13
http://ajs.sagepub.com/ Medicine The American Journal of Sports http://ajs.sagepub.com/content/37/6/1223 The online version of this article can be found at: DOI: 10.1177/0363546509335947 2009 37: 1223 originally published online May 5, 2009 Am J Sports Med Daniel S. Heckman, George S. Gluck and Selene G. Parekh Tendon Disorders of the Foot and Ankle, Part 2 : Achilles Tendon Disorders Published by: http://www.sagepublications.com On behalf of: American Orthopaedic Society for Sports Medicine can be found at: The American Journal of Sports Medicine Additional services and information for http://ajs.sagepub.com/cgi/alerts Email Alerts: http://ajs.sagepub.com/subscriptions Subscriptions: http://www.sagepub.com/journalsReprints.nav Reprints: http://www.sagepub.com/journalsPermissions.nav Permissions: at UNIV OF OTTAWA LIBRARY on December 22, 2010 ajs.sagepub.com Downloaded from

Tendon Disorders of Foot and Ankle Part 2- Achilles

Embed Size (px)

DESCRIPTION

2009 37: 1223 originally published online May 5, 2009 Am J Sports Med http://ajs.sagepub.com/content/37/6/1223 Published by: The online version of this article can be found at: On behalf of: American Orthopaedic Society for Sports Medicine http://www.sagepublications.com can be found at: The American Journal of Sports Medicine Additional services and information for http://www.sagepub.com/journalsPermissions.nav Permissions: http://www.sagepub.com/journalsReprints.nav Reprints:

Citation preview

Page 1: Tendon Disorders of Foot and Ankle Part 2- Achilles

http://ajs.sagepub.com/Medicine

The American Journal of Sports

http://ajs.sagepub.com/content/37/6/1223The online version of this article can be found at:

 DOI: 10.1177/0363546509335947

2009 37: 1223 originally published online May 5, 2009Am J Sports MedDaniel S. Heckman, George S. Gluck and Selene G. Parekh

Tendon Disorders of the Foot and Ankle, Part 2 : Achilles Tendon Disorders  

Published by:

http://www.sagepublications.com

On behalf of: 

  American Orthopaedic Society for Sports Medicine

can be found at:The American Journal of Sports MedicineAdditional services and information for     

  http://ajs.sagepub.com/cgi/alertsEmail Alerts:

 

http://ajs.sagepub.com/subscriptionsSubscriptions:  

http://www.sagepub.com/journalsReprints.navReprints:  

http://www.sagepub.com/journalsPermissions.navPermissions:  

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 2: Tendon Disorders of Foot and Ankle Part 2- Achilles

1223

†Address correspondence to Selene G. Parekh, MD, MBA, North Carolina Orthopaedic Clinic, 3609 Southwest Durham Dr, Durham, NC 27707 (e-mail: [email protected]).

No potential conflict of interest declared.

The American Journal of Sports Medicine, Vol. 37, No. 6DOI: 10.1177/0363546509335947© 2009 The Author(s)

Tendon Disorders of the Foot and Ankle, Part 2

Achilles Tendon Disorders

Daniel S. Heckman,* MD, George S. Gluck,* MD, and Selene G. Parekh,†‡§ MD, MBAFrom the *Department of Orthopaedic Surgery, University of North Carolina School of Medicine, Chapel Hill, North Carolina, ‡Division of Orthopaedic Surgery, and §Fuqua Business School, Duke University, Durham, North Carolina

Disorders of the Achilles tendon include both acute and chronic ruptures as well as a spectrum of chronic overuse injuries involving inflammatory and degenerative changes within the tendon and surrounding tissues. These injuries are relatively common in ath-letes as well as among the general population. There is no consensus on the optimal treatment of Achilles tendon disorders. The goals of this review are to develop a current understanding of the anatomy and diagnostic evaluation of the Achilles tendon, and to present current treatment options and the authors’ preferred surgical techniques for operative management of Achilles tendon disorders.

Keywords: Achilles; ruptures; insertional tendinitis; Haglund; bursitis

and soleus muscles.57 As this confluence courses toward its insertion on the calcaneal tuberosity, the fibers twist 90°.19 The Achilles tendon does not have a true synovial sheath; rather, the tendon is encased in a paratenon, which is com-posed of thin membranes that permit gliding of the tendon within the surrounding tissues.61 Anteriorly, the paratenon consists of highly vascularized areolar tissue that provides blood supply to the middle portion of the tendon via a series of mesotenal vincula. Additional blood supply to the tendon arises proximally at the musculotendinous junction and distally at the osseous insertion. A watershed region of decreased vascularity is located approximately 2 to 6 cm proximal to the calcaneal insertion, and is a common location of Achilles tendon lesions.7

Deep and just proximal to the Achilles tendon insertion lies the horseshoe-shaped retrocalcaneal bursa. Posteriorly, this bursa is adherent to the Achilles paratenon, and ante-riorly it abuts the calcaneal tuberosity (Figure 1).57,74 Abnormal prominence of the posterosuperior aspect of the calcaneal tuberosity is termed a Haglund’s deformity, and it is often associated with disorders of the retrocalcaneal bursa and the Achilles tendon insertion.73

ACHILLES TENDON OVERUSE INJURIES

Achilles tendon overuse injuries are a spectrum of disorders that are characterized by inflammatory and degenerative changes within the tendon and surrounding tissues. The lit-erature is confusing regarding the nomenclature and classi-fication of these painful conditions.61 According to Puddu et al,68 noninsertional Achilles tendon lesions include (1) intratendinous degeneration without inflammation

Clinical Sports Medicine Update

Disorders of the Achilles tendon include both acute and chronic ruptures as well as a spectrum of chronic overuse injuries involving inflammatory and degenerative changes within the tendon and surrounding tissues. Achilles ten-don injuries are relatively common among athletes as well as within the general public. They are frequently encoun-tered in individuals who participate in strenuous physical activities that involve repetitive, eccentric impact loading and jumping. Lysholm and Wiklander39 reported a 7% to 9% annual incidence of Achilles disorders in top-level run-ners. In athletes, the most common cause of Achilles ten-don injury is training errors, including a sudden increase in intensity, changes of terrain or surface, changes in train-ing schedules, or use of inappropriate footwear.71 Lower extremity malalignments, such as a hyperpronated or cavus foot and forefoot varus, have also been implicated in Achilles tendon injuries.29,35,56

ANATOMY

The Achilles tendon is the largest and strongest tendon in the body. It is approximately 15 cm in length and is formed from a confluence of fibers derived equally from the gastrocnemius

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 3: Tendon Disorders of Foot and Ankle Part 2- Achilles

1224 Heckman et al The American Journal of Sports Medicine

(tendinosis), (2) inflammation isolated to the Achilles paratenon (paratenonitis), and (3) inflammation of the paratenon with associated intratendinous degeneration (paratenonitis with tendinosis). As these distinct patho-logic entities are frequently coexistent, it is useful to con-sider them as a continuum of disease. Including painful conditions of the Achilles insertion and the retrocalcaneal bursa, the spectrum of Achilles tendon overuse injuries consists of 5 disorders: (1) tendinosis, (2) paratenonitis, (3) paratenonitis with tendinosis, (4) insertional tendinitis, and (5) retrocalcaneal bursitis.71

Tendinosis

Tendinosis is a noninflammatory condition that involves intratendinous degeneration and atrophy due to repetitive microtrauma, aging, or a combination of these. Initially, Achilles tendinosis can be asymptomatic. The presence of pain may signify a partial rupture in the area of degener-ate tendon, leading to focal tenderness, thickening, and nodularity within the tendon. Mobility of the intratendi-nous thickening or nodule with ankle dorsiflexion and plantar flexion (painful arc sign) distinguishes Achilles tendinosis from paratenonitis.11

Ultrasonography visualizes Achilles tendinosis as a hypoechogenic lesion with or without intratendinous calci-fication. The MRI findings with tendinosis include tendon thickening on sagittal images and altered signal appear-ance within the tendon tissue.61

Achilles tendinosis usually responds to nonoperative measures. Rest, a slight heel lift, an Achilles heel pad, activ-ity modifications, and correction of training errors should be part of the initial treatment. Heel lifts or shock-absorbing orthotics can be used to correct associated hindfoot malalign-ment. Older, less active patients and those with severe pain may respond to immobilization in a controlled ankle-motion

(CAM) boot or molded ankle/foot orthosis (MAFO) and physical therapy directed at eccentric heel cord stretching and calf-muscle strengthening. Eccentric heel cord stretches are performed by standing upright with the heel over the edge of a step and the ankle in plantar flexion.2 The Achilles tendon is eccentrically loaded by slowly lowering the heel below the level of the step. Three sets of 15 heel drops are performed twice per day for 12 weeks. Eccentric stretches are intended to be painful, and the patient is instructed to increase exercise intensity to maintain a painful stretch. Mafi et al47 reported that 82% of patients were satisfied and returned to preinjury levels of activity by 12 weeks using this protocol.

Nonsteroidal anti-inflammatory drugs (NSAIDs) and cryotherapy are helpful to reduce any associated inflam-mation. Corticosteroid injections are generally not rec-ommended because of adverse effects on the mechanical properties of the tendon and increased risk of tendon rup-ture. Topical glyceryl trinitrate therapy has been shown to result in significantly decreased Achilles tendon pain with activity and at night, as well as improved functional outcomes in patients with noninsertional Achilles tendi-nopathy. However, painful headaches can limit patient compliance.63 Injection of a sclerosing agent (polidocanol) is thought to decrease pain associated with neovascular-ization in areas of tendinosis. Good short-term results have been reported in a small number of patients.58

The focus of operative treatment for tendinosis is removal of the degenerate tendon. The tendon itself is incised longi-tudinally over the thickened or nodular regions, the foci of degenerative tendon are completely excised, and the resul-tant defect is repaired side to side with absorbable sutures.71 Alternatively, Maffulli et al45 described a percutaneous technique in which 4 or 5 longitudinal tenotomies are made in the symptomatic areas using a No. 11 blade. This percu-taneous technique is indicated for isolated Achilles tendino-sis without paratenonitis and a well-defined nodular lesion that is less than 2.5 cm long. In cases of moderate tendino-sis or partial rupture (20% to 40%), reinforcement of the tendon can be performed using expanded plantaris, or an aponeurosis turn-down flap.54,71 When greater than 50% to 75% of the tendon has been debrided, augmentation of the tendon using autogenous tendon transfer or allograft reconstruction has been recommended.13,71

Perhaps the most popular approach for tendon augmen-tation in severe cases of Achilles tendinosis is transfer of the flexor hallucis longus (FHL) tendon. In this technique, the FHL is harvested through a medial midfoot incision and transferred to the calcaneus anterior to the diseased Achilles tendon.79 Wilcox et al81 reported no reruptures, tendinosis recurrence, or wound complications using this technique in their series of 20 patients with chronic Achilles tendinosis. Using a single-incision modification, Den Hartog13 reported excellent results in 23 of 26 patients treated with FHL augmentation for chronic Achilles tendi-nosis. In this technique, the FHL tendon is harvested from its fibroosseous canal, with the great toe and ankle held in maximal plantar flexion.

Few studies have reported operative management of Achilles tendinosis in athletes, and no criteria explicitly

Figure 1. Posterior ankle anatomy. (Reproduced with per-mission from Stephens MM. Haglund’s deformity and retro-calcaneal bursitis. Orthop Clin North Am. 1994;25(1):41-46.)

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 4: Tendon Disorders of Foot and Ankle Part 2- Achilles

Vol. 37, No. 6, 2009 Achilles Tendon Disorders 1225

guide return to sports. Maffulli et al45 reported a series of 52 middle- and long-distance runners with Achilles tendi-nosis treated with percutaneous tenotomy; 37 (71%) had good or excellent results, with calf strength returning to 90% of normal by 6 months. Schepsis et al72 reported a series of 66 patients, 80% of whom were competitive run-ners. Sixty-seven percent of their patients treated opera-tively for Achilles tendinosis had satisfactory results. Notably, the literature on FHL tendon transfer consists primarily of outcomes in older, less active individuals, and thus there are no return-to-sports guidelines for this procedure when performed in athletes.

Authors’ Preference for Tendinosis. Patients with Achilles tendinosis are initially treated with nonoperative modali-ties, such as NSAIDS, heel lifts, physical therapy, and pos-sibly a period of immobilization if extremely tender. Physical therapy may include eccentric strengthening exercises and other modalities employed by the therapist, including massage therapy, ultrasound, electrical stimula-tion, and iontophoresis.

After 3 to 6 months of failed nonoperative care, patients are offered a surgical debridement. The patient is placed supine on the operating room table. A bump may be placed under the contralateral hip to aid in visualization of the posteromedial ankle. A thigh tourniquet is placed. A longi-tudinal incision is placed 1 cm anterior to the medial bor-der of the Achilles, starting from 1 to 2 cm proximal to the area of concern and extending to the area of the insertion, if needed (Figure 2). The dissection is deepened to the paratenon of the Achilles. Retractors are not placed into the wound until the paratenon is opened, to allow for retraction of full-thickness flaps. The Achilles tendon and the paratenon are inspected. The paratenon is freed from the underlying Achilles and excised. The anterior aspect of the Achilles must be protected to prevent injury to the anterior blood supply. The tendon is inspected for areas of tendinosis. A longitudinal incision is made over the degenerated tendon area. This area is then debrided sharply and repaired with nonabsorbable suture. If less than 50% of the tendon

remains after the debridement, a tendon transfer is required (see later section on FHL tendon transfer). If more than 50% of the tendon remains, the repaired tendon is left as is. The subcutaneous tissues are closed with absorbable suture, and the skin with staples. The extrem-ity is placed in slight plantar flexion in a bulky Jones splint to control edema.

Postoperatively, patients are made nonweightbearing. If a tendon transfer was not performed, staples are removed after 2 weeks and patients are placed in a short-leg weightbearing cast or boot. Range of motion and strength-ening activities are started 4 to 6 weeks after surgery. Return to competitive sports is allowed when strength has been regained, typically by 5 to 6 months postoperatively, although return to play may require up to 1 year with extensive tendon involvement.71

Paratenonitis

Isolated paratenonitis is characterized by inflammation that is limited to the paratenon. It is especially common in middle- and long-distance runners, who report localized pain after strenuous activity. Diffuse tenderness and swell-ing is usually present on both sides of the Achilles tendon, although the medial side is more frequently involved. This is because the medial fibers are subjected to increased stress, due in part to the 90° lateral twist of the Achilles tendon and also to hyperpronation, which is commonly associated with Achilles paratenonitis. Nodularity and crepitus within the paratenon may also be detected on palpation. With acute paratenonitis, the area of swelling and tenderness typically remains fixed with ankle dorsi-flexion.61,71

Ultrasonography and MRI are both useful imaging modalities for Achilles paratenonitis.69 Ultrasonography frequently reveals fluid surrounding the tendon in cases of acute paratenonitis, while peritendinous adhesions that are visualized as thickening of the hypoechogenic paratenon may be seen in cases of chronic paratenonitis. Ultrasonography offers advantages that include short pro-cedure times, avoidance of patient exposure to ionizing radiation, low cost, and excellent evaluation of the internal structure of the tendon. However, ultrasonography is highly user-dependent, and may be unreliable when it fails to detect adhesions or when patients with few adhe-sions have false-negative results.62

Magnetic resonance imaging findings will show a thick-ened paratenon on T1-weighted images and increased sig-nal within the paratenon on T2-weighted images (halo sign).70 Magnetic resonance imaging is less user-dependent and has a wider field of view relative to ultrasonography; furthermore, it offers multiplanar imaging that can pro-vide extensive information on the internal structure of the tendon and surrounding structures. Similar to ultrasonog-raphy, MRI results have demonstrated excellent correla-tion with intraoperative pathologic findings.72

As with tendinosis, the initial nonoperative treatment of symptomatic Achilles paratenonitis is directed toward relieving symptoms, correcting training errors, modifying

Figure 2. Intraoperative photograph demonstrating the pos-teromedial incision used to approach the Achilles tendon.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 5: Tendon Disorders of Foot and Ankle Part 2- Achilles

1226 Heckman et al The American Journal of Sports Medicine

limb malalignment with orthoses, and improving flexibility.61 Interventions such as eccentric heel-cord stretching, topi-cal glyceryl trinitrate therapy, and sclerosing therapy may be helpful; however, these modalities have not been directly studied in patients with paratenonitis. Cryotherapy and NSAIDs are helpful to reduce inflammation. Corticosteroid injections are generally not recommended because of adverse effects on the mechanical properties of the tendon and increased risk of tendon rupture. For chronic or refrac-tory cases of paratenonitis, brisement can be helpful. This technique involves infiltration of saline or dilute local anesthetic into the paratenon sheath to break up adhe-sions.24 Patients who have failed to improve with a 3- to 6-month period of nonoperative management are candi-dates for operative treatment.

Operative treatment is considered for chronic cases that do not respond to nonoperative measures. Open de bridement and lysis of adhesions are performed through a medial longitudinal incision. Full-thickness flaps of skin, subcutaneous tissue, and crural fascia are developed. Thickened portions of the paratenon are excised, taking care to avoid the anterior blood supply. The crural fascia is closed to decrease subcutaneous scarring.34,71 Endoscopic release of the constricting paratenon has also been described, with promising preliminary results.51 The endoscopic technique may reduce early postoperative morbidity; however, there are no studies that compare different operative methods for treatment of Achilles paratenonitis.

Postoperatively, patients begin weightbearing when pain and swelling subside, usually by 7 to 10 days. Strengthening exercises are begun after 2 to 3 weeks, and running is introduced at 6 to 10 weeks. Athletes can expect to return to competition between 3 and 6 months postoperatively.69 Schepsis et al72 reported satisfactory results in 87% of competitive runners treated operatively for Achilles paratenonitis.

Authors’ Preference for Paratenonitis. Achilles parateno-nitis is initially treated nonoperatively, similarly to Achilles tendinosis. After 3 to 6 months of failed nonoperative care, patients are offered a surgical debridement. The surgical technique and postoperative protocol are the same as that previously described for Achilles tendinosis.

Insertional Tendinitis

Insertional tendinitis is a true inflammatory process within the tendinous insertion of the Achilles, and is often diagnosed in obese patients and older or recre-ational athletes.71 It is frequently associated with Haglund’s deformity or retrocalcaneal bursitis, and, in athletes, may be exacerbated by hill running, interval pro-grams, and training errors. Insertional tendinitis presents with morning ankle stiffness, posterior heel pain, and swelling that worsens with activity. Examination reveals tenderness at the bone-tendon interface and limited ankle dorsiflexion.33

Plain radiographs are helpful and should be obtained in patients with insertional tendinitis. Radiographic findings may include a prominence of the posterior calcaneal

tuberosity, possible intratendinous calcification, or a cal-caneal spur (Figure 3).52

Ultrasonography may reveal intratendinous calcifica-tions or heterogeneity in the structure of the tendon fibers. Magnetic resonance imaging may demonstrate increased signal within the retrocalcaneal bursa on T2-weighted images and degenerative or inflammatory changes at the tendon insertion (Figure 4). It can also be used to rule out intratendinous abnormalities within the Achilles tendon.71

Similar to other Achilles overuse injuries, the majority of cases of insertional tendinitis will resolve with nonopera-tive therapy such as rest, cryotherapy, correction of train-ing errors, and correction of malalignment with heel lifts or orthoses. Stretching programs can be helpful; however, eccentric calf-muscle training is reported to be effective in only 32% of cases of insertional tendinitis compared with 89% of cases of noninsertional tendinopathy.15 A modified eccentric calf-muscle training program was recently described in which the patient is instructed to avoid load-ing the Achilles tendon into dorsiflexion. With this modifi-cation, 67% of patients were satisfied and had returned to preinjury activity levels by 12 weeks.26 These preliminary results suggest that moving into dorsiflexion may exacer-bate insertional tendinopathy due to bony impingement on the anterior surface of the Achilles tendon. Sclerosing therapy with polidocanol injections has showed promis-ing results in the treatment of insertional tendinitis.58 McGarvey et al52 found that nonoperative treatment of insertional tendinitis was successful in 89% of cases; there-fore operative intervention should be reserved for cases that have failed exhaustive nonoperative measures for at least 3 to 6 months.

Operative treatment is directed at underlying patho-logic changes, and consists of debridement of the

Figure 3. Lateral ankle radiograph in a patient with Achilles insertional tendinitis demonstrating prominence of the posterior calcaneal tuberosity and intratendinous calcification (arrow).

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 6: Tendon Disorders of Foot and Ankle Part 2- Achilles

Vol. 37, No. 6, 2009 Achilles Tendon Disorders 1227

Achilles insertion, excision of the retrocalcaneal bursa, and posterosuperior calcaneal ostectomy.33 There is clinical and biomechanical evidence that up to 50% of the Achilles insertion can be safely resected without compromising plantar flexion strength or increasing the risk of rupture.6,31 When greater than 50% of the tendon is debrided or detached, repair with suture anchors46,78 or augmentation with plantaris or FHL tendon14 has been recommended.

Numerous surgical approaches have been described, including medial, lateral, posterior, and combined approaches. The central tendon-splitting approach uses an incision 2 cm proximal to the Achilles insertion and extend-ing 6 cm distally.52 The tendon insertion is incised in its midline and any calcific or degenerative areas are excised. If debridement of greater than 50% of the tendon insertion is required, the plantaris tendon is used for augmentation. Excision of the retrocalcaneal bursa and posterosuperior calcaneal ostectomy are performed as necessary. McGarvey et al52 reported an 82% satisfaction rate with this technique in 21 patients. For recalcitrant cases, complete detachment and excision of the diseased Achilles insertion may be nec-essary. This technique is completed with a V-Y lengthening and reattachment of the tendon with suture anchors. Wagner et al78 compared this technique (39 patients) with tendon debridement (22 patients) and reported similar patient satisfaction and complication rates for both proce-dures. However, patients in the detached group required a significantly longer period of immobilization postopera-tively (12 weeks vs 4 weeks). The authors concluded that complete detachment of the Achilles insertion is a reli-able and effective treatment for severe or chronic cases of insertional tendinitis.78

Minimally invasive surgical approaches to the retrocal-caneal space have demonstrated excellent outcomes with reduced morbidity when compared with open techniques. van Dijk et al77 performed endoscopic calcaneoplasty in 20 patients (21 heels) with retrocalcaneal bursitis, and they reported good or excellent results in 19 patients with 2- to 6-year follow-up. Ortmann and McBryde59 reported 3-year follow-up results on 28 patients (30 heels) who underwent endoscopic calcaneoplasty for retrocalcaneal bursitis. Twenty-nine of 30 heels had good or excellent results. One patient had a postoperative Achilles tendon rupture that was repaired through an open procedure. In a prospective cohort study, Leitze et al37 compared the results of endoscopic (33 heels) versus open (17 heels) decompression of the retrocalcaneal space. Postoperative American Orthopaedic Foot and Ankle Society (AOFAS) scores and recovery times were similar between the 2 groups; however, the endoscopic group had shorter opera-tive times and fewer wound complications.

Athletes are allowed to begin gentle training at 6 weeks and full sports activity at 20 to 24 weeks postoperatively, although return to full competition is not recommended before 6 months postoperatively.33 Schepsis et al72 reported 86% satisfactory results in their series of competitive run-ners treated operatively for insertional tendinitis. van Dijk et al77 reported a return to sporting activities at 4 to 8 weeks postoperatively in their series of 20 patients treated with endoscopic debridement.

Retrocalcaneal Bursitis

Unlike paratenonitis and tendinosis, retrocalcaneal bursi-tis is a distinct entity that involves inflammation of the retrocalcaneal bursa and is characterized by pain anterior to the Achilles tendon. The bursa is situated between the calcaneus and the Achilles tendon, and is compressed by these structures during ankle dorsiflexion. Not surpris-ingly, retrocalcaneal bursitis is frequently encountered in runners training uphill. It is also associated with Haglund’s deformity, which can increase mechanical irritation. The bursa can become inflamed, hypertrophied, and adherent to the Achilles tendon, which can lead to degenerative changes within the Achilles insertion.71 Retrocalcaneal bursitis is diagnosed by the 2-finger squeeze test, in which pain is elicited by applying pressure medially and laterally just anterior to the Achilles tendon.

Imaging of retrocalcaneal bursitis, as with insertional tendinitis, should include plain radiographs. Radiographs may demonstrate a prominence of the posterosuperior cal-caneal tuberosity, which can be measured on the lateral radiograph using parallel pitch lines.65 Ultrasonography or MRI can be useful for identifying fluid within the retrocal-caneal bursa as well as associated intratendinous lesions, such as peritendinous thickening, tendinosis, calcification, or partial Achilles tendon tears.71

Nonoperative treatment is effective in about 90% of cases, and consists of rest, NSAIDs, physical therapy, activ-ity modification, and avoidance of shoes with a rigid heel counter.37 Corticosteroid injections administered into the retrocalcaneal bursa should be avoided due to the significant

Figure 4. Sagittal MRI image in a patient with Achilles inser-tional tendinitis demonstrating increased T2-weighted signal within the retrocalcaneal bursa (RB) and at the Achilles ten-don insertion.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 7: Tendon Disorders of Foot and Ankle Part 2- Achilles

1228 Heckman et al The American Journal of Sports Medicine

risk of tendon rupture. For refractory cases, a short period of immobilization in a short-leg walking cast may prove beneficial.25

Those who require operative treatment for retrocalca-neal bursitis are usually either patients with Haglund’s deformity who have failed prolonged nonoperative therapy of 6 months, or long-distance runners who are unwilling to discontinue or modify their training protocol. Operative management focuses on excision of the retrocalcaneal bursa and partial calcaneal ostectomy with debridement of the Achilles tendon if associated insertional tendinitis is present.25 Techniques for endoscopic decompression of the retrocalcaneal space have demonstrated excellent results with shorter operative time, less morbidity, and fewer com-plications when compared with open procedures.37,59,77

Patients with retrocalcaneal bursitis tend to be older, less active individuals; thus very few studies have reported outcomes in athletes. McGarvey et al52 reported an 82% satisfaction rate in 22 patients of various activity levels. Twenty patients resumed daily activities within 3 months; however, only 13 were able to return to sports. Schepsis et al72 reported 75% satisfactory results in their series of competitive runners treated operatively for retrocalcaneal bursitis.

Authors’ Preference for Insertional Tendinitis and Retrocalcaneal Bursitis. Insertional Achilles tendinitis patients are initially treated with nonoperative modali-ties as described previously in the section on Achilles tendinosis. After 3 to 6 months of failed nonoperative care, patients are offered an open surgical debridement. Patient positioning is the same as that described for ten-dinosis. A longitudinal incision is placed 1 cm anterior to the medial border of the Achilles, starting from 2 to 3 cm proximal to the insertion site and extending to the area of the insertion. The dissection is deepened anterior to the Achilles. Retractors are not placed into the wound until the deeper tissues are dissected. The bursa and retrocal-caneal soft tissues are debrided. A 0.25-in or 0.5-in curved osteotome is placed in the interval between the anterior Achilles tendon and the calcaneus. Care must be taken to not retract aggressively and compromise the Achilles insertion site. With the ankle maximally plantar flexed, a microsagittal saw is used to resect the Haglund’s defor-mity. If a large enthesophyte or significant intratendi-nous calcification is present and needs to be resected, a longitudinal incision is made over the enthesophyte and the calcified tissues are resected. There is evidence that up to 50% of the tendon insertion can be detached with-out detriment.6,31 However, as a safeguard, we prefer to secure the tendon with 1 or more suture anchors if greater than 1/3 of its insertion is detached. If more than 50% of the Achilles tendon is debrided, a tendon transfer is performed.

The preferred tendon to transfer, if needed, is the FHL tendon (Figure 5). A single incision technique is used. The incision is extended 2 to 3 cm proximally. Next, the fascia overlying the deep posterior compartment is identified and opened carefully at its midportion. The FHL muscle belly is identified and followed distally until the FHL tendon is visualized. A mosquito forceps is placed deep to

the FHL tendon and is used to place traction on the ten-don. Care is taken to ensure that the FHL tendon, and not the tibial nerve, is retracted. This is performed by pulling on the tendon and confirming flexion of the great toe. Next, the ankle and toe are maximally plantar flexed. This allows for maximal length of the tendon to be pulled through the posterior wound. With the toe and ankle in this position, tenotomy scissors are used to transect the tendon as distally as possible in the fibroosseous tunnel. A 0-Vicryl tag suture (Ethicon Inc, Johnson & Johnson, Somerville, New Jersey) is then placed in the FHL tendon end. The tendon is then measured using the Biotenodesis screw system (Arthex, Naples, Florida). A guide pin is placed just anterior to the Achilles insertion in the calca-neus. A reamer the size of the tendon is used to overdrill the pin. The reamer is taken through the inferior aspect of the calcaneus. With the ankle held in at least 10° of plantarflexion, a Keith needle is used to pull the tag suture and tendon through the calcaneus and the plantar surface of the foot. Traction is placed on the tag suture. A Biotenodesis screw is then placed to hold the FHL tendon in place. Nonabsorbable suture is then used to tack the FHL tendon and muscle belly to the Achilles tendon. The subcutaneous tissues are closed with absorbable suture, and the skin with staples. The extremity is placed in slight plantar flexion in a bulky Jones splint to control edema.

Postoperatively, patients are made nonweightbearing. Two weeks later, staples are removed and patients are placed in a short-leg weightbearing cast or boot. Immobilization con-tinues 4 weeks after the surgery, unless a tendon transfer has been performed. In the FHL tendon-transfer patients, immobilization persists until 6 weeks after surgery. Range of motion and strengthening activities are started 4 to 6 weeks after surgery.

Acute Achilles Tendon Rupture

Acute ruptures of the Achilles tendon usually occur in middle-aged men during athletic activities. Most ruptures occur when pushing off with the weightbearing foot while extending the knee; however, injury can also occur during sudden or violent dorsiflexion of a plantar flexed foot (eccentric contracture). The Achilles tendon most fre-quently ruptures in the watershed region, approximately 2 to 6 cm proximal to its insertion. Degenerative changes are frequently present within spontaneously ruptured Achilles tendons and may lead to rupture under physiologic load-ing.28 Local or systemic corticosteroids48 and fluoroquino-lones53 have also been implicated as risk factors for Achilles ruptures. Patients with Achilles tendon ruptures typically report a sudden, painful pop or snap in their calf with subsequent weakness and gait disturbances. Examination findings include a palpable defect in the ten-don, an increase in passive ankle dorsiflexion, a positive Thompson calf-squeeze test, and decreased plantar-flexion strength.71

In general, MRI and ultrasonography have similar diag-nostic utility in confirming the presence and extent of Achilles tendon ruptures, although ultrasonography is

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 8: Tendon Disorders of Foot and Ankle Part 2- Achilles

Vol. 37, No. 6, 2009 Achilles Tendon Disorders 1229

more cost-effective.43 Ultrasonography reveals a hypoecho-genic area over the ruptured tendon, and MRI visualizes a complete rupture as disruption of the signal within the tendon on T1-weighted images. However, the diagnosis of acute Achilles tendon ruptures is based on clinical exami-nation, and imaging is generally not indicated except in cases of equivocal examination findings.71

Dynamic ultrasonography may have a potential role in selecting appropriate patients for nonoperative treatment based on the degree of tendon apposition. Kotnis et al32 reported a series of 125 Achilles tendon ruptures in which patients with a tendon gap of less than 5 mm on ultra-sonography were treated nonoperatively and patients with a tendon gap of 5 mm or more were treated operatively. There was no difference in rerupture rates between the operative and nonoperative groups. Hufner et al22 reported 74% good to excellent results and a rerupture rate of 6.4%

in a series of 168 Achilles tendon ruptures that were treated nonoperatively based on ultrasonography findings of a tendon gap of less than 10 mm.

There is no consensus on the optimal treatment of acute Achilles tendon ruptures. Favorable functional outcomes have been reported with both operative and nonoperative methods. In general, operative management has demon-strated lower rerupture rates (3.1% vs 13%), greater plantar-flexion strength (87% vs 78%), better return to athletic activities (71% vs 63%), improved ankle motion, and fewer complaints.4,9 However, wound complications (4.7%) and sural nerve injury (6%) are not uncommon after operative repair.4,9

Nonoperative management of Achilles tendon ruptures consists of plaster immobilization in plantar flexion for 4 weeks followed by immobilization in a neutral position for 4 weeks. Early functional rehabilitation using splints

Figure 5. Intraoperative photographs demonstrating transfer of the flexor hallucis longus (FHL) tendon. A, the FHL tendon is identified and retracted with a hemostat. Plantar flexion of the ankle and great toe allows for maximal excursion of the FHL ten-don into the wound. B, the FHL tendon is tagged with a 0-Vicryl suture. C, a guide pin is placed just anterior to the Achilles insertion in the calcaneus. A reamer the size of the tendon is used to overdrill the pin to the inferior aspect of the calcaneus. With the ankle held in at least 10° of plantar flexion, a Keith needle is used to pull the tag suture and tendon through the calcaneus and the plantar surface of the foot. D, the FHL tendon and muscle belly are sutured to the Achilles tendon.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 9: Tendon Disorders of Foot and Ankle Part 2- Achilles

1230 Heckman et al The American Journal of Sports Medicine

or braces with a gradual progression from equinus to a neutral foot position has also been successful.82

Numerous methods of operative treatment of acute Achilles tendon ruptures have been described. Open end-to-end repair with Krackow,80 Bunnell, or Kessler-type sutures placed 2 cm to 4 cm from the ruptured tendon ends can be performed through a medial longitudinal incision under regional or even local anesthesia.8 Ma and Griffith40 first described the percutaneous technique using 3 medial and 3 lateral stab incisions. Many modifications of this method have subsequently been reported,16,17 including endoscopically assisted repair.18,75 Percutaneous techniques offer shorter operative times and lower risk of infection30; however, studies have shown higher rates of rerupture (3.7% vs 2.8%) and sural nerve injury (4.5% vs 2.8%) compared with open repair.12,42,82 Limited open repair techniques have been developed using a 1.5- to 2-cm incision over the tendon gap, which provides visualization and precise control of the tendon ends while avoiding excessive dissection.3,27 A recent bio-mechanical study showed that a limited open repair using the Achillon device (Wright Medical, Memphis, Tennessee) had a significantly higher load to failure than a standard open 4-strand Krackow repair.21 Augmentation of the repair can be performed using fascial turn-down flaps,10 tendon transfers,50,79 or synthetic grafts20,23,60,64; however, these techniques are usually reserved for rerup-tures or neglected cases.

Authors’ Preference for Acute Achilles Rupture. Surgical management for an Achilles tendon rupture is the preferred treatment modality. The patient is placed prone on the oper-ating room table and a thigh tourniquet is placed. A longitu-dinal incision is placed 1 cm anterior to the medial border of the Achilles, starting from 1 cm proximal to the Achilles defect and extending approximately 3 cm. The dissection is deepened to the paratenon of the Achilles. Retractors are not placed into the wound until the paratenon is opened, to allow for retraction of full-thickness flaps. The proximal and distal ends of the Achilles tendon are identified. The Achillon device is placed over the proximal stump of tendon, but deep to the paratenon (Figure 6). The limbs of the device are adjusted to fit the tendon snugly. Three nonabsorbable sutures are then passed sequentially through the limbs of the device and through the proximal tendon using the included needles. The device is then removed, leaving these sutures transversely through the tendon, deep to the paratenon. A similar preparation is performed for the distal stump. With the ankle held at a resting tension matching the uninjured limb, the sutures are then tied with the ten-don ends reapproximated. Next, a running nonabsorbable suture is placed on the medial and lateral ends of the proxi-mal and distal stump, using a Krackow technique. Finally, a running 0-Vicryl suture is placed at the rupture site to fur-ther enhance the repair. The paratenon is then closed with a running 0-Vicryl suture. The subcutaneous tissues are closed with absorbable suture, and the skin with staples. The extremity is placed at resting plantar flexion in a bulky Jones splint to control edema.

Postoperatively, patients are made nonweightbearing. Two weeks later, staples are removed and patients are

placed in a short-leg nonweightbearing cast or boot, at the resting plantar flexion angle. At week 4, the patient is placed in a CAM boot and allowed to bear weight. Range of motion and strengthening activities are started 4 to 6 weeks after surgery.

Athletes may benefit from an accelerated rehabilitation program, as described by Mandelbaum et al.49 Range of motion exercises begin 72 hours postoperatively, ambula-tion in a boot is allowed at 2 weeks, and full weightbearing is allowed at 6 weeks. Ninety-three percent of athletes (27 of 29) treated with this protocol returned to sports by 6 months, with only 2% to 3% deficit of preinjury power and strength.

Chronic Achilles Tendon Rupture

The definition of a chronic Achilles tendon rupture is vari-able, although the most commonly used time frame is 4 to 10 weeks from injury to presentation or treatment.1,67 Up to 20% of Achilles ruptures are missed by the first exam-ining physician.41 Diagnosis of chronic ruptures can be difficult, as pain and swelling may have subsided. Additionally, the palpable defect found in acute injuries may be less apparent or absent in chronic ruptures, as fibrous scar tissue replaces the gap between the proximal and distal ends of the Achilles tendon. Patients with chronic Achilles tendon ruptures will have plantar-flexion weakness and a limp. A positive Thompson squeeze test may not be present as bridging fibrous tissue may allow for a false-negative test. Other examination findings include a positive Matles test, O’Brien needle test, and Copeland sphygmomanometer test.43

Ultrasonography reveals a hypoechogenic area over the tendon gap and thickened irregular edges that correspond to retracted tendon stumps. Chronic ruptures that have filled with scar tissue may appear as an area of increased echogenicity. Magnetic resonance imaging allows determi-nation of the extent of the rupture and evaluation of the condition of the tendon ends. Chronic ruptures appear as decreased signal on T1-weighted images and increased signal on T2-weighted images. Tendon discontinuity, fray-ing, and retraction of the tendon edges are best evaluated with sagittal T2-weighted MRI.43

Nonoperative management of chronic Achilles tendon ruptures is rarely indicated, but may be considered in patients without a functional deficit and in those with a significant risk of wound complications, anesthetic con-traindications for surgery, significant comorbidities, and those who bear minimal weight. Christensen10 reported satisfactory results in 10 of 18 patients (56%) treated nonoperatively with MAFO bracing, and noted that func-tional improvement occurred slowly, sometimes over several years.

Operative management of chronic Achilles tendon rup-tures differs from that of acute ruptures. The tendon ends typically have retracted, making end-to-end repair difficult. Interposing scar tissue must be excised, and the tendon ends must be freshened to promote healing. Furthermore, restora-tion of function depends on reestablishing the resting length and tension of the gastrocnemius-soleus complex.43

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 10: Tendon Disorders of Foot and Ankle Part 2- Achilles

Vol. 37, No. 6, 2009 Achilles Tendon Disorders 1231

Various techniques have been described to reconstruct the Achilles musculotendinous unit. The V-Y tendon advancement typically can make up a deficit of 2 to 3 cm, although this technique has been reported to achieve end-to-end anastomosis for defects as large as 10 cm.1,36 Fascial turn-down flaps with gastrocnemius aponeurosis provide a strong suture line that can accommodate anatomic repair while decreasing the risk of skin adhesions.10 For larger defects where primary anastomosis is not possible, aug-mentation of the repair has been described using fascia lata,5 gracilis,44 or plantaris38 autografts, or Achilles ten-don allograft.55 The use of synthetic grafts has also been described20,23,60,64; although these techniques avoid donor-site morbidity and have provided encouraging results, they have not found widespread use in the orthopaedic community.

Local tissue transfer can be used not only to recon-struct chronic Achilles tendon defects, but also to pro-vide enhanced contractile strength and vascularity to the Achilles musculotendinous complex. The peroneus brevis tendon can be routed through drill holes in the calcaneus66 or through the distal Achilles stump.76 Mann et al50 described a technique using the flexor digitorum longus tendon and postulated that it better reestab-lished the more medial pull of the Achilles while avoid-ing the potential muscle imbalance created by peroneus brevis transfer. The FHL transfer was described by Wapner et al,79 and it provides a long and durable ten-don source, a stronger muscle that fires in phase and along a similar axis as the Achilles tendon, close ana-tomic proximity, and maintenance of the normal muscle balance of the foot.

Figure 6. Intraoperative photographs demonstrating primary repair of an acute Achilles tendon rupture using the Achillon device. A, the Achillon is placed over the proximal stump of tendon and deep to the paratenon. Three nonabsorbable sutures are passed through the device and through the proximal tendon. B, the device is removed, leaving the sutures through the proximal tendon and deep to the paratenon. C, the Achillon is placed over the distal stump of tendon, and sutures are passed in the same fash-ion as the proximal stump sutures. D, with the ankle held at resting tension, the tendon ends are reapproximated and the sutures are tied.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 11: Tendon Disorders of Foot and Ankle Part 2- Achilles

1232 Heckman et al The American Journal of Sports Medicine

Authors’ Preference for Chronic Achilles Rupture. Surgical management for a chronic Achilles tendon rupture is per-formed on patients with symptomatic tears or persistent weakness that inhibits activities. A V-Y advancement and FHL tendon transfer are the preferred methods for treat-ing this condition. A thigh tourniquet is placed. The patient is placed is prone on the operating room table.

For the V-Y advancement, a longitudinal incision is placed 1 cm anterior to the medial border of the Achilles, starting from the musculotendinous junction and extend-ing distally approximately 3 cm distal to the rupture. The dissection is deepened to the paratenon of the Achilles. Retractors are not placed into the wound until the paratenon is opened, to allow for retraction of full-thick-ness flaps. The diseased portion of the tendon is identified and excised. The tendon ends are prepared with nonab-sorbable suture, run in a Krackow pattern. A V-shaped incision is made in the aponeurosis of the musculotendi-nous junction (Figure 7). The apex is proximal, with the limbs of the V 1.5 to 2 times longer than the width of the gap in the Achilles tendon. Traction is placed on the distal aspect of the proximal tendon. Typically, about 2 to 3 cm of advancement can be achieved with this technique, although correction of larger deficits has been reported in the litera-ture. The tendon tension is checked and the tendon ends tied together. The V-Y gap is closed with 0-Vicryl suture. The paratenon is then opposed with a running 0-Vicryl suture. The subcutaneous tissues are closed with absorb-able suture, and the skin with staples. The extremity is placed at resting plantar flexion in a bulky Jones splint to control edema.

Postoperatively, patients are made nonweightbearing. Two weeks later, staples are removed and patients are placed in a short-leg nonweightbearing cast or boot, at the resting plantar flexion angle. At week 4, the patient is placed in a CAM boot and allowed to bear weight. Range of motion activities are started 6 weeks after surgery. Strengthening activities are started 8 weeks after surgery

The FHL tendon transfer technique is described in detail in an earlier section.

SUMMARY

Achilles tendon disorders are frequently encountered in athletes, and are most commonly attributed to training errors. The spectrum of pathologic disorders includes chronic overuse injuries as well as acute and chronic tendon ruptures. Most Achilles overuse injuries respond to nonoperative treatment such as NSAIDS, orthoses, activ-ity modification, correction of training errors, and eccentric calf-muscle strengthening and heel cord–stretching pro-grams. Topical glyceryl trinitrate and sclerosing polido-canol injections are newer modalities with promising short-term results. Corticosteroid injections are not recom-mended for Achilles tendon disorders. Operative treatment of Achilles overuse injuries, consisting primarily of debri-dement of the degenerated or inflamed tissues, is reserved for patients who fail to improve after 3 to 6 months of nonoperative measures. In cases that require extensive

debridement, tendon repair with suture anchors, soft tis-sue augmentation, or FHL tendon transfers may be required, depending on the amount of tissue loss.

Treatment of Achilles tendon ruptures is controversial, and good outcomes have been reported with both operative and nonoperative measures. Dynamic ultrasonography may be useful for identifying patients who are appropriate for nonoperative measures. Numerous surgical techniques have been described for Achilles tendon repair, and there is no consensus on the optimal treatment for these injuries. Chronic Achilles tendon ruptures usually require opera-tive techniques to reconstruct the soft tissue defect. These techniques include soft tissue augmentation, V-Y advance-ment flaps, and local tendon transfers.

Although Achilles tendon injuries are relatively com-mon, there is a paucity of high-quality data regarding the treatment of these disorders, particularly in athletic patient populations. Rehabilitation protocols and return-to-sports guidelines are highly variable and poorly sup-ported in the literature. Further research is needed to define the optimal management of Achilles tendon injuries in athletes.

REFERENCES

1. Abraham E, Pankovich AM, Jansey F. False aneurysm of the profunda femoris artery resulting from intertrochanteric fracture: a case report. J Bone Joint Surg Am. 1975;57(6):871.

2. Alfredson H, Pietila T, Jonsson P, Lorentzon R. Heavy-load eccentric calf muscle training for the treatment of chronic Achilles tendinosis. Am J Sports Med. 1998;26(3):360-366.

3. Assal M, Jung M, Stern R, Rippstein P, Delmi M, Hoffmeyer P. Limited open repair of Achilles tendon ruptures: a technique with a new instrument and findings of a prospective multicenter study. J Bone Joint Surg Am. 2002;84(2):161-170.

Figure 7. Intraoperative photograph demonstrating V-Y adva ncement of a chronic Achilles tendon rupture. The inci-sion is made in the aponeurosis at the musculotendinous junction with the apex directed proximally. The limbs of the V incision are 1.5 to 2 times the length of the Achilles tendon defect.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 12: Tendon Disorders of Foot and Ankle Part 2- Achilles

Vol. 37, No. 6, 2009 Achilles Tendon Disorders 1233

4. Bhandari M, Guyatt GH, Siddiqui F, et al. Treatment of acute Achilles tendon ruptures: a systematic overview and metaanalysis. Clin Orthop Relat Res. 2002;400:190-200.

5. Bugg EI Jr, Boyd BM. Repair of neglected rupture or laceration of the Achilles tendon. Clin Orthop Relat Res. 1968;56:73-75.

6. Calder JD, Saxby TS. Surgical treatment of insertional Achilles tendi-nosis. Foot Ankle Int. 2003;24(2):119-121.

7. Carr AJ, Norris SH. The blood supply of the calcaneal tendon. J Bone Joint Surg Br. 1989;71(1):100-101.

8. Cetti R, Christensen SE. Surgical treatment under local anesthesia of Achilles tendon rupture. Clin Orthop Relat Res. 1983;173:204-208.

9. Cetti R, Christensen SE, Ejsted R, Jensen NM, Jorgensen U. Operative versus nonoperative treatment of Achilles tendon rupture: a prospec-tive randomized study and review of the literature. Am J Sports Med. 1993;21(6):791-799.

10. Christensen I. Rupture of the Achilles tendon: analysis of 57 cases. Acta Chir Scand. 1953;106(1):50-60.

11. Coughlin MJ. Disorders of tendons. In: Coughlin MJ, Mann RA, eds. Surgery of the Foot and Ankle. St. Louis, MO: Mosby; 1999:835-840.

12. Cretnik A, Kosanovic M, Smrkolj V. Percutaneous versus open repair of the ruptured Achilles tendon: a comparative study. Am J Sports Med. 2005;33(9):1369-1379.

13. Den Hartog BD. Flexor hallucis longus transfer for chronic Achilles tendonosis. Foot Ankle Int. 2003;24(3):233-237.

14. DeOrio MJ, Easley ME. Surgical strategies: insertional Achilles tendi-nopathy. Foot Ankle Int. 2008;29(5):542-550.

15. Fahlstrom M, Jonsson P, Lorentzon R, Alfredson H. Chronic Achilles tendon pain treated with eccentric calf-muscle training. Knee Surg Sports Traumatol Arthrosc. 2003;11(5):327-333.

16. FitzGibbons RE, Hefferon J, Hill J. Percutaneous Achilles tendon repair. Am J Sports Med. 1993;21(5):724-727.

17. Gorschewsky O, Pitzl M, Putz A, Klakow A, Neumann W. Percutaneous repair of acute Achilles tendon rupture. Foot Ankle Int. 2004;25(4):219-224.

18. Halasi T, Tallay A, Berkes I. Percutaneous Achilles tendon repair with and without endoscopic control. Knee Surg Sports Traumatol Arthrosc. 2003;11(6):409-414.

19. Hoffman B, Nunley J. Achilles tendon torsion has no effect on percu-taneous triple-cut tenotomy results. Foot Ankle Int. 2006;27(11): 960-964.

20. Howard CB, Winston I, Bell W, Mackie I, Jenkins DH. Late repair of the calcaneal tendon with carbon fibre. J Bone Joint Surg Br. 1984;66(2):206-208.

21. Huffard B, O’Loughlin PF, Wright T, Deland J, Kennedy JG. Achilles tendon repair: Achillon system vs. Krackow suture: an anatomic in vitro biomechanical study. Clin Biomech (Bristol, Avon). 2008;23(9): 1158-1164.

22. Hufner TM, Brandes DB, Thermann H, Richter M, Knobloch K, Krettek C. Long-term results after functional nonoperative treatment of Achilles tendon rupture. Foot Ankle Int. 2006;27(3):167-171.

23. Jennings AG, Sefton GK. Chronic rupture of tendo Achillis: long-term results of operative management using polyester tape. J Bone Joint Surg Br. 2002;84(3):361-363.

24. Johnston E, Scranton P Jr, Pfeffer GB. Chronic disorders of the Achilles tendon: results of conservative and surgical treatments. Foot Ankle Int. 1997;18(9):570-574.

25. Jones DC, James SL. Partial calcaneal ostectomy for retrocalcaneal bursitis. Am J Sports Med. 1984;12(1):72-73.

26. Jonsson P, Alfredson H, Sunding K, Fahlstrom M, Cook J. New regi-men for eccentric calf-muscle training in patients with chronic inser-tional Achilles tendinopathy: results of a pilot study. Br J Sports Med. 2008;42(9):746-749.

27. Kakiuchi M. A combined open and percutaneous technique for repair of tendo Achillis: comparison with open repair. J Bone Joint Surg Br. 1995;77(1):60-63.

28. Kannus P, Jozsa L. Histopathological changes preceding spontane-ous rupture of a tendon: a controlled study of 891 patients. J Bone Joint Surg Am. 1991;73(10):1507-1525.

29. Kaufman KR, Brodine SK, Shaffer RA, Johnson CW, Cullison TR. The effect of foot structure and range of motion on musculoskeletal over-use injuries. Am J Sports Med. 1999;27(5):585-593.

30. Khan RJ, Fick D, Brammar TJ, Crawford J, Parker MJ. Interventions for treating acute Achilles tendon ruptures. Cochrane Database Syst Rev. 2004;3:CD003674.

31. Kolodziej P, Glisson RR, Nunley JA. Risk of avulsion of the Achilles tendon after partial excision for treatment of insertional tendonitis and Haglund’s deformity: a biomechanical study. Foot Ankle Int. 1999;20(7):433-437.

32. Kotnis R, David S, Handley R, Willett K, Ostlere S. Dynamic ultra-sound as a selection tool for reducing Achilles tendon reruptures. Am J Sports Med. 2006;34(9):1395-1400.

33. Krishna Sayana M, Maffulli N. Insertional Achilles tendinopathy. Foot Ankle Clin. 2005;10(2):309-320.

34. Kvist H, Kvist M. The operative treatment of chronic calcaneal paratenonitis. J Bone Joint Surg Br. 1980;62(3):353-357.

35. Kvist M. Achilles tendon injuries in athletes. Sports Med. 1994;18(3): 173-201.

36. Leitner A, Voigt C, Rahmanzadeh R. Treatment of extensive aseptic defects in old Achilles tendon ruptures: methods and case reports. Foot Ankle. 1992;13(4):176-180.

37. Leitze Z, Sella EJ, Aversa JM. Endoscopic decompression of the retrocalcaneal space. J Bone Joint Surg Am. 2003;85(8):1488-1496.

38. Lynn TA. Repair of the torn Achilles tendon, using the plantaris ten-don as a reinforcing membrane. J Bone Joint Surg Am. 1966;48(2): 268-272.

39. Lysholm J, Wiklander J. Injuries in runners. Am J Sports Med. 1987;15(2):168-171.

40. Ma GW, Griffith TG. Percutaneous repair of acute closed ruptured Achilles tendon: a new technique. Clin Orthop Relat Res. 1977; 128:247-255.

41. Maffulli N. Clinical tests in sports medicine: more on Achilles tendon. Br J Sports Med. 1996;30(3):250.

42. Maffulli N. Rupture of the Achilles tendon. J Bone Joint Surg Am. 1999;81(7):1019-1036.

43. Maffulli N, Ajis A, Longo UG, Denaro V. Chronic rupture of tendo Achillis. Foot Ankle Clin. 2007;12(4):583-596.

44. Maffulli N, Leadbetter WB. Free gracilis tendon graft in neglected tears of the Achilles tendon. Clin J Sport Med. 2005;15(2):56-61.

45. Maffulli N, Testa V, Capasso G, Bifulco G, Binfield PM. Results of per-cutaneous longitudinal tenotomy for Achilles tendinopathy in middle- and long-distance runners. Am J Sports Med. 1997;25(6):835-840.

46. Maffulli N, Testa V, Capasso G, Sullo A. Calcific insertional Achilles tendinopathy. Reattachment with bone anchors. Am J Sports Med. 2004;32(1):174-182.

47. Mafi N, Lorentzon R, Alfredson H. Superior short-term results with eccentric calf muscle training compared to concentric training in a randomized prospective multicenter study on patients with chronic Achilles tendinosis. Knee Surg Sports Traumatol Arthrosc. 2001;9(1):42-47.

48. Mahler F, Fritschy D. Partial and complete ruptures of the Achilles tendon and local corticosteroid injections. Br J Sports Med. 1992;26(1):7-14.

49. Mandelbaum BR, Myerson MS, Forster R. Achilles tendon ruptures: a new method of repair, early range of motion, and functional rehabilita-tion. Am J Sports Med. 1995;23(4):392-395.

50. Mann RA, Holmes GB Jr, Seale KS, Collins DN. Chronic rupture of the Achilles tendon: a new technique of repair. J Bone Joint Surg Am. 1991;73(2):214-219.

51. Maquirriain J, Ayerza M, Costa-Paz M, Muscolo DL. Endoscopic surgery in chronic Achilles tendinopathies: a preliminary report. Arthroscopy. 2002;18(3):298-303.

52. McGarvey WC, Palumbo RC, Baxter DE, Leibman BD. Insertional Achilles tendinosis: surgical treatment through a central tendon split-ting approach. Foot Ankle Int. 2002;23(1):19-25.

53. McGarvey WC, Singh D, Trevino SG. Partial Achilles tendon ruptures associated with fluoroquinolone antibiotics: a case report and litera-ture review. Foot Ankle Int. 1996;17(8):496-498.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from

Page 13: Tendon Disorders of Foot and Ankle Part 2- Achilles

1234 Heckman et al The American Journal of Sports Medicine

54. Nelen G, Martens M, Burssens A. Surgical treatment of chronic Achilles tendinitis. Am J Sports Med. 1989;17(6):754-759.

55. Nellas ZJ, Loder BG, Wertheimer SJ. Reconstruction of an Achilles tendon defect utilizing an Achilles tendon allograft. J Foot Ankle Surg. 1996;35(2):144-148; discussion 190.

56. Nigg BM. The role of impact forces and foot pronation: a new para-digm. Clin J Sport Med. 2001;11(1):2-9.

57. O’Brien M. The anatomy of the Achilles tendon. Foot Ankle Clin. 2005;10(2):225-238.

58. Ohberg L, Alfredson H. Sclerosing therapy in chronic Achilles tendon insertional pain-results of a pilot study. Knee Surg Sports Traumatol Arthrosc. 2003;11(5):339-343.

59. Ortmann FW, McBryde AM. Endoscopic bony and soft-tissue decompression of the retrocalcaneal space for the treatment of Haglund deformity and retrocalcaneal bursitis. Foot Ankle Int. 2007;28(2):149-153.

60. Ozaki J, Fujiki J, Sugimoto K, Tamai S, Masuhara K. Reconstruction of neglected Achilles tendon rupture with Marlex mesh. Clin Orthop Relat Res. 1989;238:204-208.

61. Paavola M, Kannus P, Jarvinen TA, Khan K, Jozsa L, Jarvinen M. Achilles tendinopathy. J Bone Joint Surg Am. 2002;84(11):2062-2076.

62. Paavola M, Paakkala T, Kannus P, Jarvinen M. Ultrasonography in the differential diagnosis of Achilles tendon injuries and related disorders: a comparison between pre-operative ultrasonography and surgical findings. Acta Radiol. 1998;39(6):612-619.

63. Paoloni JA, Appleyard RC, Nelson J, Murrell GA. Topical glyceryl trinitrate treatment of chronic noninsertional achilles tendinopathy: a randomized, double-blind, placebo-controlled trial. J Bone Joint Surg Am. 2004;86(5):916-922.

64. Parsons JR, Rosario A, Weiss AB, Alexander H. Achilles tendon repair with an absorbable polymer-carbon fiber composite. Foot Ankle. 1984;5(2):49-53.

65. Pavlov H, Heneghan MA, Hersh A, Goldman AB, Vigorita V. The Haglund syndrome: initial and differential diagnosis. Radiology. 1982;144(1):83-88.

66. Perez Teuffer A. Traumatic rupture of the Achilles tendon: reconstruc-tion by transplant and graft using the lateral peroneus brevis. Orthop Clin North Am. 1974;5(1):89-93.

67. Porter DA, Mannarino FP, Snead D, Gabel SJ, Ostrowski M. Primary repair without augmentation for early neglected Achilles tendon rup-tures in the recreational athlete. Foot Ankle Int. 1997;18(9):557-564.

68. Puddu G, Ippolito E, Postacchini F. A classification of Achilles tendon disease. Am J Sports Med. 1976;4(4):145-150.

69. Saltzman CL, Tearse DS. Achilles tendon injuries. J Am Acad Orthop Surg. 1998;6(5):316-325.

70. Saxena A, Cheung S. Surgery for chronic Achilles tendinopathy: review of 91 procedures over 10 years. J Am Podiatr Med Assoc. 2003;93(4):283-291.

71. Schepsis AA, Jones H, Haas AL. Achilles tendon disorders in ath-letes. Am J Sports Med. 2002;30(2):287-305.

72. Schepsis AA, Wagner C, Leach RE. Surgical management of Achilles tendon overuse injuries: a long-term follow-up study. Am J Sports Med. 1994;22(5):611-619.

73. Sella EJ, Caminear DS, McLarney EA. Haglund’s syndrome. J Foot Ankle Surg. 1998;37(2):110-114; discussion 173.

74. Stephens MM. Haglund’s deformity and retrocalcaneal bursitis. Orthop Clin North Am. 1994;25(1):41-46.

75. Thermann H, Tibesku CO, Mastrokalos DS, Passler HH. Endoscopically assisted percutaneous Achilles tendon suture. Foot Ankle Int. 2001;22(2):158-160.

76. Turco V, Spinella AJ. Team physician #2: peroneus brevis transfer for Achilles tendon rupture in athletes. Orthop Rev. 1988;17(8):822-824, 827-828.

77. van Dijk CN, van Dyk GE, Scholten PE, Kort NP. Endoscopic calca-neoplasty. Am J Sports Med. 2001;29(2):185-189.

78. Wagner E, Gould JS, Kneidel M, Fleisig GS, Fowler R. Technique and results of Achilles tendon detachment and reconstruction for insertional Achilles tendinosis. Foot Ankle Int. 2006;27(9): 677-684.

79. Wapner KL, Pavlock GS, Hecht PJ, Naselli F, Walther R. Repair of chronic Achilles tendon rupture with flexor hallucis longus tendon transfer. Foot Ankle. 1993;14(8):443-449.

80. Watson TW, Jurist KA, Yang KH, Shen KL. The strength of Achilles tendon repair: an in vitro study of the biomechanical behavior in human cadaver tendons. Foot Ankle Int. 1995;16(4): 191-195.

81. Wilcox DK, Bohay DR, Anderson JG. Treatment of chronic Achilles ten-don disorders with flexor hallucis longus tendon transfer/augmentation. Foot Ankle Int. 2000;21(12):1004-1010.

82. Wong J, Barrass V, Maffulli N. Quantitative review of operative and nonoperative management of Achilles tendon ruptures. Am J Sports Med. 2002;30(4):565-575.

For reprints and permission queries, please visit SAGE’s Web site at http://www.sagepub.com/journalsPermissions.nav.

at UNIV OF OTTAWA LIBRARY on December 22, 2010ajs.sagepub.comDownloaded from