitant Diseases

1.2 Continuation of Usual Treatment or Not?

Cataract is mainly an elderly patients health concern. In a series of >55 000 cases, the median patient age was 75 years.[6] As in all elderly patients, cataract patients may have multiple concomitant diseases. Kelly observed that up to 63% of cataract patients have a significant underlying disease.[7] Stupp et al. observed a 97% rate of pre-existing risk factors for intraoperative adverse events requiring medical intervention in patients undergoing cataract surgery.[8] The most frequently described co-morbidity is cardiovascular disease, but neurological or metabolic diseases or conditions, including diabetes mellitus or chronic corticosteroid therapy, are also reported. In a prospective study, Sharwood et al. reported the following incidences: hypertension (51%), gastrooesophageal reflux (19%), angina and myocardial infarction (14%), diabetes (13%) and asthma (9%).[9] Indeed, diabetes and corticosteroid therapy are classical causes of cataracts. The elderly have also specific characteristics that may interfere with anaesthetic agents, particularly GA, to produce exaggerated adverse effects:  Decreased drug elimination may lead to accumulation, relative overdosage and subsequently delayed recovery or prolonged residual effect with mental dysfunction.  Increased susceptibility to anaesthetic agents may lead to more pronounced drug effects than in healthy adults with the same dosage (e.g. deeper hypotension induced by GA or delayed wakening).  Effects of anaesthesia such as arterial hypotension may be less well tolerated in the elderly patient because of multiple underlying diseases (e.g. myocardial or cerebral infarction resulting from coronary or carotid stenosis).  Polymedication may lead to significant drug interactions with anaesthetic agents, especially in patients given GA.
2010 Adis Data Information BV. All rights reserved.

As with any other surgical procedure, the golden rule is not to change a patients usual treatment if he or she is stable on it. This is especially true for antihypertensive agents, and moreover for b-adrenoceptor antagonists (b-blockers) in particular, because of the risk of a rebound effect.[10] Therefore, the patients usual treatment should be given on the morning of surgery. Concerning insulin treatment of diabetic patients, when the patient is fasted, two options are possible: (i) as for any other surgery, because the patient is fasting, the normal insulin dose is given as usual under the cover of an intravenous glucose infusion (glucose 50 g/L, 1 mL/g/h); or (ii) the morning insulin dose is avoided. When the patient is not fasting, the usual insulin is continued. The most debated problem is whether to continue anticoagulants/antiplatelet therapy. On the one hand, continuing such treatments is frequently thought to increase both the risk for surgical bleeding and the risk of haematoma from eye block puncture. On the other hand, discontinuing those therapies may unacceptably increase the risk for thrombosis in at-risk patients, for example, those with atrial fibrillation, non-tissue prosthetic heart valves, carotid stenosis or coronary drug-eluting stents. These risks include cerebral ischaemic stroke, myocardial infarction and death. In fact, cataract surgery is a totally bloodless procedure and can be performed safely in patients taking any anticoagulant/antiplatelet agents.[11] So these drugs should generally be continued. When choosing the eye block, the option of TA must be taken into account. All deep blocks (both needle block and cannula STB) carry a low risk for haematoma. Most of them are benign with only short-term aesthetic consequences (i.e. subconjunctival or palpebral haematoma), the exception being the classical but very infrequent compressive retrobulbar haemorrhage.[12,13] However, it appears that anticoagulants/antiplatelets do not significantly increase the risk of such a haemorrhage.[14] Therefore, discontinuation of anticoagulant/antiplatelet therapy for a deep block is no longer routinely recommended. British guidelines state that, the INR [international normalized ratio] should be within the therapeutic ratio which is determined by the
Drugs Aging 2010; 27 (1)

24

Nouvellon et al.

condition for which the patient is being anticoagulated,[15] and French guidelines recommend continuation of aspirin (acetylsalicylic acid).[16] Canadian practice appears to be similar.[17] The only drug for which there are no recommendations is clopidogrel. In this case, a decision must be taken on an individual basis. 2. Requests from the Surgeon Classically, for cataract surgery, as with any other open globe surgery, the request from the surgeon is for analgesia, akinesia and hypotonia of the eyeball. This remains true for the traditional extracapsular cataract extraction (ECE) technique, which requires a large (68 mm) incision to remove the entire lens. By contrast, with the development of PKE, methods have radically changed.[4] This procedure consists of fragmentation of the lens nucleus with an ultrasound probe and evacuation of the fragments by an irrigation-aspiration system. A small incision (34 mm) is sufficient to allow a foldable prosthetic lens to be introduced. Because the small incision is self-sealing, there is no need for any suture and the duration of the procedure is reduced to as little as 35 minutes in easy cases. As the small incision is obturated by the ultrasound probe, PKE can be considered a closed eye procedure, meaning lowering the IOP is much less important. Akinesia is rarely if ever required, because the ultrasound probe allows immobilization of the eyeball. Finally, as the lens is free of any sensory innervation, only the corneal incision may be painful, and this only rarely. Indeed, PKE is possible for selected patients without any anaesthesia at all.[18] The surgeon may also require that other general conditions be prevented. Acute peak arterial hypertension, for instance, may cause catastrophic choroidal expulsive haemorrhage. Tremor or restlessness may impair the procedure, for obvious reasons. Coughing gives rise to very acute and high peak IOP, which can impair surgery and therefore must be prevented. 3. Anatomical Considerations The orbit is a cavity in the shape of a truncated square pyramid, with its apex posterior and its
2010 Adis Data Information BV. All rights reserved.

base corresponding to the anterior aperture. The orbit is filled mainly by adipose tissue, and the globe is suspended in its anterior part. The four rectus muscles of the eye are inserted anteriorly near the equator of the globe. Posteriorly, they are inserted together, at the apex, on the Zinn tendinous annulus, through which the optic nerve enters the orbit. The four rectus muscles delimit the retrobulbar cone, which is not sealed by any intermuscular membrane.[19,20] Sensory innervation of the globe is supplied by the ophthalmic nerve, the first branch of the trigeminal nerve (cranial nerve [CN] V), which passes through the muscular cone. The trochlear nerve (CN IV) provides the motor command to the superior oblique muscles, the abducens nerve (CN VI) does the same to the lateral rectus muscle, and the oculomotor nerve (CN III) provides motor signals to all the other extraocular muscles. All these nerves except the trochlear nerve pass through the muscular conus. Therefore, injecting local anaesthetic inside the cone can logically be expected to provide anaesthesia and akinesia of the globe and of the extraocular muscles. Only the motor command of the orbicularis muscle of the eyelids has an extraorbital course, arising as it does from the superior branch of the facial nerve (CN VII). Many major structures are located in the muscular conus and are therefore vulnerable to the risk of needle injury; these include the optic nerve with its meningeal sheaths, most of the arteries of the orbit, and the autonomic, sensory and motor innervation of the globe. The facial sheath of the eyeball also called Tenons capsule is a fibroelastic layer that surrounds the entire scleral portion of the globe. It delimits the episcleral space or sub-Tenons space, a potential space with no actual volume, although fluid can be injected into it. Some authors assimilate it into the articular capsule of the globe. Near the equator, Tenons capsule is perforated by the tendons of the oblique and rectus muscles before they insert into the sclera. At this point there is a continuity between Tenons capsule and the fascial sheath of the muscles. Anteriorly, Tenons capsule merges with the bulbar conjunctiva before both insert together into the corneal limbus.
Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

25

4. Regional Anaesthesia (RA): Conventional Blocks As all needle blocks can be used for cataract surgery, the choice of needle is largely based on efforts to prevent complications.
4.1 Complications of Needle Blocks

The most common cause of needle block complications is needle misplacement. Although some anatomical features may increase the risk of complications, the main risk factor is poor training and limited experience of the physician performing the block. Spread of anaesthesia in the CNS may involve two mechanisms. Firstly, an inadvertent intraarterial injection may reverse blood flow in the ophthalmic artery up to the anterior cerebral or the internal carotid artery,[21] such that an injected volume as small as 4 mL can produce seizures. In such cases, symptomatic treatment (oxygen supply, injected antiepileptic drugs such as thiopental sodium or benzodiazepine, face mask ventilation, and, if required, tracheal intubation for mechanical ventilation with muscle relaxants) usually allows rapid recovery without after-effects. Secondly, an inadvertent injection under the dura mater sheath of the optic nerve or directly through the optic foramen may result in subarachnoid spread of the local anaesthetic. This causes partial or total, progressive, brainstem anaesthesia.[22,23] Katsev et al. have demonstrated that the apex of the orbit may be reached with a 40 mm long needle.[24] Depending on the dose and volume of local anaesthetic spreading towards the brainstem, possibilities include bilateral block, CN palsy with sympathetic activation, confusion and restlessness, or total spinal anaesthesia with tetraparesis, arterial hypotension, bradycardia and eventually respiratory and cardiac arrest. Symptomatic treatment (oxygen supply, vasopressors, and, if required, tracheal intubation and ventilation) should result in total recovery within hours. Inadvertent globe perforation and rupture is the most devastating complication of eye blocks. The complication has a poor prognosis, particu 2010 Adis Data Information BV. All rights reserved.

larly in cases of delayed diagnosis. The incidence is between 1/350 and 7/50 000 cases.[25,26] Risk factors are classically inexperience of the physician and a highly myopic eye (i.e. long eyeball).[26] In a series of 50 000 cases, Edge and Navon observed that myopic staphyloma was the greatest risk factor for scleral perforation.[25] This suggests that isolated high myopia may not be a risk factor per se, but acts as a confounding factor, as myopic staphyloma occurs only in myopic eyes. Using ultrasound imaging, Vohra and Good observed that the probability of staphyloma is greater in highly myopic than slightly myopic eyes.[27] Moreover, staphyloma is more frequently located at the posterior pole of the globe (accounting for perforations after RBA) or in the inferior area of the globe (accounting for perforations after inferior and temporal punctures, both peri- or retrobulbar). As a result, at least in myopic patients and ideally in all patients, ultrasound measurement of the axial length of the globe (biometry) should be performed. A highly myopic eye (axial length >26 mm) remains the classical contraindication to eye block. However, this contraindication may be circumvented if B-mode ultrasound is conducted to assess the presence and location of a staphyloma. Injury to an extraocular muscle may cause diplopia and ptosis. Several mechanisms can be involved: direct injury by the needle resulting in intramuscular haematoma, high pressure due to injection into the muscle sheath, or myotoxicity of the local anaesthetic.[28] The injury may progress in three steps: initially the muscle is paralysed, then it appears to recover, and finally, a retractile scar develops. Retrobulbar haemorrhage results from an inadvertent arterial puncture. This complication may lead to a compressive haematoma, which can threaten retinal perfusion. Surgical decompression may be required, but, in most cases, all that is needed is postponement of surgery.[12] The main risk factor is arterial fragility (diabetes and atheroma), rather than clotting disorders. Venous puncture leads to noncompressive haematoma, the consequences of which are much less severe, so that, in most cases, surgery can be continued.
Drugs Aging 2010; 27 (1)

26

Nouvellon et al.

Direct optic nerve trauma by the needle is very rare but causes blindness. CT scan imaging usually shows optic nerve enlargement due to intraneural haematoma.[29]
4.2 Retrobulbar Anaesthesia (RBA)

Historically, RBA has been the gold standard of eye block and is achieved by injecting a small volume of local anaesthetic agent (35 mL) inside the muscular cone.[30] The main hazard of RBA is the risk of injury to the globe or one of the many vulnerable elements located in the muscular cone. Near the apex, these structures are packed in a very small volume and are fixed by the tendon of Zinn, which prevents them from moving away from the needle. The resulting potential complications are detailed in section 4.1. To prevent such complications, some authors have proposed avoiding introduction of a needle into the muscular cone, and prefer to keep the needle in the extraconal space, a technique that is theoretically less hazardous.[31,32] Although used from the beginning of the twentieth century, RBA was formally described only in 1936.[30] An additional facial nerve block is required to prevent blinking, with the technique that is most frequently used being the van Lindt eyelid block.[33] The Atkinson up and in position of the gaze was abandoned after Liu et al.[34] and Unso ld et al.[35] confirmed that this increases the risk of optic nerve injury. RBA is used less frequently nowadays because of its complications.
4.3 Peribulbar Anaesthesia (PBA)

Fig. 1. Conventional peribulbar injection: (a) inferior and temporal injection and (b) superior and nasal injection.

The long-used technique of peribulbar anaesthesia (PBA) was highlighted by the work of Bloomberg,[32] Davis and Mandel.[31,36] The main reason for use of this technique is to reduce the risk of injury to major structures in the intraconal space. PBA consists of introducing the needle into the extraconal space, which appears to be less hazardous. The volume of local anaesthetic injected is larger than that with an RBA injection (up to12 mL). A large volume is required to allow the local anaesthetic to spread into the entire corpus adiposum of the orbit, including the intraconal space, where the nerves to be blocked are
2010 Adis Data Information BV. All rights reserved.

located. Additionally, such a large volume allows an anterior spread to the eyelids, providing a block of the orbicularis muscle of the eyelids, thereby avoiding the need for additional eyelid block. The most classical PBA technique involves two injections, one inferiorly and temporally at the same site as for RBA, and the second superiorly and nasally (figure 1). Many alternative techniques have been described and these cannot be extensively reported here. However, advances in PBA techniques may be summarized in terms of a few guidelines: 1. Using a single injection technique. As the space where the local anaesthetic spreads is unique,
Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

27

increasing the injected volume is sufficient to provide efficient anaesthesia. Comparative studies have confirmed that, provided the injected volume is sufficient, the single injection technique is as effective as the double injection technique.[37] Moreover, as the first injection may impair identification of anatomical landmarks, it has been suggested that the second injection may lead to complications more frequently than the first.[38] A second injection should be performed only as a supplement when the first injection has failed. 2. Limiting the depth of needle insertion (usually 25 mm). Posteriorly to the globe, the rectus muscles are in contact with the orbital walls, so that the extraconal space totally disappears and becomes virtual. Increasing the depth of needle insertion is expected to change a peribulbar into a retrobulbar injection.[39] Some posterior PBAs are in fact inadvertent retrobulbar injections. This fact probably explains the occurrence of complications such as optic nerve injury after an attempted peribulbar injection. Moreover, a long needle introduced totally into the orbit may reach the apex of the orbit, a hazardous zone.[40] In an anatomical study, Katsev et al. demonstrated that inserting the needle up to a 40 mm depth will result in an injection directly through the optical foramen in 11% of cases.[24] More recently, limiting the needle insertion to a 15 mm depth was proposed; this has a similar efficacy to a 25 mm depth.[41] 3. Avoiding the superior and nasal site of puncture. At this level, the distance between the orbital roof and the globe is reduced, theoretically increasing the risk of globe perforation. Additionally, the superior oblique muscle may be injured by the needle. Inferior and temporal puncture remains the gold standard. An alternative site of puncture for PBA is the medial canthus (figure 2).[42] The needle is introduced at the medial junction of the eyelids, nasally to the lachrymal caruncle, in a strictly posterior direction, at a depth of 15 mm (no more). At this level, the space between the orbital wall and the globe is as large as with the inferior and temporal approach, and is free from blood
2010 Adis Data Information BV. All rights reserved.

2 3 1

Fig. 2. Sites of needle introduction for the most frequently used blocks: (1) medial canthus peribulbar anaesthesia;[42] (2) lachrymal caruncle; (3) semilunaris fold of the conjunctiva; (4) medial canthus episcleral anaesthesia;[43-79] and (5) inferior and temporal peribulbar anaesthesia.[31,32]

vessels. Moreover, myopic staphyloma, an anatomical anomaly that represents a risk factor for perforation, is infrequently located on the nasal side of the globe.[25,27] 4. Using a thin needle (25 gauge) to limit pain. Use of short-bevelled needles should presumably enhance safety by increasing the tactile perception of resistance to be overcome during needle insertion. For example, shortbevelled needles need more pressure to perforate the sclera on cadavers.[43] However, because of the very low incidence of perforation, this has not been confirmed in patients. 5. Using compression to lower IOP, which increases after injection. Compression has not been shown to enhance the quality of the block. A 30 mmHg pressure applied for 1015 minutes is sufficient. In all cases, spreading of local anaesthetics in the corpus adiposum of the orbit remains somewhat unpredictable, leading to the need to increase the injected volume to prevent an imperfect block. Depending on the surgeons request for akinesia, an additional injection is required in 150% of cases.[31,32,36,44] This poor reproducibility in block efficacy is the main disadvantage of PBA.
Drugs Aging 2010; 27 (1)

28

Nouvellon et al.

4.4 RBA versus PBA Controversy

Although RBA is classically assumed to be more efficacious than PBA, it appears that provided a sufficient volume of local anaesthetic is injected both techniques have similar efficacies.[45] This can be explained by the fact that there is no intermuscular membrane that separates extraconal from intraconal spaces, so that both form a unique space for spreading of local anaesthetic (figure 3).[19,20] If efficacy is similar, the only basis on which to choose one technique over the other is safety. RBA, theoretically, carries a higher risk of complications (optic nerve injury, brainstem anaesthesia, retrobulbar haemorrhage) because of intraconal introduction of the needle. However, the expected greater safety of PBA has never been confirmed. This may be because of the very low rate of complications and the subsequent lack of power of comparative studies including large series or meta-analyses.[46,47] In our opinion, the similar efficacy of the two techniques and the possibly higher risk of complications with RBA compared with PBA should result in discontinuation of use of RBA. 5. Recent RA Techniques
5.1 Topical Anaesthesia

Instillation of local anaesthetic eyedrops provides corneal analgesia, thus allowing cataract surgery by PKE when akinesia is not required.[4] It is quick and simple to perform and avoids the potential hazards of needle techniques. However, the rate of complications with RA is so low that it was never possible to confirm a decrease in the RA-related sight-threatening complication rate with TA, except when compared with RBA.[46] As it does not require deep insertion of a sharp needle, TA is the technique of choice for patients receiving anticoagulant or antiplatelet therapy.[48] This technique is increasingly being used for cataract surgery worldwide, accounting for up to 60% of procedures in some series but remaining at around 20% in some others.[1-3,46,49] Some surgeons prefer TA for routine PKE in >90% of their procedures. However, the efficacy
2010 Adis Data Information BV. All rights reserved.

Fig. 3. Schematic of spread of local anaesthetic for (a) both retrobulbar and peribulbar block or (b) sub-Tenons injection. The figures show coronal sections passing through the globe. In (a), it is important to note that local anaesthetic spreads through the whole corpus adiposum of the orbit regardless of whether the injection is peribulbar or retrobulbar anaesthesia. Inhomogeneous spreading because of small septa explains why retrobulbar and peribulbar anaesthesia may produce incomplete blocks. By contrast, in (b), subTenons injection corresponds to a very specific space that guides the local anaesthetic all around the globe, thus explaining the very reproducible anaesthesia of the globe. Because there is a continuation between Tenons capsule and the muscle sheath, when the injected volume is increased (>4 mL), excess local anaesthetic flows towards the rectus muscles, thereby explaining the very good akinesia (reproduced from Ripart et al.,[80] with permission from Lippincott Williams & Wilkins, http://lww.com).

Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

29

of TA is limited. Firstly, analgesia may be incomplete. Patients randomly subjected to RBA or TA for one eye and the other technique for the other eye preferred RBA (71% vs 10%).[50] Intraoperative comfort is more consistently obtained with RBA[50-52] or sub-Tenons[53,81] than with TA, which appears to be no more effective than no anaesthesia at all in selected cases involving an experienced surgeon.[18] Secondly, the lack of akinesia and IOP control associated with the short duration of the procedure may theoretically make surgery more hazardous.[52,54] Only one study has observed an advantage for TA in terms of the surgical complication rate. In a randomized, non-blind, comparative study of unselected patients, Jacobi et al. observed only one significant difference between TA and RBA, namely, a surprising decrease in vitreous issue rate in the TA group (0.4% vs 2.5%).[55] At the same time, the investigators observed a nonsignificant increase in iris prolapse (1.7% vs 0.4%), possibly reflecting eye hypertonia due to the lack of akinesia in the TA group. In a 2382case survey, Shaw et al. observed an acceptably low rate of surgical complications of cataract surgery performed under TA.[56] By contrast, a more recent meta-analysis has shown that, when compared with STB, TA is associated with a 2-fold increase in posterior capsule rupture requiring anterior vitrectomy.[54] Similarly, TA was identified as a risk factor for displacement of nuclear fragments into the vitreous.[57] Therefore, TA should be limited to planned easy procedures performed by experienced surgeons in selected patients. For manual ECE, akinesia is still required and TA is questionable.[4] This may be the case in institutions where PKE is not available for technical reasons.[58,59] Efforts have been made to improve TA efficacy in many ways. Use of long-acting local anaesthetics such as levobupivacaine or ropivacaine appears more efficacious than lidocaine (lignocaine).[60,61] Intracameral injection of local anaesthetic has been proposed to enhance analgesia.[62] This entails injecting small volumes (0.1 mL) of local anaesthetic in the anterior chamber at the beginning of surgery. The safety of this technique in relation to local anaesthetic
2010 Adis Data Information BV. All rights reserved.

toxicity to corneal endothelium, which is not able to regenerate, has been confirmed,[63,64] but any significant analgesic benefit of intracameral injection versus simple TA has never been established by properly designed trials.[18,64-68] This is not surprising, as analgesia is not correlated with intracameral local anaesthetic concentration.[69] For these reasons, intracameral injection cannot be recommended. The efficacy of sponges soaked with local anaesthetic inserted into the conjunctival fornices and soluble local anaesthetic inserts needs further documentation.[70,71] Instilling lidocaine jelly instead of eyedrops appears to clearly enhance the quality of analgesia of the anterior segment,[53,69,72] and is being increasingly used. However, lidocaine jelly has been associated with an increase in postoperative incidence of endophthalmitis, as described by the French sanitary agency in 2004.[73] This might explain why TA has been associated with a 3.8-fold increase in endophthalmitis rate compared with RBA.[74] The most plausible explanation for these cases is that if the jelly was applied first on the eye, it would have acted as a barrier, preventing disinfectant applied later from reaching the conjunctiva, thereby resulting in insufficient eye disinfection. Thus, the problem is probably the wrong sequence of application rather than the jelly itself. Specifically designed topical lidocaine or tetracaine jellies should replace current jellies, which were originally designed for use in urology.[75,76]
5.2 Perilimbal (Subconjunctival) Anaesthesia

Subconjunctival injection of local anaesthetic may provide analgesia of the anterior segment without any akinesia,[77] but has not gained wide popularity.
5.3 Sub-Tenons Block

Also called episcleral or parabulbar anaesthesia, STB is achieved by injecting into the episcleral space. This allows the local anaesthetic to spread circularly around the scleral portion of the globe, thereby achieving high-quality analgesia of the whole globe with injection of relatively
Drugs Aging 2010; 27 (1)

30

Nouvellon et al.

low volumes (usually 35 mL) [figure 3].[78,79] In addition, use of a larger volume (up to 811 mL) means the local anaesthetic will spread to the extraocular muscle sheaths, producing an effective and reproducible akinesia (figure 3).[44,78,79,82,83] Chemosis (subconjunctival spread of the local anaesthetic) occurs frequently after injection of such large volumes[81,83,84] and is easily explained by anatomical features. Its occurrence confirms the sub-Tenons location of the injection and may require compression to be resolved. Sub-Tenons space may be injected using either a needle technique or surgical dissection followed by introduction of a blunt cannula.
5.3.1 Needle Technique

In the needle technique,[44,78,79,83] the needle is introduced into the fornix between the semilunaris fold of the conjunctiva and the globe, tangentially to the globe (figure 2). After it has encroached on the conjunctiva, the needle is slightly shifted medially and advanced strictly posteriorly, thereby pulling on the globe, which results in directing the gaze medially. At a 1015 mm depth, after a small loss of resistance (click) is perceived, the globe returns to its primary gaze position. This serves as a depth marker, thus allowing injection. Using a large volume (510 mL) with this technique results in good globe and lid akinesia, and the results are more reproducible than with classical PBA.[44] This technique is simple to learn and perform, with acceptable safety of use. In our experience of 2000 cases, we encountered no serious complications.[83] However, as with all needle techniques, the risk of misplacement of the needle and its subsequent complications must be kept in mind. This technique has not been widely used outside France. A variant technique using only forceps and a pencil-point needle instead of scissors and cannula was recently proposed by Allman et al.[85] This technique might be considered as a blend of needle and cannula techniques. It remains to be assessed on a large scale.
5.3.2 Surgical Approach with a Blunt Cannula

Use of a blunt cannula was first proposed as an intraoperative complement to RBA.[86] This tech 2010 Adis Data Information BV. All rights reserved.

nique has subsequently been proposed as a sole anaesthetic technique,[87] and has been used in up to 50% of cases in the UK.[3] Under TA, the bulbar conjunctiva is grasped with a small forceps, 510 mm away from the limbus. Blunt Wescott scissors are used to open a small buttonhole into the conjunctiva and Tenons capsule to gain access to the episcleral space. A blunt cannula is then inserted into the episcleral space to allow the injection. Several types of cannulas have been proposed: smooth curved metallic cannulas, ultrashort metallic cannulas, and silicon or plastic cannulas.[88,89] When no specific cannula is available, a short intravenous catheter (18 or 20 gauge) without its needle can be used. This technique is usually used with injection of low volumes (25 mL) of local anaesthetic, which provides good globe analgesia but only partial akinesia of the globe and lids. To obtain acceptable akinesia, the injected volume must be increased to 11 mL.[82] Being a non-akinesia technique, lowvolume STB carries the same limitations as TA. Indeed, it has been identified as a risk factor for displacement of nuclear fragments into the vitreous.[57] Another limitation is the relatively high rate of minor incidents, with a 2.3-fold increase compared with RBA and PBA having been observed.[3] In a 6000-case series, Guise reported that 6% of cases had chemosis and 7% had subconjunctival haematoma, with only one case requiring cancellation of surgery.[84] However, STB efficacy is excellent for globe analgesia: 96% of the blocks were scored as perfect or good. Moreover, use of small volumes causes a very small increase in IOP, such that preoperative compression of the globe may be unnecessary. Similarly, episcleral injection of a small volume of local anaesthetic may be used for an open globe. Therefore, it is the technique of choice as a supplemental injection when required intraoperatively. Finally, the main advantage of the technique is its safety, because it avoids the blind introduction of a needle in the orbit. Some complications of STB, although rare, have been described. Nevertheless, despite its proven safety record in large series, any needle block complication may occur after STB, including strabismus, eyeball perforation and sepsis.[13] One inadvertent perforation
Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

31

occurred during dissection, possibly due to synechias between the sclera and Tenons capsule in a previously operated eye.[90] However, in large series, complications of STB are very rare. Guise reported no serious complication in 6000 cases.[84] In the series of 375 000 procedures by Eke and Thompson, STB appeared to cause a lower incidence of serious complications than RBA and PBA, although this difference was significant only for life-threatening (not sight-threatening) complications.[46] In the British database series of >55 000 cases, STB was associated with a 2.5-fold decreased risk in serious complications compared with needle techniques.[3] Sub-Tenons anaesthesia also leads to higher patient and surgeon satisfaction than RBA and TA,[91,92] but is only slightly superior to PBA in this respect.[93]
5.4 Local Anaesthetics and Adjuvant Agents for Eye Blocks

All available local anaesthetics have been used for eye block, either alone or as a mixture of two agents. Many publications have compared various local anaesthetic mixtures and concentrations, and these cannot be detailed here. The choice of local anaesthetics should be based on the pharmacological properties and availability of the drugs, taking into account particularly the requirement for quick onset (lidocaine, mepivacaine, articaine), prolonged effect, postoperative residual block for analgesia (ropivacaine, bupivacaine) and akinesia (higher concentration). The most frequently used local anaesthetics are lidocaine, bupivacaine, ropivacaine and mepivacaine, or a combination of two of these.[94] The main differences found between these local anaesthetic combinations are in accordance with their known pharmacological properties. Articaine, usually used more by dental surgeons than anaesthesiologists, has also been proposed for both PBA and STB because of its quick onset and short duration.[95,96] Articaine is also considered to cause less myotoxicity and diplopia than other drugs.[97] There is no definitive magic bullet in terms of the ideal local anaesthetic for eye block. As the volume of local anaesthetic injected is usually small (310 mL), systemic toxicity is not a major concern.
2010 Adis Data Information BV. All rights reserved.

Hyaluronidase is an enzyme that facilitates wider spreading of local anaesthetics.[98] The only significant, although very rare, complication of hyaluronidase administration is immediate or delayed allergy. This may cause orbital oedema or a pseudotumoural orbitopathy, with a subsequent rise in IOP that may mimic an expulsive choroidal haemorrhage[99] or lead to surgical complications (posterior capsule rupture or vitreous loss).[100-104] As it is probably not due to hyaluronidase itself, this problem should be prevented by more effective purification[105] or, ideally, by use of recombinant hyaluronidase, which is available in the US. Use of hyaluronidase is classically assumed to shorten the onset of the block and enhance its quality.[106] However, the literature is somewhat controversial, showing a limited magnitude of benefit concerning akinesia.[107] This may be because of use of varying concentrations in different studies. Doses varying from 3.75 to 300 IU/mL of local anaesthetic have been reported, with only a tendency towards a greater benefit with higher doses.[108] Given the wide range of concentrations used and the absence of clear dose-efficacy ranging data, 2550 IU/mL can be proposed. Use of hyaluronidase also permits a decrease in the amount of local anaesthetic required to achieve the same efficacy.[109] Other benefits of hyaluronidase include a smaller intraoperative increase in IOP, which may otherwise dramatically hinder surgery.[108] The last consideration in terms of the benefits of hyaluronidase is its ability to decrease the incidence of postoperative strabismus, possibly by limiting local anaesthetic myotoxicity because of a quicker spread.[110,111] Other adjuvants are less frequently used. Clonidine, for instance, enhances intraoperative and postoperative analgesia when added to local anaesthetic. At a dose of 0.51 mg/kg, clonidine does not increase the incidence of systemic adverse events such as hypotension or excessive sedation.[112] Moreover, clonidine may help to prevent intraoperative arterial hypertension and may lower IOP. Adrenaline (epinephrine) is sometimes added to increase the duration of eye block. However, postoperative pain is not a major concern after cataract surgery and the
Drugs Aging 2010; 27 (1)

32

Nouvellon et al.

availability of a new long-acting local anaesthetic has become the focus of less interest. Alkalinization has been proposed to decrease pain during injection and to accelerate block onset, but has limited efficacy. The optimal concentration of bicarbonate is difficult to determine because of the large pH range of local anaesthetic solutions provided by pharmaceutical firms. Moreover, the local anaesthetic may precipitate when there is excessive alkalinization.[113] Small doses of muscle relaxant may enhance akinesia, but concern has been expressed about their potential risk for systemic effects.[114] Opioids do not appear to be more efficient when administered via a regional ophthalmic route than by systemic administration.[115] Warming the local anaesthetic may decrease pain on injection and enhance block efficacy, but no clinically relevant benefits appear to occur.[116]
5.5 Controversy: Who Should Perform the Block?

anaesthesiologists have demonstrated their ability to perform eye blocks, as with other RA techniques.[3,42,83,84] However, anaesthesiologists are not available for eye block everywhere. The cost effectiveness of this practice may vary among various institutions, depending on the payment system.
5.6 Supporting Therapies

Since the 1980s, anaesthesiologists have become increasingly involved in eye blocks that were previously undertaken by surgeons. However, there are very great differences in anaesthesiologists involvement in eye blocks both between and within countries. In some institutions, anaesthesiologists are not available, and surgeons have to manage the block themselves.[117,118] In other areas, anaesthesiologists provide only monitored anaesthesia care while the surgeon performs the block. Finally, in many countries, such as France and the UK, anaesthesiologists are increasingly performing the blocks. Some reports have emphasized complications after blocks performed by anesthesia personnel.[119] In fact, some of these complications were associated with blocks performed by nurse anaesthetists or operating room nurses directly supervised by surgeons. It must be stressed that eye blocks are, like any other RA, relatively safe but potentially dangerous. Anaesthesiologists should theoretically be the most appropriate persons to perform eye blocks, provide monitored anaesthesia care and manage life-threatening complications. Provided that they have been correctly taught and trained,
2010 Adis Data Information BV. All rights reserved.

Eye surgery, especially cataract surgery, has very little impact on perioperative morbidity and mortality.[120,121] Eye block is associated with lower perioperative morbidity than GA for ophthalmic surgery, provided that no heavy sedation is added.[121-125] As a result, some standard safety measures such as monitoring, fasting and preoperative evaluation, which are considered basic recommendations for other types of block, are sometimes circumvented for eye blocks.[126] The role of preoperative fasting remains controversial. When complications occur or there is a need to convert to GA, a full stomach may constitute an aggravating factor. On the other hand, strict fasting is uncomfortable and may be considered too heavy-handed given the very low incidence of such situations. Practices appear to vary widely among different countries. Old patients undergoing eye surgery frequently have coexisting disease. Therefore, a preoperative assessment to ensure coexisting diseases are properly controlled appears well advised. Premedication is frequently used but should be utilized cautiously so as not to prevent early discharge because of residual excessive sedation in an ambulatory setting. The most frequently used drugs include benzodiazepines, hydroxyzine, clonidine, dexmedetomedine and melatonin.[127-130] Antibacterial prophylaxis is limited to sepsis highrisk situations such as in the treatment of diabetic patients, in which situation the antibacterial will be chosen for its intraocular penetration properties. A fluoroquinolone is a standard choice.[131] Short sedation for performing the block might include small amounts of propofol or opioids.[44,83] Care must be taken in patients with mental confusion and restlessness attributable to benzodiazepines or any other drug overdosage. Excessive sedation and
Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

33

subsequent mental confusion have been claimed to have caused patient movement during performance of RA with subsequent eyeball perforation.[132] Because perioperative systemic adverse events of supporting therapies are so rare that their incidence cannot be measured, some physicians advise against any intraoperative monitoring or monitored anaesthesia care.[133] However, given the underlying co-morbidities in many patients, adverse events are possible.[134] Even if their incidence is low, in the context of the very large amount of cataract surgery performed each year, the absolute value of such events is probably significant. In the authors opinion, as in French practice, intraoperative monitoring should include basic monitoring, i.e. ECG, pulse oximetry and automated noninvasive blood pressure measurement. Various British guidelines are less strict and recommend only continuous monitoring of ventilation and circulation by clinical observation and pulse oximetry.[15,135] In patients with life-threatening complications, the ability to perform resuscitation is essential, and the same guidelines recommend that there should always be at least one person present who has Advanced Life Support (ALS) training or equivalent. Intravascular access is clearly required. Patient immobility is required during surgery. Anxiety and residual pain occur relatively frequently during eye surgery under local anaesthesia, and the presence of drapes over the head may increase anxiety. Therefore, the patient should be positioned as comfortably as possible, with sufficient space to allow free breathing. Additional fresh air flow is preferable to oxygen as a means of improving ventilation in a confined atmosphere. Intraoperative sedation can be used to limit anxiety and pain, in addition to or instead of premedication. However, heavy sedation has been associated with an increase in complications such as restlessness or hypoxia, and has even been associated with increased mortality.[122] Excessive sedation can be avoided by cautious use of sedatives (if needed), which are titrated before the patient is draped. The most frequently used drugs are small doses of propofol or remifentanil infusion.[136,137] Continuous infusion of dexmedetomedine or midazolam has also been proposed,
2010 Adis Data Information BV. All rights reserved.

without enjoying widespread use.[138] To avoid excessive sedation, patient-controlled or patienttarget-controlled infusion have been proposed, again without achieving great popularity.[137,139] 6. Controversy: General Anaesthesia versus RA Anterior segment ophthalmic surgery is the only human surgical procedure that has no detectable impact on patient survival.[123] There is a lack of well conducted prospective large studies comparing GA and RA for cataract surgery. Light planes of GA may lead to eye divergency with subsequent eccentric eye position and impaired surgical access to the globe.[140] Postoperative cognitive dysfunction (POCD) in the elderly patient remains an unsolved problem after GA for major surgery and is a predictive factor for mortality.[141] Both anaesthesia duration and the technique used for postoperative analgesia (particularly avoidance of RA, i.e. continuous epidural or peripheral nerve blocks) may play a role in increasing POCD.[5] This is an argument for avoiding GA for cataract surgery, which typically is totally painless postoperatively. Most cataract surgeries are performed under RA, a choice based on the assumption that this is less hazardous to elderly patients. Although this assumption seems logical for many, it is based on very weak evidence or none at all. In a 20-year noncomparative survey, Quigley was not able to demonstrate any difference in mortality rate between GA and RA.[124] Similarly, Hoskings et al. were not able to show any difference in short-term mortality and morbidity between GA and RA in elderly patients (age >90 years).[125] However, in this series, patients age was greater in the RA group, possibly because physicians more frequently chose elderly and disabled patients, being convinced that RA is safer for fragile patients. Only Glantz et al. observed fewer myocardial ischaemic events after RA than after GA.[121] However, trying to impose use of RA in all cases is probably pointless. Neurological diseases and psychiatric disorders (e.g. deafness, Parkinsons disease or mental confusion) may impair patient cooperation and the ability to lie
Drugs Aging 2010; 27 (1)

34

Nouvellon et al.

supine quietly during surgery under RA. Similarly, rheumatic pain or prostate adenoma dysuria may prevent the patient from lying immobile under RA. The required time of immobility is around 15 minutes, but may be longer, depending on the surgeons skill and the characteristics of the cataract. Sedation is frequently presented as a solution but excessive sedation may lead to restlessness, sleeping and snoring, or to respiratory depression, which, in the absence of any airway access, may be catastrophic. In cases where intraoperative immobility cannot be guaranteed, GA might be considered as a first choice, rather than RA combined with hazardous heavy sedation. 7. Future of Cataract Anaesthesia Cataract anaesthesia in the elderly is achieved mainly by RA or TA. Research should continue the quest for the Holy Grail that will provide analgesia (PKE) and total akinesia when required (other surgical techniques), enabling procedures to be carried out in absolute safety. In terms of efficacy, STB appears to be the gold standard. However, most anaesthesiologists are unfamiliar with cannula techniques, which require two-hand experience. Because of safety considerations, needle techniques will probably continue to lose popularity in the future. TA prevents anaesthesia complications but some concerns may still be expressed over surgical difficulties caused by the absence of akinesia. Specific local anaesthetic jelly mixtures for TA should be developed. Hyaluronidase, although useful, is not available in many countries, and further efforts should be made to make it including the recombinant form more available. 8. Conclusion Cataract surgery is the most frequent surgical procedure requiring anaesthesia in developed countries. Over the last 20 years, anaesthesiologists have played an increasing role in performing eye blocks. RBA is being progressively replaced by PBA. The requirement for very deep block with total akinesia has greatly decreased with the use of PKE for cataract surgery, expanding the
2010 Adis Data Information BV. All rights reserved.

place for TA or low-volume STB. Needle block carries a low but real risk of complications, mainly due to needle misplacement. Correct teaching and training are required to prevent complications. The main patient risk factor for inadvertent globe perforation is the presence of a myopic staphyloma. Use of a surgical approach to gain access to sub-Tenons space avoids needle block, but does not totally prevent complications. When deep anaesthesia is required, STB, performed either by the needle technique or using a surgical approach, appears to be the technique of choice. Acknowledgements
No sources of funding were used to assist in the preparation of this review. Jacques Ripart has acted as a consultant to , France. The AstraZeneca, France, and Air Liquide Sante other authors have no conflicts of interest that are directly relevant to the content of this review. The authors acknowledge Serge Albertini for his English editing of the article.

References
1. Leaming DV. Practice styles and preferences of ASCRS members: 2003 survey. J Cataract Refract Surg 2004; 30: 892-900 2. Pick ZS, Leaming DV, Elder MJ. The fourth New Zealand cataract and refractive surgery survey: 2007. Clin Experiment Ophthalmol 2008; 36: 604-19 3. El-Hindy N, Johnston RL, Jaycock P, et al., UK EPR User Group. The Cataract National Dataset electronic multicentre audit of 55 567 operations: anaesthetic techniques and complications. Eye 2009; 23: 50-5 4. Schutz JS, Mavrakanas NA. What degree of anaesthesia is necessary for intraocular surgery? It depends on whether surgery is open or closed. Br J Ophthalmol. In press 5. Cohendy R, Brougere A, Cuvillon P. Anaesthesia in the older patient. Curr Opin Clin Nutr Metab Care 2005; 8: 17-21 6. Narendran N, Jaycock P, Johnston RL, et al. The Cataract National Dataset electronic multicentre audit of 55 567 operations: risk stratification for posterior capsule rupture and vitreous loss. Eye 2009; 23: 31-7 7. Kelly JM. Preoperative assessment and medication. In: Mostafa SM, editor. Anaesthesia for ophthalmic surgery. New York: Oxford University Press, 1991: 95-105 8. Stupp T, Hassouna I, Soppart K, et al. Systemic adverse events: a comparison between topical and peribulbar anaesthesia in cataract surgery. Ophthalmologica 2007; 221: 320-5 9. Sharwood PL, Thomas D, Roberts TV. Adverse medical events associated with cataract surgery performed under

Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

35

10.

11.

12. 13. 14.

15.

16.

17.

18.

19.

20.

21.

22.

23.

24. 25.

26.

topical anaesthesia. Clin Experiment Ophthalmol 2008; 36: 842-6 Landoni G, Zambon M, Zangrillo A. Reducing perioperative myocardial infarction with anesthetic drugs and techniques. Curr Drug Targets. In press Jonas JB, Pakdaman B, Sauder G. Cataract surgery under systemic anticoagulant therapy with coumarin. Eur J Ophthalmol 2006; 16: 30-2 Edge KR, Nicoll MV. Retrobulbar hemorrhage after 12 500 retrobulbar blocks. Anesth Analg 1993; 76: 1019-22 Ru schen H, Bremner FD, Carr C. Complications after subTenons eye block. Anesth Analg 2003; 96: 273-7 Kumar N, Jivan S, Thomas P, et al. Sub-Tenons anesthesia with aspirin, warfarin, and clopidogrel. J Cataract Refract Surg 2006; 32: 1022-5 The Royal College of Anaesthetists and the Royal College of Ophthalmologists. Local anaesthesia for intraocular surgery. London: Royal College of Anaesthetists, 2001 [online]. Available from URL: http://www.rcoa.ac.uk/ docs/RCARCOGuidelines.pdf [Accessed 2009 Sep 29] te Francaise dAnesthe sie et de Re animation et Socie te Socie matologie. Antiplatelet agents and the periFrancaise dHe operative period, 2001 [online]. Available from URL: http:// www.sfar.org/pdf/aapconfexp.pdf [Accessed 2009 Sep 29] Ong-Tone L, Paluck EC, Hart-Mitchell RD. Perioperative use of warfarin and aspirin in cataract surgery by Canadian Society of Cataract and Refractive Surgery members: survey. J Cataract Refract Surg 2005; 31: 991-6 Pandey SK, Werner L, Apple DJ, et al. No-anesthesia clear corneal phacoemulsification versus topical and topical plus intracameral anesthesia: randomized clinical trial. J Cataract Refract Surg 2001; 27: 1643-50 Koornneef L. Details of the orbital connective tissue in the adult: the architecture of the musculo-fibrous apparatus in the human orbit. Acta Morphol Neerl-Scand 1977; 15: 1-64 Ripart J, Lefrant JY, de La Coussaye JE, et al. Peribulbar versus retrobulbar anesthesia for ophthalmic surgery: an anatomical comparison of extraconal and intraconal injections. Anesthesiology 2001; 94: 56-62 Aldrete JA, Romo-Salas F, Arora S, et al. Reverse arterial blood flow as a pathway for central nervous system toxic responses following injection of local anesthetics. Anesth Analg 1978; 57: 428-33 Nicoll JMV, Acharya PA, Ahlen K, et al. Central nervous system complication after 6000 retrobulbar blocks. Anesth Analg 1987; 66: 1298-302 Loken RG, Mervyn Kirker GE, Hamilton RC. Respiratory arrest following peribulbar anesthesia for cataract surgery: case report and review of the literature. Can J Ophthalmol 1998; 33: 225-6 Katsev DA, Drews RC, Rose BT. Anatomic study of retrobulbar needle path length. Ophthalmology 1989; 96: 1221-4 Edge R, Navon S. Scleral perforation during retrobulbar and peribulbar anesthesia: risk factor and outcome in 50 000 consecutive injections. J Cataract Refract Surg 1999; 25: 1237-44 Duker JS, Belmont JB, Benson WE, et al. Inadvertent globe perforation during retrobulbar and peribulbar anesthesia. Ophthalmology 1991; 98: 519-26

27. 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: 242-5 28. Carlson BM, Rainin EA. Rat extraocular muscle regeneration: repair of local anesthetic-induced damage. Arch Ophthalmol 1985; 103: 1373-7 29. Hersch M, Baer G, Diecker JP, et al. Optic nerve enlargement and central retinal-artery occlusion secondary to retrobulbar anesthesia. Ann Ophthalmol 1989; 21: 195-7 30. Atkinson WS. Retrobulbar injection of anesthetic within the muscular cone (cone injection). Arch Ophthalmol 1936; 16: 495-503 31. Davis DB, Mandel MR. Posterior peribulbar anesthesia: an alternative to retrobulbar anesthesia. J Cataract Refract Surg 1986; 12: 182-4 32. Bloomberg LB. Administration of periocular anesthesia. J Cataract Refract Surg 1986; 12: 677-9 brale transitoire provoque e 33. Van Lindt M. Paralysie palpe ration de la cataracte. Ann Ocul 1914; 151: dans lope 420-4 34. Liu C, Youl B, Moseley I. Magnetic resonance imaging of the optic nerve in the extremes of gaze: implications for the positioning of the globe for retrobulbar anaesthesia. Br J Ophthalmol 1992; 76: 728-33 35. Unso ld R, Stanley JA, Degroot J. The CT topography of retrobulbar anesthesia: anatomical correlation of implications and suggestion of a modified technique. Albrecht Von Graefes Arch Klin Exp Ophthalmol 1981; 217: 125-36 36. Davis DB, Mandel MR. Efficacy and complication rate of 16 224 consecutive peribulbar blocks: a prospective multicenter study. J Cataract Refract Surg 1994; 20: 327-37 37. Demirok A, Simsek S, Cinal A, et al. Peribulbar anesthesia: one versus two injections. Ophthalmic Surg Lasers 1997; 28: 998-1001 38. Ball JL, Woon WH, Smith S. Globe perforation by the second peribulbar injection. Eye 2002; 16: 663-5 39. Sarvela J, Nikki P. Comparison of two needle lengths in regional ophthalmic anesthesia with etidocaine and hyaluronidase. Ophthalmic Surg 1992; 23: 742-5 40. Karampatakis V, Natsis K, Gisgis P, et al. The risk of optic nerve injury in retrobulbar anesthesia: a comparative study on 35 and 40 mm retrobulbar needles in 12 cadavers. Eur J Ophthalmol 1998; 8: 184-7 41. Riad W, Ahmed N. Single injection peribulbar anesthesia with a short needle combined with digital compression. Anesth Analg 2008; 107: 1751-3 42. Hustead RF, Hamilton RC, Loken RG. Periocular local anesthesia: medial orbital as an alternative to superior nasal injection. J Cataract Refract Surg 1994; 20: 197-201 43. Waller SG, Taboada J, OConnor P. Retrobulbar anesthesia risk: do sharp needles really perforate the eye more easily than blunt needles? Ophthalmology 1993; 100: 506-10 44. Ripart J, Lefrant JY, Vivien B, et al. Ophthalmic regional anesthesia: canthus episcleral anesthesia is more efficient than peribulbar anesthesia: a double blind randomized study. Anesthesiology 2000; 92: 1278-85 45. Demediuk OM, Dhaliwal RS, Papworth DP, et al. A comparison of peribulbar and retrobulbar anesthesia

2010 Adis Data Information BV. All rights reserved.

Drugs Aging 2010; 27 (1)

36

Nouvellon et al.

46.

47.

48.

49.

50.

51.

52.

53.

54.

55.

56. 57.

58.

59.

60.

61.

62.

for vitreoretinal surgical procedures. Arch Ophthalmol 1995; 113: 908-13 Eke T, Thompson JR. Serious complications of local anaesthesia for cataract surgery: a 1 year national survey in the United Kingdom. Br J Ophthalmol 2007; 91: 470-5 Alhassan MB, Kyari F, Ejere HO. Peribulbar versus retrobulbar anaesthesia for cataract surgery. Cochrane Database Syst Rev 2008; 16 (3): CD004083 Barequet IS, Sachs D, Priel A, et al. Phacoemulsification of cataract in patients receiving coumadin therapy: ocular and hematologic risk assessment. Am J Ophthalmol 2007; 144: 719-72 Wagle AA, Wagle AM, Bacsal K, et al. Practice preferences of ophthalmic anaesthesia for cataract surgery in Singapore. Singapore Med J 2007; 48: 287-90 Boezaart A, Berry R, Nell M. Topical anesthesia versus retrobulbar block for cataract surgery: the patients perspective. J Clin Anesth 2000; 12: 58-60 Rebolleda G, Munoz-Negrete FJ, Gutierrez-Ortiz C. Topical plus intracameral lidocaine versus retrobulbar anesthesia in phacotrabeculectomy: prospective randomized study. J Cataract Refract Surg 2001; 27: 1214-20 Gombos K, Jakubovits E, Kolos A, et al. Cataract surgery anaesthesia: is topical anaesthesia really better than retrobulbar? Acta Ophthalmol Scand 2007; 85: 309-16 Sekundo W, Dick HB, Schmidt JC. Lidocaine-assisted xylocaine jelly anesthesia versus one quadrant sub-Tenon infiltration for self-sealing sclero-corneal incision routine phacoemulsification. Eur J Ophthalmol 2004; 14: 111-6 Davison M, Padroni S, Bunce C, et al. Sub-Tenons anaesthesia versus topical anaesthesia for cataract surgery. Cochrane Database Syst Rev 2007; 18 (3): CD006291 Jacobi PC, Dietlein TS, Jacobi FK. A comparative study of topical vs retrobulbar anesthesia in complicated cataract surgery. Arch Ophthalmol 2000; 118: 1037-43 Shaw AD, Ang GS, Eke T. Phacoemulsification complication rates. Ophthalmology 2007; 114: 2101-2 Mahmood S, von Lany H, Cole MD, et al. Displacement of nuclear fragments into the vitreous complicating phacoemulsification surgery in the UK: incidence and risk factors. Br J Ophthalmol 2008; 92: 488-92 Waddell KM, Reeves BC, Johnson GI. A comparison of anterior and posterior chamber lenses after cataract extraction in rural Africa: a within patient randomized trial. Br J Ophthalmol 2004; 88: 734-9 Bourne RR, Minassian DC, Dart JK, et al. Effect of cataract surgery on the corneal endothelium: modern phacoemulsification compared with extracapsular surgery. Ophthalmology 2004; 11: 679-85 Borazan M, Karalezli A, Akova YA, et al. Comparative clinical trial of topical anaesthetic agents for cataract surgery with phacoemulsification: lidocaine 2% drops, levobupivacaine 0.75% drops, and ropivacaine 1% drops. Eye 2008; 22: 425-9 ndez SA, Dios E, Diz JC. Comparative study of toFerna pical anaesthesia with lidocaine 2% vs levobupivacaine 0.75% in cataract surgery. Br J Anaesth 2009; 102: 216-20 Karp CL, Cox TA, Wagoner MD, et al. Intracameral anesthesia: a report by the American Academy of Ophthalmology. Ophthalmology 2001; 108: 1704-10

63. Heuerman T, Hartman C, Anders N. Long term endothelial cell loss after phacoemulsification: peribulbar anesthesia versus intracameral lidocaine 1%: prospective randomized study. J Cataract Refract Surg 2002; 28: 638-43 64. Boulton JE, Lopatazidis A, Luck J, et al. A randomized controlled trial of intracameral lidocaine during phacoemulsification under topical anesthesia. Ophthalmology 2000; 107: 68-71 65. Roberts T, Boytell K. A comparison of cataract surgery under topical anaesthesia with and without intracameral lignocaine. Clin Experiment Ophthalmol 2002; 30: 19-22 66. Pang MP, Fujimoto DK, Wilkens LR. Pain, photophobia, and retinal and optic nerve function after phacoemulsification with intracameral lidocaine. Ophthalmology 2001; 108: 2018-25 67. Ezra DG, Nambiar A, Allan BD. Supplementary intracameral lidocaine for phacoemulsification under topical anesthesia: a meta-analysis of randomized controlled trials. Ophthalmology 2008; 115: 455-87 68. Ezra DG, Allan BD. Topical anaesthesia alone versus topical anaesthesia with intracameral lidocaine for phacoemulsification. Cochrane Database Syst Rev 2007 Jul 18; (3): CD005276 69. Bardocci A, Lofoco G, Perdicaro S, et al. Lidocaine 2% gel versus lidocaine 4% unpreserved drops for topical anesthesia in cataract surgery: a randomized controlled trial. Ophthalmology 2003; 110: 144-9 70. Aziz ES. Deep topical fornix nerve block versus peribulbar block in one-step adjustable suture horizontal strabismus surgery. Br J Anaesth 2002; 88: 129-32 I, Mouly S, Jarrin I, et al. Efficacy and safety of three 71. Mahe ophthalmic inserts for topical anaesthesia of the cornea: an exploratory comparative dose-ranging, double-blind, randomized trial in healthy volunteers. Br J Clin Pharmacol 2005; 59: 220-6 72. Barequet IS, Soriano ES, Green WR, et al. Provision of anesthesia with single application of lidocaine gel. J Cataract Refract Surg 1999; 25: 626-31 curite Sanitaire et des Produits 73. Agence Francaise pour la Se (AFSSAPS): Me susage de Xyloca ne 2%, gel de Sante tral en seringue pre -remplie: endophtalmies rapporte es ure lors du traitement chirurgical de la cataracte. Vigilances 2004 Oct; (2): 2 [online]. Available from URL: http:// www.afssaps.fr/var/afssaps_site/storage/original/applica tion/9095a6c4be02687266bd4b042dad2229.pdf [Accessed Sep 29] 74. Garcia-Arumi J, Fonollosa A, Sararols L, et al. Topical anesthesia: possible risk factor for endophthalmitis after cataract extraction. J Cataract Refract Surg 2007; 33: 989-92 75. Perone JM, Popovici A, Ouled-Moussa R, et al. Safety and efficacy of two ocular anesthetic methods for phacoemulsification: topical anesthesia and viscoanesthesia (VisThesia). Eur J Ophthalmol 2007; 17: 171-7 76. Amiel H, Koch PS. Tetracaine hydrochloride 0.5% versus lidocaine 2% jelly as a topical anesthetic agent in cataract surgery: comparative clinical trial. J Cataract Refract Surg 2007; 33: 98-100

2010 Adis Data Information BV. All rights reserved.

Drugs Aging 2010; 27 (1)

Anaesthesia for Cataract Surgery

37

77. Cagini C, De Carolis A, Fiore T, et al. Limbal anaesthesia versus topical anaesthesia for clear corneal phacoemulsification. Acta Ophthalmol Scand 2006; 84: 105-9 78. Ripart J, Prat-Pradal D, Charavel P, et al. Medial canthus single injection episcleral (sub-Tenon) anesthesia anatomic imaging. Clin Anat 1998; 11: 390-5 79. Ripart J, Metge L, Prat-Pradal D, et al. Medial canthus single injection episcleral (sub-Tenon) anesthesia computed tomography imaging. Anesth Analg 1998; 87: 43-5 80. Ripart J, Benbabaali M, LHermitte J, et al. Ophthalmic blocks at the medial canthus (reply). Anesthesiology 2001; 95: 1533-5 81. Zafirakis P, Voudouri A, Rowe S, et al. Topical versus subTenons anesthesia without sedation in cataract surgery. J Cataract Refract Surg 2001; 27: 873-9 82. Li HK, Abouleish A, Grady J, et al. Sub-Tenons injection for local anesthesia in posterior segment surgery. Ophthalmology 2000; 107: 41-7 83. Nouvellon E, LHermite J, Chaumeron A, et al. Ophthalmic regional anesthesia: medial canthus episcleral (subTenons) single injection block. Anesthesiology 2004; 100: 370-4 84. Guise P. SubTenons anesthesia: a prospective study of 6000 blocks. Anesthesiology 2003; 98: 964-8 85. Allman KG, Theron AD, Byles DB. A new technique of incisionless minimally invasive sub-Tenons anaesthesia. Anaesthesia 2008; 63: 782-3 86. Mein CE, Flynn HW. Augmentation of local anesthesia during retinal detachment surgery. Arch Ophthalmol 1989; 107: 1084 87. Stevens JD. A new local anaesthesia technique for cataract extraction by one quadrant sub-Tenons infiltration. Br J Ophthalmol 1992; 76: 670-4 88. Kumar CM, Mac Neela BJ. Ultrasonic localization of anaesthetic fluids using sub Tenons cannulae of three different lengths. Eye 2003; 17: 1-5 89. MacNeela BJ, Kumar CM. Sub-Tenons block using ultrashort cannula. J Cataract Refract Surg 2004; 30: 858-62 90. Friedman BJ, Friedberg MA. Globe perforation associated with sub-Tenons anesthesia. Am J Ophthalmol 2001; 131: 520-1 91. Rodrigues PA, Vale PJ, Cruz LM, et al. Topical anesthesia versus sub-Tenon block for cataract surgery: surgical conditions and patient satisfaction. Eur J Ophthalmol 2008; 18: 356-60 92. Ryu JH, Kim M, Bahk JH, et al. A comparison of retrobulbar block, sub-Tenon block, and topical anesthesia during cataract surgery. Eur J Ophthalmol 2009; 19: 240-6 93. Budd JM, Brown JP, Thomas J, et al. A comparison of subTenons with peribulbar anaesthesia in patients undergoing sequential bilateral cataract surgery. Anaesthesia 2009; 64: 19-22 94. Borazan M, Karalezli A, Oto S, et al. Comparison of a bupivacaine 0.5% and lidocaine 2% mixture with levobupivacaine 0.75% and ropivacaine 1% in peribulbar anaesthesia for cataract surgery with phacoemulsification. Acta Ophthalmol Scand 2007; 85: 844-7 95. Ozdemir M, Ozdemir G, Zencirci B, et al. Articaine versus lidocaine plus bupivacaine for peribulbar anaesthesia in cataract surgery. Br J Anaesth 2004; 92: 231-4

96. Gouws P, Galloway P, Jacob J, et al. Comparison of articaine and bupivacaine/lidocaine for sub-Tenons anaesthesia in cataract extraction. Br J Anaesth 2004; 92: 228-30 97. Raman SV, Barry JS, Murjaneh S, et al. Comparison of 4% articaine and 0.5% levobupivacaine/2% lidocaine mixture for sub-Tenons anaesthesia in phacoemulsification cataract surgery: a randomised controlled trial. Br J Ophthalmol 2008; 92: 496-9 98. Khandwala M, Ahmed S, Goel S, et al. The effect of hyaluronidase on ultrasound-measured dispersal of local anaesthetic following sub-Tenon injection. Eye 2008; 22: 1065-8 99. Minning CA. Hyaluronidase allergy simulating expulsive choroidal hemorrhage. Arch Ophthalmol 1994; 112: 585-6 100. Agrawal A, Mclure HA, Dabbs TR. Allergic reaction to hyaluronidase after a peribulbar injection [letter]. Anaesthesia 2003; 58: 814-5 101. Etesse B, Beaudroit L, Deleuze M, et al. Hyaluronidase: here we go again. Ann Fr Anesth Reanim 2009; 28: 658-65 102. Kumar CM, Dowd TC, Dodds C, et al. Orbital swelling following peribulbar and sub-Tenons anaesthesia. Eye 2004; 18: 418-20 103. Quhill F, Bowling B, Packard RB. Hyaluronidase allergy after peribulbar anesthesia with orbital inflammation. J Cataract Refract Surg 2004; 30: 916-7 104. Ahluwalia HS, Lukaris A, Lane CM. Delayed allergic reaction to hyaluronidase: a rare sequel to cataract surgery. Eye 2003; 17: 263-6 105. Eberhart AH, Weiler CR, Erie JC. Angioedema related to the use of hyaluronidase in cataract surgery. Am J Ophthalmol 2004; 138: 142-3 106. Remy M, Pinter F, Nentwich MM, et al. Efficacy and safety of hyaluronidase 75 IU as an adjuvant to mepivacaine for retrobulbar anesthesia in cataract surgery. J Cataract Refract Surg 2008; 34: 1966-9 107. Alwitryy A, Chaudhary S, Gopee K, et al. Effect of hyaluronidase on ocular motility in sub Tenons anesthesia: randomized controlled trial. J Cataract Refract Surg 2002; 28: 1420-3 108. Dempsey GA, Barrett PJ, Kirby IJ. Hyaluronidase and peribulbar block. Br J Anaesth 1997; 78: 671-4 109. Schulenburg HE, Sri-Chandana C, Lyons G, et al. Hyaluronidase reduces local anaesthetic volumes for subTenons anaesthesia. Br J Anaesth 2007; 99: 717-20 110. Brown SM, Coats DK, Collins MLZ, et al. Second cluster of strabismus cases after periocular anesthesia without hyaluronidase. J Cataract Refract Surg 2001; 27: 1872-5 111. Hamada S, Devys JM, Xuan TH, et al. Role of hyaluronidase in diplopia after cataract surgery. Ophthalmology 2005; 112: 879-82 112. Bharti N, Madan R, Kaul HL, et al. Effect of addition of clonidine to local anaesthetic mixture for peribulbar block. Anaesth Intens Care 2002; 30: 438-41 113. Hinshaw KD, Fiscella R, Sugar J. Preparation of pHadjusted local anesthetics. Ophthalmic Surg 1995; 26: 194-9 114. Ku cu kyavuz Z, Arici MK. Effects of atracurium added to local anesthetics on akinesia in peribulbar block. Reg Anesth Pain Med 2002; 27: 487-90

2010 Adis Data Information BV. All rights reserved.

Drugs Aging 2010; 27 (1)

38

Nouvellon et al.

115. Hemmerling TM, Budde WM, Koppert W, et al. Retrobulbar versus systemic application of morphine during titratable regional anesthesia via retrobulbar catheter in intraocular surgery. Anesth Analg 2000; 91: 585-8 116. Krause M, Weindler J, Ruprecht KW. Does warming of anesthetic solutions improve analgesia and akinesia in retrobulbar anesthesia? Ophthalmology 1997; 104: 429-32 117. Hansen TE. Practice styles and preferences of Danish cataract surgeons: 1995 survey. Acta Ophthalmol Scand 1996; 74: 56-9 118. Royal College of Anaesthetists Guidance on the provision of ophthalmic anaesthesia services [online]. Available from URL: http://www.rcoa.ac.uk/docs/GPAS-Ophth.pdf [Accessed 2009 Sep 29] 119. Grizzard WS, Kirk NM, Pavan PR, et al. Perforating ocular injuries caused by anesthesia personnel. Ophthalmology 1991; 98: 1011-6 120. Eke T, Thompson JR. The national survey of local anesthesia for ocular surgery II: safety profile of local anesthesia techniques. Eye 1999; 13: 196-204 121. Glantz L, Drenger B, Gozal Y. Perioperative myocardial ischemia in cataract surgery patients: general vs local anesthesia. Anesth Analg 2000; 91: 1415-9 122. Katz J, Feldman MA, Bass EB. Adverse intraoperative medical events and their association with anesthesia management strategies in cataract surgery. Ophthalmology 2001; 108: 1721-6 123. Edwards AE, Seymour DG, McCarthy JM, et al. A 5-year survival study of general surgical patients aged 65 years and over. Anaesthesia 1996; 51: 3-10 124. Quigley HA. Mortality associated with ophthalmic surgery: a 20-year experience at the Wilmer institute. Am J Ophthalmol 1974; 77: 517-24 125. Hoskings MP, Warner MA, Lobdell CM, et al. Outcomes of surgery in patients 90 years of age and older. JAMA 1989; 261: 1909-15 126. Eke T, Thompson JR. The national survey of local anesthesia for ocular surgery: survey methodology and current practice. Eye 1999; 13: 189-95 127. Filos KS, Patroni O, Goudas LC, et al. A dose-response study of orally administered clonidine as premedication in the elderly: evaluating hemodynamic safety. Anesth Analg 1993; 77: 1185-92 128. Barioni MF, Lauretti GR, Lauretti-FO A, et al. Clonidine as a coadjuvant in eye surgery: comparison of peribulbar vs oral administration. J Clin Anaesth 2002; 14: 140-5 129. Erdurmus M, Aydin B, Usta B, et al. Patient comfort and surgeon satisfaction during cataract surgery using topical

130.

131.

132.

133.

134. 135.

136.

137.

138.

139.

140.

141.

anesthesia with or without dexmedetomidine sedation. Eur J Ophthalmol 2008 May-Jun; 18 (3): 361-7 Ismail SA, Mowafi HA. Melatonin provides anxiolysis, enhances analgesia, decreases intraocular pressure, and promotes better operating conditions during cataract surgery under topical anesthesia. Anesth Analg 2009; 108: 1146-51 te Francaise dAnesthe sie et de Re animaation. ReSocie commandations pour la pratique de lantibioprophylaxie en chirurgie [online]. Available from URL: http://www. sfar.org/antibiofr.html [Accessed 2009 Sep 29] Schaack E, Diallo B, Devys JM. Inadvertent globe perforation during peribulbar anaesthesia and sedation with propofol. Ann Fr Anaesth Reanim 2006; 25: 43-5 Rocha G, Turner C. Safety of cataract surgery under topical anesthesia with oral sedation without anesthetic monitoring. Can J Ophthalmol 2007; 42: 288-94 Quantock CL, Goswami T. Death potentially secondary to sub-Tenons block. Anaesthesia 2007; 62: 175-7 The Royal College of Ophthalmologists. Commissioning cataract surgery: an outline of good practice. 2004 [online]. Available from URL: http://www.rcophth.ac.uk/docs/ publications/published-guidelines/CommissioningCataract Surgery-April2004.pdf [Accessed 2009 Sep 29] Yun MJ, Oh AY, Kim KO, et al. Patient-controlled sedation vs anaesthetic nurse-controlled sedation for cataract surgery in elderly patients. Int J Clin Pract 2008; 62: 776-8 Akgu l A, Aydin ON, Dayanir V, et al. Usage of remifentanil and fentanyl in intravenous patient-controlled sedo-analgesia. Agri 2007; 19: 39-46 Alhashemi JA. Dexmedetomidine vs midazolam for monitored anaesthesia care during cataract surgery. Br J Anaesth 2006; 96: 722-6 Pac-Soo CK, Deacock S, Lockwood G, et al. Patientcontrolled sedation for cataract surgery using peribulbar block. Br J Anaesth 1996; 77: 370-4 Rossiter JD, Wood M, Lockwood A, et al. Operating conditions for ocular surgery under general anaesthesia: an eccentric problem. Eye 2006; 20: 55-8 Steinmetz J, Christensen KB, Lund T, et al., and the ISPOCD Group. Long-term consequences of postoperative cognitive dysfunction. Anesthesiology 2009; 110: 548-55

partement d Correspondence: Prof. Jacques Ripart, De sie-Douleur, Centre Hospitalier Universitaire de Anesthe mes, Ho pital Caremeau, Place du Pr Debre , 30029 N mes, N CEDEX 09, France. E-mail: jacques.ripart@chu-nimes.fr

2010 Adis Data Information BV. All rights reserved.

Drugs Aging 2010; 27 (1)

Copyright of Drugs & Aging is the property of ADIS International Limited and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.

S-ar putea să vă placă și